[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (385)

Search Parameters:
Keywords = Drp-1

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 1679 KiB  
Article
Omega-3 Fatty Acids Modify Drp1 Expression and Activate the PINK1-Dependent Mitophagy Pathway in the Kidney and Heart of Adenine-Induced Uremic Rats
by Dong Ho Choi, Su Mi Lee, Bin Na Park, Mi Hwa Lee, Dong Eun Yang, Young Ki Son, Seong Eun Kim and Won Suk An
Biomedicines 2024, 12(9), 2107; https://doi.org/10.3390/biomedicines12092107 (registering DOI) - 15 Sep 2024
Abstract
Mitochondrial homeostasis is controlled by biogenesis, dynamics, and mitophagy. Mitochondrial dysfunction plays a central role in cardiovascular and renal disease and omega-3 fatty acids (FAs) are beneficial for cardiovascular disease. We investigated whether omega-3 fatty acids (FAs) regulate mitochondrial biogenesis, dynamics, and mitophagy [...] Read more.
Mitochondrial homeostasis is controlled by biogenesis, dynamics, and mitophagy. Mitochondrial dysfunction plays a central role in cardiovascular and renal disease and omega-3 fatty acids (FAs) are beneficial for cardiovascular disease. We investigated whether omega-3 fatty acids (FAs) regulate mitochondrial biogenesis, dynamics, and mitophagy in the kidney and heart of adenine-induced uremic rats. Eighteen male Sprague Dawley rats were divided into normal control, adenine control, and adenine with omega-3 FA groups. Using Western blot analysis, the kidney and heart expression of mitochondrial homeostasis-related molecules, including peroxisome proliferator-activated receptor gamma coactivator-1 alpha (PGC-1α), dynamin-related protein 1 (Drp1), and phosphatase and tensin homolog-induced putative kinase 1 (PINK1) were investigated. Compared to normal, serum creatinine and heart weight/body weight in adenine control were increased and slightly improved in the omega-3 FA group. Compared to the normal controls, the expression of PGC-1α and PINK1 in the kidney and heart of the adenine group was downregulated, which was reversed after omega-3 FA supplementation. Drp1 was upregulated in the kidney but downregulated in the heart in the adenine group. Drp1 expression in the heart recovered in the omega-3 FA group. Mitochondrial DNA (mtDNA) was decreased in the kidney and heart of the adenine control group but the mtDNA of the heart was recovered in the omega-3 FA group. Drp1, which is related to mitochondrial fission, may function oppositely in the uremic kidney and heart. Omega-3 FAs may be beneficial for mitochondrial homeostasis by activating mitochondrial biogenesis and PINK1-dependent mitophagy in the kidney and heart of uremic rats. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in the expression of factors related to mitochondrial biogenesis including PGC-1α, SIRT1/3, and Nrf2 in the kidney (<b>A</b>) and heart (<b>B</b>). * <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the control group). <sup>a</sup> <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the adenine group).</p>
Full article ">Figure 2
<p>Changes in the expression of factors related to mitochondrial fusion and fission including OPA1, Drp1, and Mfn1/2 in the kidney (<b>A</b>) and heart (<b>B</b>). * <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from normal control group). <sup>a</sup> <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the adenine group).</p>
Full article ">Figure 3
<p>Changes in the expression of factors related to mitochondrial mitophagy including PINK1, BNIP3, and NIX in the kidney (<b>A</b>) and heart (<b>B</b>). * <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the control group). <sup>a</sup> <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the adenine group).</p>
Full article ">Figure 4
<p>Relative mitochondrial DNA (mtDNA) content in the kidney (<b>A</b>) and heart (<b>B</b>). * <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the control group). <sup>a</sup> <span class="html-italic">p</span> value &lt; 0.05 (mean values are significantly different from the adenine group).</p>
Full article ">
16 pages, 4530 KiB  
Article
Bcl-2 Orthologues, Buffy and Debcl, Can Suppress Drp1-Dependent Age-Related Phenotypes in Drosophila
by Azra Hasan and Brian E. Staveley
Biomolecules 2024, 14(9), 1089; https://doi.org/10.3390/biom14091089 - 30 Aug 2024
Viewed by 319
Abstract
The relationship of Amyotrophic Lateral Sclerosis, Parkinson’s disease, and other age-related neurodegenerative diseases with mitochondrial dysfunction has led to our study of the mitochondrial fission gene Drp1 in Drosophila melanogaster and aspects of aging. Previously, the Drp1 protein has been demonstrated to interact [...] Read more.
The relationship of Amyotrophic Lateral Sclerosis, Parkinson’s disease, and other age-related neurodegenerative diseases with mitochondrial dysfunction has led to our study of the mitochondrial fission gene Drp1 in Drosophila melanogaster and aspects of aging. Previously, the Drp1 protein has been demonstrated to interact with the Drosophila Bcl-2 mitochondrial proteins, and Drp1 mutations can lead to mitochondrial dysfunction and neuronal loss. In this study, the Dopa decarboxylase-Gal4 (Ddc-Gal4) transgene was exploited to direct the expression of Drp1 and Drp1-RNAi transgenes in select neurons. Here, the knockdown of Drp1 seems to compromise locomotor function throughout life but does not alter longevity. The co-expression of Buffy suppresses the poor climbing induced by the knockdown of the Drp1 function. The consequences of Drp1 overexpression, which specifically reduced median lifespan and diminished climbing abilities over time, can be suppressed through the directed co-overexpression of pro-survival Bcl-2 gene Buffy or by the co-knockdown of the pro-cell death Bcl-2 homologue Debcl. Alteration of the expression of Drp1 acts to phenocopy neurodegenerative disease phenotypes in Drosophila, while overexpression of Buffy can counteract or rescue these phenotypes to improve overall health. The diminished healthy aging due to either the overexpression of Drp1 or the RNA interference of Drp1 has produced novel Drosophila models for investigating mechanisms underlying neurodegenerative disease. Full article
(This article belongs to the Special Issue Molecular Advances in Mechanism and Regulation of Lifespan and Aging)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Drp1</span> is evolutionarily conserved between Drosophila and humans. (<b>a</b>) Clustal Omega multiple sequence alignment of <span class="html-italic">D. melanogaster</span> Drp1 (NP_608694.2) protein with the <span class="html-italic">H. sapiens</span> (NP_001265392.1) shows evolutionarily conserved domains identified using the NCBI Conserved Domain Database (CDD) and is further confirmed by the Eukaryotic Linear Motif (ELM) resource. The two well-documented phosphorylation sites are identified: S606 and S627 in dynamin-1-like protein (DLP-1) isoform 4 of <span class="html-italic">H. sapiens</span> and S616 and T637 in Drp1 of <span class="html-italic">D. melanogaster</span>. The asterisks indicate the residues that are identical, the colons indicate the conserved substitutions, and the dots indicates the semi-conserved substitutions. Colour differences indicate the chemical nature of amino acids: red indicates small hydrophobic (includes aromatic) residues; blue indicates acidic; magenta indicates basic; and green indicates basic with hydroxyl or amine groups. (<b>bi</b>) The original Dynamin-1-like protein (DLP-1) structure of H. sapiens (NP_001265392.1) from the NCBI structure database. (<b>bii</b>) The Phyre2 web portal for protein modelling, prediction, and analysis mediated the development of a model of the Drp1 protein of <span class="html-italic">D. melanogaster</span> (NP_608694.2) from a 76% identical protein with a confidence of 100%. The N terminus is coloured in Magenta; C terminus is coloured in Red, and a consensus ATG8 binding region at N terminus is coloured in orange.</p>
Full article ">Figure 1 Cont.
<p><span class="html-italic">Drp1</span> is evolutionarily conserved between Drosophila and humans. (<b>a</b>) Clustal Omega multiple sequence alignment of <span class="html-italic">D. melanogaster</span> Drp1 (NP_608694.2) protein with the <span class="html-italic">H. sapiens</span> (NP_001265392.1) shows evolutionarily conserved domains identified using the NCBI Conserved Domain Database (CDD) and is further confirmed by the Eukaryotic Linear Motif (ELM) resource. The two well-documented phosphorylation sites are identified: S606 and S627 in dynamin-1-like protein (DLP-1) isoform 4 of <span class="html-italic">H. sapiens</span> and S616 and T637 in Drp1 of <span class="html-italic">D. melanogaster</span>. The asterisks indicate the residues that are identical, the colons indicate the conserved substitutions, and the dots indicates the semi-conserved substitutions. Colour differences indicate the chemical nature of amino acids: red indicates small hydrophobic (includes aromatic) residues; blue indicates acidic; magenta indicates basic; and green indicates basic with hydroxyl or amine groups. (<b>bi</b>) The original Dynamin-1-like protein (DLP-1) structure of H. sapiens (NP_001265392.1) from the NCBI structure database. (<b>bii</b>) The Phyre2 web portal for protein modelling, prediction, and analysis mediated the development of a model of the Drp1 protein of <span class="html-italic">D. melanogaster</span> (NP_608694.2) from a 76% identical protein with a confidence of 100%. The N terminus is coloured in Magenta; C terminus is coloured in Red, and a consensus ATG8 binding region at N terminus is coloured in orange.</p>
Full article ">Figure 2
<p>Altered <span class="html-italic">Drp1</span> expression under the control of <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> influences the survival and climbing ability of flies. (<b>A</b>) The GraphPad prism8 generated graph of the longevity assay for the expression of <span class="html-italic">Drp1</span>, <span class="html-italic">Drp1-RNAis</span> under the control of <span class="html-italic">Ddc-Gal4</span> transgene. The directed expression results in decreased median lifespan of 56 days compared to 68 days of control, as calculated by Log-rank Mantel–Cox test, with Bonferroni correction. The knockdown of <span class="html-italic">Drp1</span> under the control of the <span class="html-italic">Ddc-Gal4</span> transgene results in lifespan of 70 days with <span class="html-italic">UAS-Drp1-RNAi1</span> and 72 days with <span class="html-italic">UAS-Drp-RNAi2</span> compared to 68 days of control performed by Log-rank Mantel–Cox test, with Bonferroni correction. (<b>B</b>) The GraphPad prism8 generated graph of the climbing abilities of flies with overexpression of <span class="html-italic">Drp1</span>, the <span class="html-italic">Drp1-RNAis</span>, and the <span class="html-italic">lacZ</span> control. The climbing ability of <span class="html-italic">Drp1</span> overexpression and <span class="html-italic">Drp1</span>-<span class="html-italic">RNAi</span>s flies have decreased compared to control as determined in non-linear fitting of the climbing curve by 95% confidence interval (<span class="html-italic">p</span>-value &lt; 0.0001).</p>
Full article ">Figure 3
<p>The ectopic expression of <span class="html-italic">Drp1-RNAi</span>, directed by <span class="html-italic">Ddc-Gal4<sup>4.</sup></span><sup>3D</sup>, can increase median lifespan and decrease climbing. (<b>A</b>) In control, <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-lacZ</span> critical class males resulted in a median life span of 62 days (n = 308). Expression of <span class="html-italic">Drp1</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> resulted in a median life span of 56 days (n = 310), much lower than the <span class="html-italic">lacZ</span>-expressing control; expression of <span class="html-italic">Drp1</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1-RNAi</span> resulted in a median life span of 64 days (n = 250) very similar to control (<span class="html-italic">Ddc/lacZ</span>) as determined by the Log-rank Mantel–Cox test (<span class="html-italic">p</span> value = 0.0633) with Bonferroni correction. The graph of the longevity assay was generated by GraphPad prism8. (<b>B</b>) The <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> flies express <span class="html-italic">UAS-lacZ</span> in control flies. The climbing abilities of <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> expressing flies have decreased compared to control as determined in the non-linear fitting of the climbing curve by a 95% confidence interval (<span class="html-italic">p</span> &lt; 0.0001). The flies’ climbing ability expressing <span class="html-italic">Drp1</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1-RNAi</span> transgene is similar to control as determined in the non-linear fitting of the climbing curve by a 95% confidence interval at <span class="html-italic">p</span> value = 1.309. The graph of longevity assay was generated by GraphPad prism8 non-linear regression curve. (<b>C</b>) In control, <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-lacZ</span> critical class males resulted in a median life span of 62 days (n = 308). Knockdown of <span class="html-italic">Drp1-RNAi</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> resulted in a median life span of 70 days (n = 321), much higher compared to the control; knockdown of <span class="html-italic">Drp1-RNAi</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> resulted in a median life span of 64 days (n = 327), very similar to control (<span class="html-italic">Ddc/lacZ</span>) as determined by the Log-rank Mantel–Cox test (<span class="html-italic">p</span> value = 0.0582) with Bonferroni correction. The graph of the longevity assay was generated by GraphPad prism8. (<b>D</b>) The <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> flies express <span class="html-italic">UAS-lacZ</span> in control flies. The climbing abilities of <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1-RNAi</span> expressing flies have decreased compared to control as determined in the non-linear fitting of the climbing curve by a 95% confidence interval (<span class="html-italic">p</span> &lt; 0.0001). The flies’ climbing ability expressing <span class="html-italic">Drp1-RNAi</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> transgene is similar to control (<span class="html-italic">Ddc/lacZ</span>) as determined in the non-linear fitting of the climbing curve by a 95% confidence interval at <span class="html-italic">p</span> value = 0.0027. The graph of longevity assay was generated by GraphPad prism8 non-linear regression curve.</p>
Full article ">Figure 3 Cont.
<p>The ectopic expression of <span class="html-italic">Drp1-RNAi</span>, directed by <span class="html-italic">Ddc-Gal4<sup>4.</sup></span><sup>3D</sup>, can increase median lifespan and decrease climbing. (<b>A</b>) In control, <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-lacZ</span> critical class males resulted in a median life span of 62 days (n = 308). Expression of <span class="html-italic">Drp1</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> resulted in a median life span of 56 days (n = 310), much lower than the <span class="html-italic">lacZ</span>-expressing control; expression of <span class="html-italic">Drp1</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1-RNAi</span> resulted in a median life span of 64 days (n = 250) very similar to control (<span class="html-italic">Ddc/lacZ</span>) as determined by the Log-rank Mantel–Cox test (<span class="html-italic">p</span> value = 0.0633) with Bonferroni correction. The graph of the longevity assay was generated by GraphPad prism8. (<b>B</b>) The <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> flies express <span class="html-italic">UAS-lacZ</span> in control flies. The climbing abilities of <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> expressing flies have decreased compared to control as determined in the non-linear fitting of the climbing curve by a 95% confidence interval (<span class="html-italic">p</span> &lt; 0.0001). The flies’ climbing ability expressing <span class="html-italic">Drp1</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1-RNAi</span> transgene is similar to control as determined in the non-linear fitting of the climbing curve by a 95% confidence interval at <span class="html-italic">p</span> value = 1.309. The graph of longevity assay was generated by GraphPad prism8 non-linear regression curve. (<b>C</b>) In control, <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-lacZ</span> critical class males resulted in a median life span of 62 days (n = 308). Knockdown of <span class="html-italic">Drp1-RNAi</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> resulted in a median life span of 70 days (n = 321), much higher compared to the control; knockdown of <span class="html-italic">Drp1-RNAi</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> resulted in a median life span of 64 days (n = 327), very similar to control (<span class="html-italic">Ddc/lacZ</span>) as determined by the Log-rank Mantel–Cox test (<span class="html-italic">p</span> value = 0.0582) with Bonferroni correction. The graph of the longevity assay was generated by GraphPad prism8. (<b>D</b>) The <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span> flies express <span class="html-italic">UAS-lacZ</span> in control flies. The climbing abilities of <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1-RNAi</span> expressing flies have decreased compared to control as determined in the non-linear fitting of the climbing curve by a 95% confidence interval (<span class="html-italic">p</span> &lt; 0.0001). The flies’ climbing ability expressing <span class="html-italic">Drp1-RNAi</span> in <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> transgene is similar to control (<span class="html-italic">Ddc/lacZ</span>) as determined in the non-linear fitting of the climbing curve by a 95% confidence interval at <span class="html-italic">p</span> value = 0.0027. The graph of longevity assay was generated by GraphPad prism8 non-linear regression curve.</p>
Full article ">Figure 4
<p>Altered expression of <span class="html-italic">Buffy</span> and <span class="html-italic">Debcl</span> can enhance and suppress climbing ability in <span class="html-italic">Drp1</span> over-expression flies. (<b>A</b>) In control, <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span>; <span class="html-italic">UAS-Drp1 UAS-lacZ</span> critical class males resulted in a median life span of 58 days (n = 282). The overexpression of <span class="html-italic">Buffy</span> results in a median lifespan of 68 days (n = 375) compares to 58 days of control (<span class="html-italic">p</span> value = 0.0002); the knockdown of <span class="html-italic">Buffy</span> directed by the <span class="html-italic">Ddc-Gal4<sup>4.3D</sup> UAS-Drp1</span> transgene results in the median lifespan of 52 (n = 274), much less compared to control, determined by Log-rank Mantel–Cox test at <span class="html-italic">p</span>-value &lt; 0.0001, with Bonferroni correction. The overexpression of <span class="html-italic">Debcl<sup>EY05743</sup></span> results in a median lifespan of 60 days (n = 331) similar to 58 days of control determined by Log-rank Mantel–Cox test at <span class="html-italic">p</span>-value 0.3293; the inhibition of <span class="html-italic">Debcl</span> directed by the <span class="html-italic">Ddc-Gal4 UAS-Drp1</span> transgene result in the median lifespan of 66 (n = 303); much higher than control, determined by Log-rank Mantel–Cox test at <span class="html-italic">p</span> value 0.0057, with Bonferroni correction. (<b>B</b>) The GraphPad prism8 generated graph of the climbing abilities of <span class="html-italic">Ddc-Gal4 UAS-Drp1</span> flies with the expression of <span class="html-italic">Buffy</span>, <span class="html-italic">Buffy-RNAi</span>, <span class="html-italic">Debcl<sup>EY05743</sup></span>, <span class="html-italic">Debcl-RNAi<sup>v47515</sup></span> and control. The climbing abilities of flies overexpressing <span class="html-italic">Buffy</span> have rescued compared to control as determined in the climbing curve’s non-linear fitting by a 95% confidence interval (<span class="html-italic">p</span> &lt; 0.0001). The climbing ability of the flies was further weakened by the knockdown of <span class="html-italic">UAS-Buffy-RNAi</span> as determined in the non-linear fitting of the climbing curve by a 95% confidence interval at a <span class="html-italic">p</span>-value 0.0125 and 0.03293, respectively (n = 50). The climbing abilities of flies expressing <span class="html-italic">Debcl-RNAi<sup>v47515</sup></span> has been rescued compared to control as determined by the non-linear fitting of the climbing curve by a 95% confidence interval (<span class="html-italic">p</span> value = 0.0057). The graph of longevity assay was generated by GraphPad prism8 non-linear regression curve.</p>
Full article ">Figure 5
<p>Altered expression of <span class="html-italic">Buffy</span> and <span class="html-italic">Debcl</span> can enhance and suppress climbing ability in <span class="html-italic">Drp1</span> knockdown flies. (<b>A</b>) In control, <span class="html-italic">Ddc-Gal4<sup>4.3D</sup></span>; <span class="html-italic">UAS-Drp1</span> transgene results in the median lifespan of 62 (n = 273), determined by Log-rank Mantel–Cox test at <span class="html-italic">p</span>-value &lt; 0.0001, with Bonferroni correction. The overexpression of <span class="html-italic">Debcl<sup>EY05743</sup></span> results in a median lifespan of 68 days (n = 331), much higher compared to control as determined by Log-rank Mantel–Cox test at <span class="html-italic">p</span>-value 0.0003; the inhibition of <span class="html-italic">Debcl</span> directed by the <span class="html-italic">Ddc-Gal4</span>; <span class="html-italic">UAS-Drp1</span> transgene result in the median lifespan of 72 (n = 303); similar to control, determined by Log-rank Mantel–Cox test at <span class="html-italic">p</span>-value 0.021, with Bonferroni correction. (<b>B</b>) The GraphPad prism8 generated graph of the climbing abilities of <span class="html-italic">Ddc-Gal4 UAS-Drp1</span> flies with the expression of <span class="html-italic">Buffy, Buffy-RNAi</span>, <span class="html-italic">Debcl<sup>EY05743</sup></span><span class="html-italic">, Debcl-RNAi<sup>v47515</sup></span> and control. The climbing abilities of flies overexpressing <span class="html-italic">Buffy</span> have rescued compared to control as determined in the climbing curve’s non-linear fitting by a 95% confidence interval (<span class="html-italic">p</span> &lt; 0.0001). The climbing ability of the flies has further diminished through the ectopic expression of <span class="html-italic">UAS-Buffy-RNAi</span> and <span class="html-italic">UAS-Debcl<sup>EY05743</sup></span> as determined in the non-linear fitting of the climbing curve by a 95% confidence interval at a <span class="html-italic">p</span>-value 0.0004 and 0.0002 respectively (n = 50). The climbing abilities of flies expressing <span class="html-italic">Debcl-RNAi<sup>v47515</sup></span> has been rescued compared to control as determined by the non-linear fitting of the climbing curve by a 95% confidence interval (<span class="html-italic">p</span> value &lt; 0.0001). The graph of longevity assay was generated by GraphPad prism8 non-linear regression curve.</p>
Full article ">
25 pages, 753 KiB  
Article
Multi-Task Multi-Objective Evolutionary Search Based on Deep Reinforcement Learning for Multi-Objective Vehicle Routing Problems with Time Windows
by Jianjun Deng, Junjie Wang, Xiaojun Wang, Yiqiao Cai and Peizhong Liu
Symmetry 2024, 16(8), 1030; https://doi.org/10.3390/sym16081030 - 12 Aug 2024
Viewed by 824
Abstract
The vehicle routing problem with time windows (VRPTW) is a widely studied combinatorial optimization problem in supply chains and logistics within the last decade. Recent research has explored the potential of deep reinforcement learning (DRL) as a promising solution for the VRPTW. However, [...] Read more.
The vehicle routing problem with time windows (VRPTW) is a widely studied combinatorial optimization problem in supply chains and logistics within the last decade. Recent research has explored the potential of deep reinforcement learning (DRL) as a promising solution for the VRPTW. However, the challenge of addressing the VRPTW with many conflicting objectives (MOVRPTW) still remains for DRL. The MOVRPTW considers five conflicting objectives simultaneously: minimizing the number of vehicles required, the total travel distance, the travel time of the longest route, the total waiting time for early arrivals, and the total delay time for late arrivals. To tackle the MOVRPTW, this study introduces the MTMO/DRP-AT, a multi-task multi-objective evolutionary search algorithm, by making full use of both DRL and the multitasking mechanism. In the MTMO/DRL-AT, a two-objective MOVRPTW is constructed as an assisted task, with the objectives being to minimize the total travel distance and the travel time of the longest route. Both the main task and the assisted task are simultaneously solved in a multitasking scenario. Each task is decomposed into scalar optimization subproblems, which are then solved by an attention model trained using DRL. The outputs of these trained models serve as the initial solutions for the MTMO/DRL-AT. Subsequently, the proposed algorithm incorporates knowledge transfer and multiple local search operators to further enhance the quality of these promising solutions. The simulation results on real-world benchmarks highlight the superior performance of the MTMO/DRL-AT compared to several other algorithms in solving the MOVRPTW. Full article
Show Figures

Figure 1

Figure 1
<p>Solution representation for the MOVRPTW. (<b>a</b>) A solution for the MOVRTPW. (<b>b</b>) The solution representation.</p>
Full article ">Figure 2
<p>The neighborhood-based parameter-transfer strategy.</p>
Full article ">Figure 3
<p>The structure of the encoder in the model.</p>
Full article ">Figure 4
<p>The structure of the decoder in the model.</p>
Full article ">Figure 5
<p>Distributions of the approximate Pareto fronts obtained by the MTMO/DRL-AT and LSMOVRPTW on the representative real-world instances: (<b>a</b>) 50-1-2 at f1-f3 plane; (<b>b</b>) 150-1-1 at f1-f3 plane; (<b>c</b>) 250-2-2 at f1-f3 plane; (<b>d</b>) 50-1-2 at f2-f3 plane; (<b>e</b>) 150-1-1 at f2-f3 plane; (<b>f</b>) 250-2-2 at f2-f3 plane.</p>
Full article ">
19 pages, 3062 KiB  
Systematic Review
Eicosapentaenoic Acid (EPA) and Docosahexaenoic Acid (DHA) Ameliorate Heart Failure through Reductions in Oxidative Stress: A Systematic Review and Meta-Analysis
by Jayant Seth, Sohat Sharma, Cameron J. Leong and Simon W. Rabkin
Antioxidants 2024, 13(8), 955; https://doi.org/10.3390/antiox13080955 - 6 Aug 2024
Viewed by 858
Abstract
The objectives of this study were to explore the role that eicosapentaenoic acid (EPA) and/or docosahexaenoic acid (DHA) plays in heart failure (HF), highlighting the potential connection to oxidative stress pathways. Following PRISMA guidelines, we conducted electronic searches of the literature in MEDLINE [...] Read more.
The objectives of this study were to explore the role that eicosapentaenoic acid (EPA) and/or docosahexaenoic acid (DHA) plays in heart failure (HF), highlighting the potential connection to oxidative stress pathways. Following PRISMA guidelines, we conducted electronic searches of the literature in MEDLINE and EMBASE focusing on serum EPA and/or DHA and EPA and/or DHA supplementation in adult patients with heart failure or who had heart failure as an outcome of this study. We screened 254 studies, encompassing RCTs, observational studies, and cohort studies that examined HF outcomes in relation to either serum concentrations or dietary supplementation of EPA and/or DHA. The exclusion criteria were pediatric patients, non-HF studies, abstracts, editorials, case reports, and reviews. Eleven studies met our criteria. In meta-analyses, high serum concentrations of DHA were associated with a lower rate of heart failure with a hazard ratio of 0.74 (CI = 0.59–0.94). High serum concentrations of EPA also were associated with an overall reduction in major adverse cardiovascular events with a hazard ratio of 0.60 (CI = 0.46–0.77). EPA and DHA, or n3-PUFA administration, were associated with an increased LVEF with a mean difference of 1.55 (CI = 0.07–3.03)%. A potential explanation for these findings is the ability of EPA and DHA to inhibit pathways by which oxidative stress damages the heart or impairs cardiac systolic or diastolic function producing heart failure. Specifically, EPA may lower oxidative stress within the heart by reducing the concentration of reactive oxygen species (ROS) within cardiac tissue by (i) upregulating nuclear factor erythroid 2-related factor 2 (Nrf2), which increases the expression of antioxidant enzyme activity, including heme oxygenase-1, thioredoxin reductase 1, ferritin light chain, ferritin heavy chain, and manganese superoxide dismutase (SOD), (ii) increasing the expression of copper–zinc superoxide dismutase (MnSOD) and glutathione peroxidase, (iii) targeting Free Fatty Acid Receptor 4 (Ffar4), (iv) upregulating expression of heme-oxygenase-1, (v) lowering arachidonic acid levels, and (vi) inhibiting the RhoA/ROCK signaling pathway. DHA may lower oxidative stress within the heart by (i) reducing levels of mitochondrial-fission-related protein DRP-1(ser-63), (ii) promoting the incorporation of cardiolipin within the mitochondrial membrane, (iii) reducing myocardial fibrosis, which leads to diastolic heart failure, (iv) reducing the expression of genes such as Appa, Myh7, and Agtr1α, and (v) reducing inflammatory cytokines such as IL-6, TNF-α. In conclusion, EPA and/or DHA have the potential to improve heart failure, perhaps mediated by their ability to modulate oxidative stress. Full article
(This article belongs to the Special Issue Oxidative Stress in Cardiovascular Diseases (CVDs))
Show Figures

Figure 1

Figure 1
<p>PRISMA flow diagram.</p>
Full article ">Figure 2
<p>Forest plot of serum DHA concentration and its association with major adverse cardiac events in patients with heart failure [<a href="#B1-antioxidants-13-00955" class="html-bibr">1</a>,<a href="#B2-antioxidants-13-00955" class="html-bibr">2</a>,<a href="#B14-antioxidants-13-00955" class="html-bibr">14</a>,<a href="#B15-antioxidants-13-00955" class="html-bibr">15</a>,<a href="#B16-antioxidants-13-00955" class="html-bibr">16</a>,<a href="#B17-antioxidants-13-00955" class="html-bibr">17</a>,<a href="#B18-antioxidants-13-00955" class="html-bibr">18</a>].</p>
Full article ">Figure 3
<p>Forest plot of serum EPA concentration and its association with major adverse cardiac events in patients with heart failure [<a href="#B1-antioxidants-13-00955" class="html-bibr">1</a>,<a href="#B2-antioxidants-13-00955" class="html-bibr">2</a>,<a href="#B14-antioxidants-13-00955" class="html-bibr">14</a>,<a href="#B15-antioxidants-13-00955" class="html-bibr">15</a>,<a href="#B16-antioxidants-13-00955" class="html-bibr">16</a>,<a href="#B17-antioxidants-13-00955" class="html-bibr">17</a>,<a href="#B18-antioxidants-13-00955" class="html-bibr">18</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Forest plot of mean differences in left ventricular ejection fraction before and after N3-PUFA supplementation. (<b>b</b>) Forest plot of mean Differences in left ventricular ejection fraction before and after EPA and/or DHA supplementation [<a href="#B20-antioxidants-13-00955" class="html-bibr">20</a>,<a href="#B22-antioxidants-13-00955" class="html-bibr">22</a>,<a href="#B23-antioxidants-13-00955" class="html-bibr">23</a>,<a href="#B24-antioxidants-13-00955" class="html-bibr">24</a>,<a href="#B25-antioxidants-13-00955" class="html-bibr">25</a>,<a href="#B26-antioxidants-13-00955" class="html-bibr">26</a>,<a href="#B27-antioxidants-13-00955" class="html-bibr">27</a>,<a href="#B28-antioxidants-13-00955" class="html-bibr">28</a>].</p>
Full article ">Figure 5
<p>ROS contribute to the pathogenesis of HF in cardiomyocytes. Under normal conditions, nitric oxide synthase (NOS<sub>3</sub>) catalyzes the conversion of O<sub>2</sub> + L-Arginine ⇌ NO + L-Citrulline. During conditions of oxidative stress, uncoupling of NOS<sub>3</sub> due to oxidation of BH<sub>4</sub> leads to the production of harmful ROS. NADPH oxidase is a cytosolic enzyme responsible for the oxidation of NADPH by the equation: NADPH + 2O<sub>2</sub> ⇌ NADP+ + 2O<sub>2</sub><sup>−</sup> + H<sup>+</sup>. This generates two superoxide (O<sub>2</sub><sup>−</sup>) radicals as a byproduct of the reaction. Xanthine oxidase (XO) first converts hypoxanthine to xanthine, then further oxidizes it to produce uric acid. This is performed by equations: (1) hypoxanthine + H<sub>2</sub>O + O<sub>2</sub> ⇌ xanthine + H<sub>2</sub>O<sub>2</sub>, and (2) xanthine + H<sub>2</sub>O + O<sub>2</sub> ⇌ uric acid + H<sub>2</sub>O<sub>2</sub>. Occasionally, mitochondria produce ROS due to the “leakage” of electrons during cellular respiration. This is mainly in the form of O<sub>2</sub><sup>−</sup> radicals. Abbreviations: RyR2 = ryanodine receptor 2, ERK = Extracellular signal-regulated kinase. JNK = Jun nuclear kinase, MAPK = mitogen-activated protein kinase, and AP-1 = activator protein-1.</p>
Full article ">Figure 6
<p>EPA’s role in reducing oxidative stress within cardiomyocytes. Under normal conditions, EPA (eicosapentaenoic acid) is involved in various cellular processes that promote mitochondrial efficiency and protect against dysfunction. During conditions of oxidative stress, EPA is shown to reduce the breakdown of cardiolipin and mitigate mitochondrial dysfunction. It improves O2 consumption efficiency by Complex IV in the mitochondrial membrane, facilitated by MnSOD (manganese superoxide dismutase). EPA also prevents the translocation of Fyn to the cell membrane, thereby inhibiting the activation of Rho Kinase. Additionally, EPA interacts with Ffar4, leading to the production of 18-HEPE. EPA also activates transcription factor Nrf2. This interaction results in altered gene expression through Nrf2 and 18-HEPE, which enhances the expression of antioxidant enzymes, including heme-oxygenase-1.</p>
Full article ">Figure 7
<p>DHA’s role in reducing oxidative stress within cardiomyocytes. DHA plays a role in alleviating oxidative stress within cardiomyocytes. DHA increases the promotion of cardiolipin within the mitochondrial membrane, which subsequently reduces mitochondrial stress and subsequent dysfunction due to oxidative damage. DHA also attenuates the Nrkb pathway and reduces the production of inflammatory cytokines and activation of genes such as appa, Myrh7, and Agtr-alpha, which have been implicated in hypertrophy and fibrosis of cardiomyocytes. DHA also enhances eNOS activity through AKT and HSP90 activation, which subsequently enhances nitric oxide production and minimizes oxidative stress.</p>
Full article ">
21 pages, 6646 KiB  
Article
Extraction, Purification, Sulfated Modification, and Biological Activities of Dandelion Root Polysaccharides
by Xiao Wu, Na Li, Zeng Dong, Qin Yin, Tong Zhou, Lixiang Zhu, Hanxi Yan, Ziping Chen and Kefeng Zhai
Foods 2024, 13(15), 2393; https://doi.org/10.3390/foods13152393 - 29 Jul 2024
Viewed by 615
Abstract
In this study, polysaccharides were extracted at a rate of 87.5% ± 1.5% from native dandelion roots, and the dandelion root polysaccharides (DRPs) were then chemically modified to obtain sulfated polysaccharides (SDRPs) with a degree of substitution of 1.49 ± 0.07. The effects [...] Read more.
In this study, polysaccharides were extracted at a rate of 87.5% ± 1.5% from native dandelion roots, and the dandelion root polysaccharides (DRPs) were then chemically modified to obtain sulfated polysaccharides (SDRPs) with a degree of substitution of 1.49 ± 0.07. The effects of modification conditions, physicochemical characterizations, structural characteristics, antioxidant properties, hypoglycemic activity, and proliferative effects on probiotics of DRP derivatives were further investigated. Results showed that the optimum conditions for sulfation of DRPs included esterification reagents (concentrated sulfuric acid: n-butanol) ratio of 3:1, a reaction temperature of 0 °C, a reaction time of 1.5 h, and the involvement of 0.154 g of ammonium sulfate. The DRPs and SDRPs were composed of six monosaccharides, including mannose, glucosamine, rhamnose, glucose, galactose, and arabinose. Based on infrared spectra, the peaks of the characteristic absorption bands of S=O and C-O-S appeared at 1263 cm−1 and 836 cm−1. Compared with DRPs, SDRPs had a significantly lower relative molecular mass and a three-stranded helical structure. NMR analysis showed that sulfated modification mainly occurred on the hydroxyl group at C6. SDRPs underwent a chemical shift to higher field strength, with their characteristic signal peaking in the region of 1.00–1.62 ppm. Scanning electron microscopy (SEM) analysis indicated that the surface morphology of SDRPs was significantly changed. The structure of SDRPs was finer and more fragmented than DRPs. Compared with DRPs, SDRPs showed better free radical scavenging ability, higher Fe2+chelating ability, and stronger inhibition of α-glucosidase and α-amylase. In addition, SDRPs had an excellent promotional effect on the growth of Lactobacillus plantarum 10665 and Lactobacillus acidophilus. Therefore, this study could provide a theoretical basis for the development and utilization of DRPs. Full article
Show Figures

Figure 1

Figure 1
<p>Separation (<b>A</b>) and purification (<b>B</b>,<b>C</b>) of dandelion polysaccharides.</p>
Full article ">Figure 2
<p>Effect of reaction time (<b>A</b>), esterifier dosage (<b>B</b>), ammonium sulfate addition (<b>C</b>), and reaction temperature (<b>D</b>) on the degree of substitution of dandelion polysaccharides.</p>
Full article ">Figure 3
<p>Contour plots (<b>A</b>,<b>C</b>,<b>E</b>) and three-dimensional response surface plots (<b>B</b>,<b>D</b>,<b>F</b>) about the effects of reaction time, esterifier dosage, the additional amount of ammonium sulfate, and interaction on the DS of sulfate of DRP.</p>
Full article ">Figure 3 Cont.
<p>Contour plots (<b>A</b>,<b>C</b>,<b>E</b>) and three-dimensional response surface plots (<b>B</b>,<b>D</b>,<b>F</b>) about the effects of reaction time, esterifier dosage, the additional amount of ammonium sulfate, and interaction on the DS of sulfate of DRP.</p>
Full article ">Figure 4
<p>FT-IR spectra of DRP, SDRP (<b>A</b>), and Congo red test (<b>B</b>).</p>
Full article ">Figure 5
<p>Chromatograms of monosaccharide compositions of standard substance mixture (<b>A</b>), DRP (<b>B</b>), and SDRP (<b>C</b>).</p>
Full article ">Figure 6
<p>The NMR spectra of DRP and SDRP.((<b>A</b>,<b>B</b>) were the <sup>13</sup>C NMR spectrum of DRPs and SDRPs respectively; (<b>C</b>,<b>D</b>) were the <sup>1</sup>H NMR spectrum of DRPs and SDRPs respectively).</p>
Full article ">Figure 7
<p>Scanning electron microscopy images of DRP (<b>A1</b>–<b>A3</b>) and SDRP (<b>B1</b>–<b>B3</b>).</p>
Full article ">Figure 8
<p>(<b>A</b>) DPPH radical scavenging assay; (<b>B</b>) superoxide anion scavenging activity assay; (<b>C</b>) hydroxyl radical scavenging ability assay; and (<b>D</b>) measurement of ferrous ion chelating capacity.</p>
Full article ">Figure 9
<p>Studies on the hypoglycemic activity of sulfated dandelion polysaccharides in vitro ((<b>A</b>) Inhibition of α-amylase by polysaccharides and (<b>B</b>) inhibition of α-glucosidase by polysaccharides).</p>
Full article ">Figure 10
<p>Probiotic proliferative effects (<b>A</b>,<b>B</b>) and probiotic growth curves (<b>C</b>,<b>D</b>). (<b>a</b>) SDRP medium OD<sub>600nm</sub>; (<b>b</b>) FOS mediumOD<sub>600nm</sub>; (<b>c</b>) SDRP medium pH; and (<b>d</b>) FOS medium pH. *: When the mass concentration of polysaccharides is the same, there is a significant difference (<span class="html-italic">p</span> &lt; 0.05) between S-DRP, DRP, and FOS.</p>
Full article ">
20 pages, 8950 KiB  
Article
The Scavenging Activity of Coenzyme Q10 Plus a Nutritional Complex on Human Retinal Pigment Epithelial Cells
by Maria Hernandez, Sergio Recalde, Jaione Bezunartea, Maite Moreno-Orduña, Idoia Belza, Ainara Chas-Prat, Elena Perugini, Alfredo Garcia-Layana and Patricia Fernández-Robredo
Int. J. Mol. Sci. 2024, 25(15), 8070; https://doi.org/10.3390/ijms25158070 - 24 Jul 2024
Viewed by 630
Abstract
Age-related macular degeneration (AMD) and diabetic retinopathy (DR) are common retinal diseases responsible for most blindness in working-age and elderly populations. Oxidative stress and mitochondrial dysfunction play roles in these pathogenesis, and new therapies counteracting these contributors could be of great interest. Some [...] Read more.
Age-related macular degeneration (AMD) and diabetic retinopathy (DR) are common retinal diseases responsible for most blindness in working-age and elderly populations. Oxidative stress and mitochondrial dysfunction play roles in these pathogenesis, and new therapies counteracting these contributors could be of great interest. Some molecules, like coenzyme Q10 (CoQ10), are considered beneficial to maintain mitochondrial homeostasis and contribute to the prevention of cellular apoptosis. We investigated the impact of adding CoQ10 (Q) to a nutritional antioxidant complex (Nutrof Total®; N) on the mitochondrial status and apoptosis in an in vitro hydrogen peroxide (H2O2)-induced oxidative stress model in human retinal pigment epithelium (RPE) cells. H2O2 significantly increased 8-OHdG levels (p < 0.05), caspase-3 (p < 0.0001) and TUNEL intensity (p < 0.01), and RANTES (p < 0.05), caspase-1 (p < 0.05), superoxide (p < 0.05), and DRP-1 (p < 0.05) levels, and also decreased IL1β, SOD2, and CAT gene expression (p < 0.05) vs. control. Remarkably, Q showed a significant recovery in IL1β gene expression, TUNEL, TNFα, caspase-1, and JC-1 (p < 0.05) vs. H2O2, and NQ showed a synergist effect in caspase-3 (p < 0.01), TUNEL (p < 0.0001), mtDNA, and DRP-1 (p < 0.05). Our results showed that CoQ10 supplementation is effective in restoring/preventing apoptosis and mitochondrial stress-related damage, suggesting that it could be a valid strategy in degenerative processes such as AMD or DR. Full article
Show Figures

Figure 1

Figure 1
<p>DNA oxidative damage analyzed as 8-OHdG levels in ARPE-19 cells’ supernatants by ELISA in basal conditions (<b>A</b>) and after the addition of H<sub>2</sub>O<sub>2</sub> (600 µM, 1 h) and antioxidant treatments in concomitance for 30 min (<b>B</b>) (* <span class="html-italic">p</span> &lt; 0.05 vs. control) (<span class="html-italic">n</span> = 3). The application of NQ showed a tendency to significantly reduce 8-OHdG levels vs. H<sub>2</sub>O<sub>2</sub> control group (<span class="html-italic">p</span> = 0.0550).</p>
Full article ">Figure 2
<p>Percentage of the fluorescence intensity of caspase-3 (green) immunolabeling in basal conditions after Q, N, and NQ showed statistical differences between control and Q (<span class="html-italic">p</span> &lt; 0.05) (<b>A</b>). Oxidative environment induced by H<sub>2</sub>O<sub>2</sub> increased caspase-3 immunofluorescence vs. control group (<span class="html-italic">p</span> &lt; 0.001) (<b>B</b>) (<span class="html-italic">n</span> = 3). After N and NQ with oxidative stress, significant differences were observed vs. H<sub>2</sub>O<sub>2</sub> group (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001). Nuclei were labeled with 4′,6-diamidino-2-phenylindole (DAPI, blue). Scale bar: 20 µm.</p>
Full article ">Figure 3
<p>Percentage of TUNEL fluorescence intensity (red) in basal conditions after Q, N, and NQ showed statistical differences between control and Q (* <span class="html-italic">p</span> &lt; 0.05) (<b>A</b>). H<sub>2</sub>O<sub>2</sub> group showed a significant increase vs. control group (*** <span class="html-italic">p</span> &lt; 0.001). After Q, N, and NQ treatments in concomitance with oxidative stress, a significant reduction was observed in Q and NQ vs. H<sub>2</sub>O<sub>2</sub> group (* <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001) (<b>B</b>) (<span class="html-italic">n</span> = 3). Nuclei were labeled with 4′,6-diamidino-2-phenylindole (DAPI, blue). Scale bar: 20 µm.</p>
Full article ">Figure 4
<p>Quantification of cytokine levels in which changes have been observed in standard conditions and under oxidative stress with treatments Q, N, and NQ (<span class="html-italic">n</span> = 4). Levels of IL17A, IL6, and RANTES in ARPE-19 cells supernatant (<b>A</b>–<b>C</b>) in standard conditions. Levels of caspase-1, IL12-p70, and RANTES in ARPE-19 cells supernatant after oxidative stress conditions (<b>D</b>–<b>F</b>) and TNFα levels in lysates after oxidative stress conditions (<b>G</b>). Lysates’ data are presented as pg/µg protein and supernatants’ data are presented as pg/mL. RANTES data are presented as RFU. For all data mean ± SEM are presented. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. H<sub>2</sub>O<sub>2</sub>. Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>.</p>
Full article ">Figure 5
<p>Quantification of <span class="html-italic">SOD2</span> gene expression of cultured ARPE-19 cells in standard conditions and under oxidative stress with Q, N, and NQ treatments (<span class="html-italic">n</span> = 4). <span class="html-italic">SOD2</span> expression in standard conditions showed a significant reduction with all antioxidant treatments (<b>A</b>). H<sub>2</sub>O<sub>2</sub> group showed a significant decrease vs. control group (* <span class="html-italic">p</span> &lt; 0.05). <span class="html-italic">SOD2</span> expression in ARPE-19 cells with 2 h of H<sub>2</sub>O<sub>2</sub> in concomitance showed no significant reduction with treatments (<b>B</b>). For all data, mean ± SEM are presented. * <span class="html-italic">p</span> &lt; 0.05 vs H<sub>2</sub>O<sub>2</sub> group. Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>.</p>
Full article ">Figure 6
<p>Quantification of <span class="html-italic">ILβ1</span> expression of cultured ARPE-19 cells in standard conditions and under oxidative stress with treatments Q, N and NQ (<span class="html-italic">n</span> = 4). <span class="html-italic">ILβ1</span> expression significantly decreased with N antioxidant treatment * <span class="html-italic">p</span> &lt; 0.05 vs. control (<b>A</b>). <span class="html-italic">ILβ1</span> expression in ARPE-19 cells with 1 h of H<sub>2</sub>O<sub>2</sub> in concomitance decreased after Q and N treatment (<b>B</b>) (* <span class="html-italic">p</span> &lt; 0.05) vs. H<sub>2</sub>O<sub>2</sub> group. Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>.</p>
Full article ">Figure 7
<p>Quantification of <span class="html-italic">CAT</span> expression of cultured ARPE-19 cells in standard conditions and under oxidative stress with treatments Q, N, and NQ (<span class="html-italic">n</span> = 4). No changes were observed in <span class="html-italic">CAT</span> expression in basal conditions with antioxidant treatments (<b>A</b>). <span class="html-italic">CAT</span> expression in ARPE-19 cells with 1 h of H<sub>2</sub>O<sub>2</sub> concomitance showed a decrease only in H<sub>2</sub>O<sub>2</sub> group vs. control (* <span class="html-italic">p</span> &lt; 0.05) (<b>B</b>). Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>.</p>
Full article ">Figure 8
<p>Percentage of mitochondrial superoxide indicator in live ARPE-19 cells measured by MitoSOX (red) in standard conditions and under oxidative stress with treatments Q, N, and NQ (<span class="html-italic">n</span> = 3). No changes in basal conditions were observed (<b>A</b>). H<sub>2</sub>O<sub>2</sub> group showed a significant increase vs. control group (* <span class="html-italic">p</span> &lt; 0.05), (<b>B</b>) and after H<sub>2</sub>O<sub>2</sub> in concomitance, only the NQ treatment decreased MitoSOX (<span class="html-italic">p</span> = 0.0533) (B). Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>. * <span class="html-italic">p</span> &lt; 0.05. Nuclei were labeled with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Scale bar: 20 µm.</p>
Full article ">Figure 9
<p>Mitochondrial DNA amount of cultured ARPE-19 cells measured by <span class="html-italic">12S</span> RT-PCR under standard conditions and under oxidative stress with treatments Q, N, and NQ (<span class="html-italic">n</span> = 4). No changes were observed in the mitochondrial DNA amount in cells treated with different treatments under basal conditions (<b>A</b>). H<sub>2</sub>O<sub>2</sub> group showed an almost significant increase vs. the control group (<span class="html-italic">p</span> = 0.0690) (<b>B</b>) and the NQ group in concomitance with H<sub>2</sub>O<sub>2</sub> significantly decreased mtDNA vs. the H<sub>2</sub>O<sub>2</sub> group * <span class="html-italic">p</span> &lt; 0.05 (<b>B</b>). Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>.</p>
Full article ">Figure 10
<p>Mitochondrial membrane potential (mtΔψ) determined by live JC-1 measurement in ARPE-19 cells under basal conditions (<b>A</b>,<b>B</b>) and in concomitance with oxidative stress conditions with antioxidants treatments (<b>C</b>,<b>D</b>) (<span class="html-italic">n</span> = 3). J-monomers, green; J-aggregates, red. No changes were observed in JC-1 under basal conditions (<b>A</b>); however, in concomitance with H<sub>2</sub>O<sub>2</sub> only, the Q treatment significantly decreased the mtΔψ vs. H<sub>2</sub>O<sub>2</sub> group (<b>B</b>) (* <span class="html-italic">p</span> &lt; 0.05). Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>. Nuclei were labeled with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Scale bar: 20 µm.</p>
Full article ">Figure 11
<p>Percentage of mitochondrial DRP-1 (red) measurement in ARPE-19 cells under basal conditions (<b>A</b>,<b>B</b>) and under oxidative stress conditions with treatments Q, N, and NQ (<b>C</b>,<b>D</b>) (<span class="html-italic">n</span> = 4). Q treatment significant increased DRP-1 under basal conditions (<b>B</b>) (** <span class="html-italic">p</span> &lt; 0.01). H<sub>2</sub>O<sub>2</sub> group showed a significant increase vs. control group (* <span class="html-italic">p</span>&lt; 0.05) (<b>B</b>). After concomitance with H<sub>2</sub>O<sub>2,</sub> only NQ treatment showed a significant decrease vs. H<sub>2</sub>O<sub>2</sub> group (* <span class="html-italic">p</span> &lt; 0.05). Q—coenzyme Q<sub>10</sub>, N—Nutrof total, NQ—Nutrof total + CoQ<sub>10</sub>. Nuclei were labeled with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Scale bar: 20 µm.</p>
Full article ">
32 pages, 7516 KiB  
Article
Novel Thienopyrimidine-Hydrazinyl Compounds Induce DRP1-Mediated Non-Apoptotic Cell Death in Triple-Negative Breast Cancer Cells
by Saloni Malla, Angelique Nyinawabera, Rabin Neupane, Rajiv Pathak, Donghyun Lee, Mariam Abou-Dahech, Shikha Kumari, Suman Sinha, Yuan Tang, Aniruddha Ray, Charles R. Ashby, Mary Qu Yang, R. Jayachandra Babu and Amit K. Tiwari
Cancers 2024, 16(15), 2621; https://doi.org/10.3390/cancers16152621 - 23 Jul 2024
Viewed by 1281
Abstract
Apoptosis induction with taxanes or anthracyclines is the primary therapy for TNBC. Cancer cells can develop resistance to anticancer drugs, causing them to recur and metastasize. Therefore, non-apoptotic cell death inducers could be a potential treatment to circumvent apoptotic drug resistance. In this [...] Read more.
Apoptosis induction with taxanes or anthracyclines is the primary therapy for TNBC. Cancer cells can develop resistance to anticancer drugs, causing them to recur and metastasize. Therefore, non-apoptotic cell death inducers could be a potential treatment to circumvent apoptotic drug resistance. In this study, we discovered two novel compounds, TPH104c and TPH104m, which induced non-apoptotic cell death in TNBC cells. These lead compounds were 15- to 30-fold more selective in TNBC cell lines and significantly decreased the proliferation of TNBC cells compared to that of normal mammary epithelial cell lines. TPH104c and TPH104m induced a unique type of non-apoptotic cell death, characterized by the absence of cellular shrinkage and the absence of nuclear fragmentation and apoptotic blebs. Although TPH104c and TPH104m induced the loss of the mitochondrial membrane potential, TPH104c- and TPH104m-induced cell death did not increase the levels of cytochrome c and intracellular reactive oxygen species (ROS) and caspase activation, and cell death was not rescued by incubating cells with the pan-caspase inhibitor, carbobenzoxy-valyl-alanyl-aspartyl-[O-methyl]-fluoromethylketone (Z-VAD-FMK). Furthermore, TPH104c and TPH104m significantly downregulated the expression of the mitochondrial fission protein, DRP1, and their levels determined their cytotoxic efficacy. Overall, TPH104c and TPH104m induced non-apoptotic cell death, and further determination of their cell death mechanisms will aid in the development of new potent and efficacious anticancer drugs to treat TNBC. Full article
(This article belongs to the Topic Recent Advances in Anticancer Strategies)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The cytotoxicity (i.e., anticancer efficacy) of TPH104c and TPH104m in different breast cancer cell lines. (<b>a</b>) The selectivity of TPH104c and TPH104m for TNBC, compared to normal, non-TNBC cell lines and TNBC, compared to normal breast cell line. (<b>b</b>) The cell viability curves of BT-20 cells after incubation for 72 h, with varying concentrations of TPH104c or TPH104m, using the MTT, CTB, or SRB assays, respectively. (<b>c</b>) Quantitative graphs of percent (%) cell viability data obtained using IncuCyte S3 software based on phase-contrast images of BT-20 cells incubated for 72 h with vehicle or varying concentrations of TPH104c, TPH104m and media. (<b>d</b>) Real-time live-cell imaging pictures of BT-20 cells after incubation with TPH104c and TPH104m for 72 hrs, in an Incucyte Cytotox green reagent—containing media. The images show the green fluorescence intensity of cytotox green dye, which stains dead or non-viable cells. (<b>e</b>) Colony formation assay for BT-20 cells that were incubated with vehicle (0 µM), 0.1, 0.3, or 1 μM of TPH104c or TPH104m. The images show the effect of TPH104c and TPH104m on colony density and size. (<b>f</b>) Bar graph summarizing the effect of different concentrations of TPH104c or TPH104m on the size of the colonies formed by BT-20 cells. The results represent the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>The effect of TPH104c or TPH104m on the cell cycle in BT-20 cells. Representative figures showing the distribution of BT-20 cells in different phases of the cell cycle after incubation with vehicle (0 μM), (<b>a</b>) TPH104c, or (<b>c</b>) TPH104m (0.5, 1, and 2 μM). BT-20 cells were stained with PI and subjected to flow cytometry. Count (<span class="html-italic">y</span>-axis) represents the cell population used in the flow cytometric analysis, and PE-A (<span class="html-italic">x</span>-axis) represents the cells stained with PI. Quantitative histograms depicting the percent change in BT-20 cells in the SubG1, G1, S, and G2 phases of the cell cycle upon treatment with (<b>b</b>) TPH104c or (<b>d</b>) TPH104m. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. The data represent the average ± SD of three separate experiments performed in triplicate.</p>
Full article ">Figure 3
<p>The effect of TPH104c on the levels of apoptotic and anti-apoptotic proteins in BT-20 cells. (<b>a</b>) Representative images featuring morphological changes in BT-20 cells (under 20× magnification) after incubation with vehicle (0 μM, media without drug), 0.1, 0.3, or 1 μM of TPH104c for 0, 24, 48 or 72 h. (<b>b</b>) Representative images of BT-20 cells with vehicle (0 μM), 2, or 5 μM of TPH104c for 24 h or paclitaxel (PTX, 1 μM, a positive control) and stained with Hoechst 33342 dye. TPH104c did not produce condensed or fragmented nuclei compared to cells incubated with paclitaxel (PTX). Scale bar = 25 μM. (<b>c</b>) Western blot images representing the levels of the apoptotic molecules, cleaved caspase-3, caspase-3, cleaved caspase-7, caspase-7, cleaved caspase-9, caspase-9, cleaved caspase-8, caspase-8, BAX, BAK, BCL-2, cleaved PARP and PARP, following incubation with vehicle (0 μM), 0.5, 1, 2 or 5 μM of TPH104c. The proteins are expressed as a ratio to β-actin, followed by normalization to the vehicle control. (<b>d</b>) The level of each protein is shown by histograms. Clvd = cleaved; Csp = caspase. The data represent the average ± SEM of four separate studies. (<b>e</b>) Caspase-Glo 3/7 assay results are represented as a bar graph and curve, showing a decrease in the levels of caspase-3 and caspase-7 by TPH104c, in a concentration-dependent manner in BT-20 cells, after 24 h of incubation. In contrast, 1 µMof PTX induced caspase- 3 and 7 activity (n = 2). (<b>f</b>) The IC<sub>50</sub> values, using the MTT assay, for TPH104c in BT-20 cells that were preincubated with zVAD-FMK (a pan-caspase inhibitor) and then incubated with varying concentrations of TPH104c for 72 h. The data were obtained from three independent experiments conducted in triplicate and represent the average ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 and ns means non-significant. Original Western Blot images can be found in <a href="#app1-cancers-16-02621" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 4
<p>The effect of TPH104c on apoptotic and anti-apoptotic proteins in BT-20 cells. (<b>a</b>) Representative images featuring morphological changes in BT-20 cells (20× magnification) after incubation with vehicle (media without the TPH compounds or paclitaxel (PTX)), 0.1, 0.3, or 1 μM of TPH104m, at 0, 24, 48 or 72 h post-incubation. (<b>b</b>) Representative images of BT-20 cells incubated with 2 or 5 μM of TPH104m or PTX (1 μM,) a positive control) or vehicle control and stained with Hoechst 33342 dye. TPH104c did not produce condensed or fragmented nuclei, compared to cells incubated with PTX. Scale bar = 25 μM. (<b>c</b>) Western blot images for the apoptotic molecules, cleaved caspase-3, caspase-3, cleaved caspase-7, caspase-7, cleaved caspase-9, caspase-9, cleaved caspase-8, caspase-8, BAX, BAK, BCL-2, cleaved PARP, and PARP, following incubation with vehicle (0 µM), 0.5, 1, 2, or 5 μM of TPH104m. The data are expressed as the ratio to β-actin, followed by normalization to the vehicle control. (<b>d</b>) The level of each protein is shown by histograms. Clvd = cleaved; Csp = caspase. The data represent the average ± SEM of four separate studies. (<b>e</b>) Caspase-Glo 3/7 assay results are presented as a bar graph and as a curve, showing that incubation of BT-20 cells with TPH104m for 24 h decreased the levels of caspase 3/7 in a concentration-dependent manner. In contrast, PTX (1 μM) increased the levels of caspase 3 and 7 (n = 2). (<b>f</b>) IC<sub>50</sub> values, using the MTT assay, for TPH104c in BT-20 cells that were preincubated with z-VADfmk and then incubated with varying concentrations of TPH104c for 72 h. The data is obtained from three independent experiments conducted in triplicates and represents the average ± SD. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 and ns means non-significant. Original Western Blot images can be found in <a href="#app1-cancers-16-02621" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 5
<p>TPH104c and TPH104m induced the loss of the mitochondrial membrane potential but did not induce oxidative stress in BT-20 cells. (<b>a</b>) Fluorescent microscopic images of BT-20 cells stained with TMRE dye after incubation with the vehicle for 24 h (0 μM), 2 or 5 μM of TPH104c or TPH104m, and CCCP as a positive control. The TMRE dye is retained in cells with normal structural and functioning mitochondria, producing a high level of red fluorescence, whereas weak or no fluorescence occurred in cells without MMP. Scale bar = 200 µm. (<b>b</b>) Quantitative bar graph illustrating the change in the percentage of red fluorescence in BT-20 cells incubated with 2, or 5 µM of TPH104c and TPH104m or CCCP, compared to cells incubated with media. The results are shown as mean ± SD in triplicate. CCCP = Carbonyl cyanide 3-chlorophenylhydrazone. (<b>c</b>) Immunofluorescence analysis of cytochrome c levels in BT-20 cells incubated with 2 or 5 μM of TPH104c or TPH104m or PTX or vehicle control (0 μM), for 24 h. PTX = Paclitaxel. Scale bar = 50 µm. (<b>d</b>) Bar graphs illustrating the fluorescence intensity of cytochrome c in BT-20 cells incubated with 2 and 5 µM TPH104c and TPH104m or vehicle control (0 μM) for 24 h. (<b>e</b>) Representative images and (<b>f</b>) bar graphs depicting the level of dichlorofluorescein (DCF) fluorescence in BT-20 cells incubated with TPH104c and TPH104m (0 μM (vehicle), 2, or 5 μM) for 24 h, or paclitaxel (2 μM) for 2 h. Images were captured at 20× magnification. Scale bar = 200 µm. Relative fluorescence units of H<sub>2</sub>DCFA in BT-20 cells. The data are expressed as the average fluorescence ± SEM of three separate experiments. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, compared to the vehicle control cells.</p>
Full article ">Figure 6
<p>The effect of TPH104c and TPH104m on the levels of mitochondrial proteins, DRP1 and phosphorylated DRP1 (p-DRP1). (<b>a</b>) Western blot images for the mitochondrial fission proteins, p-DRP1, DRP1, p-MFF, MFF, and Fis1, and the mitochondrial fusion proteins, MFN1, MFN2, or OPA1, following incubation with vehicle (0 μM), 2, or 5 μM of TPH104c and TPH104m. All proteins were expressed as a ratio to β-actin, followed by normalization to the vehicle control. (<b>b</b>) Histograms showing the ratio of phosphorylated proteins to total proteins and individual proteins. All data are presented as the mean ± SEM of 4-5 independent studies. Immunofluorescence analysis of DRP1 (<b>c</b>) and p-DRP1 (<b>e</b>) at Serine 616C in BT-20 cells incubated for 24 h with vehicle (0 μM), 2, or 5 μM of TPH104c or TPH104m. Bar graphs showing the quantification of the fluorescence intensity of DRP1 (<b>d</b>) and p-DRP1 (<b>f</b>). Scale bar = 50 µm. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. Predicted non-covalent interactions of ligands TPH104c and TPH104m. (<b>g</b>) Hydrogen bonds (yellow) shared between TPH104c and DRP-1; (<b>h</b>) Carbon-π and donor-π interactions between TPH104c and DRP-1 (<b>i</b>) Hydrogen bonds (yellow) shared between TPH104m and DRP-1; (<b>j</b>) Carbon-π and donor-π interactions between TPH104m and DRP1. Representative graphs obtained from a Nicoya SPR assay, where a direct drug-protein binding interaction occurred between the Drp1 recombinant protein and varying concentrations of (<b>k</b>) TPH104c (<b>l</b>) TPH104m. Results are shown as the mean ± SD of triplicate experiments. Original Western Blot images can be found in <a href="#app1-cancers-16-02621" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 6 Cont.
<p>The effect of TPH104c and TPH104m on the levels of mitochondrial proteins, DRP1 and phosphorylated DRP1 (p-DRP1). (<b>a</b>) Western blot images for the mitochondrial fission proteins, p-DRP1, DRP1, p-MFF, MFF, and Fis1, and the mitochondrial fusion proteins, MFN1, MFN2, or OPA1, following incubation with vehicle (0 μM), 2, or 5 μM of TPH104c and TPH104m. All proteins were expressed as a ratio to β-actin, followed by normalization to the vehicle control. (<b>b</b>) Histograms showing the ratio of phosphorylated proteins to total proteins and individual proteins. All data are presented as the mean ± SEM of 4-5 independent studies. Immunofluorescence analysis of DRP1 (<b>c</b>) and p-DRP1 (<b>e</b>) at Serine 616C in BT-20 cells incubated for 24 h with vehicle (0 μM), 2, or 5 μM of TPH104c or TPH104m. Bar graphs showing the quantification of the fluorescence intensity of DRP1 (<b>d</b>) and p-DRP1 (<b>f</b>). Scale bar = 50 µm. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. Predicted non-covalent interactions of ligands TPH104c and TPH104m. (<b>g</b>) Hydrogen bonds (yellow) shared between TPH104c and DRP-1; (<b>h</b>) Carbon-π and donor-π interactions between TPH104c and DRP-1 (<b>i</b>) Hydrogen bonds (yellow) shared between TPH104m and DRP-1; (<b>j</b>) Carbon-π and donor-π interactions between TPH104m and DRP1. Representative graphs obtained from a Nicoya SPR assay, where a direct drug-protein binding interaction occurred between the Drp1 recombinant protein and varying concentrations of (<b>k</b>) TPH104c (<b>l</b>) TPH104m. Results are shown as the mean ± SD of triplicate experiments. Original Western Blot images can be found in <a href="#app1-cancers-16-02621" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 7
<p>The cytotoxic efficacy of TPH104c and TPH104m on CRISPR (wild-type control) and partial and complete <span class="html-italic">DRP1</span> knockout (KO) PAC200 cells. (<b>a</b>) Western blot images of DRP1 levels in CRISPR wild-type (control) PAC200 cells PAC200 and complete and partial <span class="html-italic">DRP1</span> KO cells. Bar graphs depicting the IC<sub>50</sub> values of TPH104c and TPH104m in CRISPR wild-type, partial <span class="html-italic">DRP1</span> KO, and complete <span class="html-italic">DRP1</span> KO PAC200 cells after 72 h of incubation, calculated using (<b>b</b>) MTT assay, (<b>c</b>) CTB assay, (<b>d</b>) CTG assay, and (<b>e</b>) SRB assay. (<b>f</b>) Morphological images of CRISPR wild-type and complete DRP1 KO PAC200 cells incubated with 10 μM of TPH104c and TPH104m, for 72 h. Yellow arrows represent a bubble-like formation that indicates bursting. Scale bar, 100 µm. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Original Western Blot images can be found in <a href="#app1-cancers-16-02621" class="html-app">Supplementary Materials</a>.</p>
Full article ">
9 pages, 2547 KiB  
Article
De Novo DNM1L Mutation in a Patient with Encephalopathy, Cardiomyopathy and Fatal Non-Epileptic Paroxysmal Refractory Vomiting
by Beatrice Berti, Daniela Verrigni, Alessia Nasca, Michela Di Nottia, Daniela Leone, Alessandra Torraco, Teresa Rizza, Emanuele Bellacchio, Andrea Legati, Concetta Palermo, Silvia Marchet, Costanza Lamperti, Antonio Novelli, Eugenio Maria Mercuri, Enrico Silvio Bertini, Marika Pane, Daniele Ghezzi and Rosalba Carrozzo
Int. J. Mol. Sci. 2024, 25(14), 7782; https://doi.org/10.3390/ijms25147782 - 16 Jul 2024
Viewed by 858
Abstract
Mitochondrial fission and fusion are vital dynamic processes for mitochondrial quality control and for the maintenance of cellular respiration; they also play an important role in the formation and maintenance of cells with high energy demand including cardiomyocytes and neurons. The DNM1L (dynamin-1 [...] Read more.
Mitochondrial fission and fusion are vital dynamic processes for mitochondrial quality control and for the maintenance of cellular respiration; they also play an important role in the formation and maintenance of cells with high energy demand including cardiomyocytes and neurons. The DNM1L (dynamin-1 like) gene encodes for the DRP1 protein, an evolutionary conserved member of the dynamin family that is responsible for the fission of mitochondria; it is ubiquitous but highly expressed in the developing neonatal heart. De novo heterozygous pathogenic variants in the DNM1L gene have been previously reported to be associated with neonatal or infantile-onset encephalopathy characterized by hypotonia, developmental delay and refractory epilepsy. However, cardiac involvement has been previously reported only in one case. Next-Generation Sequencing (NGS) was used to genetically assess a baby girl characterized by developmental delay with spastic–dystonic, tetraparesis and hypertrophic cardiomyopathy of the left ventricle. Histochemical analysis and spectrophotometric determination of electron transport chain were performed to characterize the muscle biopsy; moreover, the morphology of mitochondria and peroxisomes was evaluated in cultured fibroblasts as well. Herein, we expand the phenotype of DNM1L-related disorder, describing the case of a girl with a heterozygous mutation in DNM1L and affected by progressive infantile encephalopathy, with cardiomyopathy and fatal paroxysmal vomiting correlated with bulbar transitory abnormal T2 hyperintensities and diffusion-weighted imaging (DWI) restriction areas, but without epilepsy. In patients with DNM1L mutations, careful evaluation for cardiac involvement is recommended. Full article
Show Figures

Figure 1

Figure 1
<p>Brain MRI pattern. (<b>a</b>) Brain MRI performed in the acute phase at 10 years of age. Coronal and Sagittal T2 weighted images shows global cerebral and cerebellar atrophy. Red arrow shows hyperintensity of the right bulbar pyramid; (<b>b</b>) brain MRI performed 20 days later shows reduction in the right corresponding bulbar pyramid lesion.</p>
Full article ">Figure 2
<p>Echocardiogram pattern. Echocardiogram performed during acute phase that showed moderate hypertrofic cardiomyopathy of the left ventricle (<b>a</b>) without outflow obstruction and with normal EF (<b>b</b>).</p>
Full article ">Figure 3
<p>Genetic and structural analysis. (<b>a</b>) The c.116G&gt;A variant identified by NGS in <span class="html-italic">DNM1L</span> was confirmed by Sanger sequencing in the patient and appeared absent in both parents; (<b>b</b>) cDNA, obtained by retrotranscription of mRNA from patient’s fibroblasts, revealed balanced expression of the two alleles; (<b>c</b>) DRP1 structure and multiple protein sequence alignment around the site of the p.Ser39Asn replacement. Crystal structure of the human dynamin-1-like protein in complex with GDP-AlF4 (Protein Data Bank code 3W6P) and protein sequence alignment around Ser39, highlighting the role of this serine in the binding of the GTP/GDP ligand and its conservation among different organisms.</p>
Full article ">Figure 4
<p>Western blotting and immunostaining analysis. (<b>a</b>) Immunoblot analysis of total lysates from controls (Ct) and patient (Pt) fibroblasts using DRP1 and GAPDH antibodies. The latter was used as loading control. The steady-state level of DRP1 protein in the patient is in the low range of normal controls. Fibroblasts from a patient with recessive <span class="html-italic">DNM1L</span> mutations (Pt-AR) were used as “positive” control. Values in the graph are given as the mean ± SD (n = 4–5). PT vs. CTs: <span class="html-italic">p</span>-value = 0.018. (<b>b</b>) Characterization of the mitochondrial network: representative images of mitochondrial morphology in control (Ct) and patient (Pt) fibroblasts grown in galactose-supplemented medium. Mitochondrial network of <span class="html-italic">DNM1L</span>-mutant fibroblasts showed an altered mitochondria morphology, with swollen, dots, and “chain-like” structures. (<b>c</b>) Characterization of the peroxisomal network: immunofluorescence staining with the anti-PMP70 antibody of fibroblasts from controls (Ct) and patient (Pt). Fibroblasts from patient displayed organelles longer and larger compared with control. (Scale bar = 10 μm).</p>
Full article ">
13 pages, 3231 KiB  
Article
3-Hydroxy-3-Methylglutaric Acid Disrupts Brain Bioenergetics, Redox Homeostasis, and Mitochondrial Dynamics and Affects Neurodevelopment in Neonatal Wistar Rats
by Josyane de Andrade Silveira, Manuela Bianchin Marcuzzo, Jaqueline Santana da Rosa, Nathalia Simon Kist, Chrístofer Ian Hernandez Hoffmann, Andrey Soares Carvalho, Rafael Teixeira Ribeiro, André Quincozes-Santos, Carlos Alexandre Netto, Moacir Wajner and Guilhian Leipnitz
Biomedicines 2024, 12(7), 1563; https://doi.org/10.3390/biomedicines12071563 - 15 Jul 2024
Viewed by 688
Abstract
3-Hydroxy-3-methylglutaric acidemia (HMGA) is a neurometabolic inherited disorder characterized by the predominant accumulation of 3-hydroxy-3-methylglutaric acid (HMG) in the brain and biological fluids of patients. Symptoms often appear in the first year of life and include mainly neurological manifestations. The neuropathophysiology is not [...] Read more.
3-Hydroxy-3-methylglutaric acidemia (HMGA) is a neurometabolic inherited disorder characterized by the predominant accumulation of 3-hydroxy-3-methylglutaric acid (HMG) in the brain and biological fluids of patients. Symptoms often appear in the first year of life and include mainly neurological manifestations. The neuropathophysiology is not fully elucidated, so we investigated the effects of intracerebroventricular administration of HMG on redox and bioenergetic homeostasis in the cerebral cortex and striatum of neonatal rats. Neurodevelopment parameters were also evaluated. HMG decreased the activity of glutathione reductase (GR) and increased catalase (CAT) in the cerebral cortex. In the striatum, HMG reduced the activities of superoxide dismutase, glutathione peroxidase, CAT, GR, glutathione S-transferase, and glucose-6-phosphate dehydrogenase. Regarding bioenergetics, HMG decreased the activities of succinate dehydrogenase and respiratory chain complexes II–III and IV in the cortex. HMG also decreased the activities of citrate synthase and succinate dehydrogenase, as well as complex IV in the striatum. HMG further increased DRP1 levels in the cortex, indicating mitochondrial fission. Finally, we found that the HMG-injected animals showed impaired performance in all sensorimotor tests examined. Our findings provide evidence that HMG causes oxidative stress, bioenergetic dysfunction, and neurodevelopmental changes in neonatal rats, which may explain the neuropathophysiology of HMGA. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of 3-hydroxy-3-methyglutaric acid (HMG) on antioxidant defenses in the neonatal rat cerebral cortex. Superoxide dismutase (SOD; (<b>A</b>)), catalase (CAT; (<b>B</b>)), glutathione peroxidase (GPx; (<b>C</b>)), glutathione S-transferase (GST; (<b>D</b>)), glutathione reductase (GR; (<b>E</b>)), and glucose-6-phosphate dehydrogenase (G6PDH; (<b>F</b>)) activities and reduced glutathione (GSH) levels (<b>G</b>) were measured. Values are means ± SD (<span class="html-italic">n</span> = 5). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to rats receiving PBS (control group) (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>Effects of 3-hydroxy-3-methyglutaric acid (HMG) on antioxidant defenses in the neonatal rat striatum. Superoxide dismutase (SOD; (<b>A</b>)), catalase (CAT; (<b>B</b>)), glutathione peroxidase (GPx; (<b>C</b>)), glutathione S-transferase (GST; (<b>D</b>)), glutathione reductase (GR; (<b>E</b>)), and glucose-6-phosphate dehydrogenase (G6PDH; (<b>F</b>)) activities and reduced glutathione (GSH) levels (<b>G</b>) were measured. Values are means ± SD (<span class="html-italic">n</span> = 5). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, compared to rats receiving PBS (control group) (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>Effects of 3-hydroxy-3-methyglutaric acid (HMG) on the activities of citric acid cycle enzymes and mitochondrial respiratory chain complexes in the neonatal rat cerebral cortex. The activities of the citric acid cycle enzymes citrate synthase (CS; (<b>A</b>)), succinate dehydrogenase (SDH; (<b>B</b>)), and malate dehydrogenase (MDH; (<b>C</b>)), as well as respiratory chain complexes II (CII; (<b>D</b>)), II–III (CII-III; (<b>E</b>)) and IV (CIV; (<b>F</b>)), were measured. Values are means ± SD (<span class="html-italic">n</span> = 5). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to rats receiving PBS (control group) (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p>Effects of 3-hydroxy-3-methyglutaric acid (HMG) on the activities of citric acid cycle enzymes and mitochondrial respiratory chain complexes in the neonatal rat striatum. The activities of the citric acid cycle enzymes citrate synthase (CS; (<b>A</b>)), succinate dehydrogenase (SDH; (<b>B</b>)), and malate dehydrogenase (MDH; (<b>C</b>)), as well as the respiratory chain complexes II (CII; (<b>D</b>)), II–III (CII-III; (<b>E</b>)), and IV (CIV; (<b>F</b>)), were measured. Values are means ± SD (<span class="html-italic">n</span> = 5). ** <span class="html-italic">p</span> &lt; 0.01, compared to rats receiving PBS (control group) (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p>Effects of 3-hydroxy-3-methyglutaric acid (HMG) on the content of DRP1 (<b>A</b>), MFN1 (<b>B</b>), and OPA1 (<b>C</b>) in the neonatal rat cerebral cortex. Representative blots are shown at the top and quantification is at the bottom. Values are means ± SD (<span class="html-italic">n</span> = 6). *** <span class="html-italic">p</span> &lt; 0.001, compared to rats receiving PBS (control group) (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Effects of 3-hydroxy-3-methyglutaric acid (HMG) on neurodevelopmental parameters in neonatal rats. Negative geotaxis (<b>A</b>), righting (<b>B</b>), gait (<b>C</b>), hindlimb suspension (<b>D</b>), cliff avoidance (<b>E</b>), and forelimb grasping (<b>F</b>) were evaluated on postnatal day 9. Values are means ± SD (<span class="html-italic">n</span> = 5–6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to rats receiving PBS (control group) (Student’s <span class="html-italic">t</span>-test or Mann–Whitney test).</p>
Full article ">
16 pages, 8620 KiB  
Article
Peanut Shell Extract Improves Mitochondrial Function in db/db Mice via Suppression of Oxidative Stress and Inflammation
by Hemalata Deshmukh, Julianna M. Santos, Matthew Bender, Jannette M. Dufour, Jacob Lovett and Chwan-Li Shen
Nutrients 2024, 16(13), 1977; https://doi.org/10.3390/nu16131977 - 21 Jun 2024
Viewed by 819
Abstract
Accumulating evidence shows a strong correlation between type 2 diabetes mellitus, mitochondrial dysfunction, and oxidative stress. We evaluated the effects of dietary peanut shell extract (PSE) supplementation on mitochondrial function and antioxidative stress/inflammation markers in diabetic mice. Fourteen db/db mice were randomly assigned [...] Read more.
Accumulating evidence shows a strong correlation between type 2 diabetes mellitus, mitochondrial dysfunction, and oxidative stress. We evaluated the effects of dietary peanut shell extract (PSE) supplementation on mitochondrial function and antioxidative stress/inflammation markers in diabetic mice. Fourteen db/db mice were randomly assigned to a diabetic group (DM in AIN-93G diet) and a PSE group (1% wt/wt PSE in AIN-93G diet) for 5 weeks. Six C57BL/6J mice were fed with an AIN-93G diet for 5 weeks (control group). Gene and protein expression in the liver, brain, and white adipose tissue (WAT) were determined using qRT-PCR and Immunoblot, respectively. Compared to the control group, the DM group had (i) increased gene and protein expression levels of DRP1 (fission), PINK1 (mitophagy), and TNFα (inflammation) and (ii) decreased gene and protein expression levels of MFN1, MFN2, OPA1 (fusion), TFAM, PGC-1α (biogenesis), NRF2 (antioxidative stress) and IBA1 (microglial activation) in the liver, brain, and WAT of db/db mice. Supplementation of PSE into the diet restored the DM-induced changes in the gene and protein expression of DRP1, PINK1, TNFα, MFN1, MFN2, OPA1, TFAM, PGC-1α, NRF2, and IBA1 in the liver, brain, and WAT of db/db mice. This study demonstrates that PSE supplementation improved mitochondrial function in the brain, liver, and WAT of db/db mice, in part due to suppression of oxidative stress and inflammation. Full article
(This article belongs to the Special Issue Dietary Manipulations: Advances in Metabolism Disease)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The impact of PSE on DRP1 gene and protein expression in the liver, brain, and WAT of mice. The data are expressed as mean ± SEM. There were 6–7 samples per group (<span class="html-italic">n</span> = 6–7). Statistical analysis was conducted using one-way ANOVA followed by Uncorrected Fisher’s LSD test using GraphPad Prism 9. Data are presented as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005 and # 0.05 &lt; <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 2
<p>The influence of PSE on FIS1 gene/protein expression in the liver, brain, and WAT of mice. The data are expressed as mean ± SEM. <span class="html-italic">n</span> = 6–7 per group. Statistical analysis was conducted with one-way ANOVA and Uncorrected Fisher’s LSD test using GraphPad Prism 9. The data are presented as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005, and **** <span class="html-italic">p</span> &lt; 0.00005.</p>
Full article ">Figure 3
<p>Result of PSE administration in mice on f MFN1 gene as well as protein expression in the liver, brain, and WAT tissue. Data are stated as mean ± SEM. <span class="html-italic">n</span> = 6–7 per group. Statistics were carried out by one-way ANOVA followed by Uncorrected Fisher’s LSD using GraphPad Prism 9. The data are shown as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, and *** <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 4
<p>PSE supplementation modulates expression of MFN2 at gene and protein levels in different tissues of mice. Data are expressed as mean ± SEM. <span class="html-italic">n</span> = 6–7 per group. Data analysis was carried out by using one-way ANOVA followed by Uncorrected Fisher’s LSD with GraphPad Prism 9. Data is stated as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, and *** <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 5
<p>PSE administration showed an effect on the expression of OPA1 in the liver, brain, and WAT tissue of mice. The data are presented as mean ± SEM with 6–7 samples per group (<span class="html-italic">n</span> = 6–7). Statistical analysis was performed using one-way ANOVA followed by Uncorrected Fisher’s LSD test using GraphPad Prism 9. Significance levels are denoted as follows: * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.005, and *** for <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 6
<p>The impact of PSE on PGC-1α gene and protein expression in the liver, brain, and (WAT) of mice. Data are expressed as mean ± SEM with 6–7 samples per group (<span class="html-italic">n</span> = 6–7). Statistical analysis was conducted using one-way ANOVA followed by Uncorrected Fisher’s LSD test using GraphPad Prism 9. Significance levels are indicated as follows: * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.005, *** for <span class="html-italic">p</span> &lt; 0.0005, and # for 0.05 &lt; <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 7
<p>The impact of PSE on expression of TFAM at gene/protein level in different tissues of mice. Data are presented as mean ± SEM with 6–7 samples per group (<span class="html-italic">n</span> = 6–7). Statistical analysis was performed using one-way ANOVA and Uncorrected Fisher’s LSD test using GraphPad Prism 9. Significance levels are denoted as follows: * for <span class="html-italic">p</span> &lt; 0.05, ** for <span class="html-italic">p</span> &lt; 0.005, and *** for <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 8
<p>Effect of PSE on PINK1 gene and protein expression in the liver, brain, and WAT of mice. The data are presented as mean ± SEM, <span class="html-italic">n</span> = 6–7 per group. Statistical analysis was carried out by using one-way ANOVA followed by Uncorrected Fisher’s LSD with GraphPad Prism 9. The analysis is presented as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005, **** <span class="html-italic">p</span> &lt; 0.00005 and # 0.05 &lt; <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 9
<p>PSE administration in mice changes the NRF2 gene and protein expression in the liver, brain, and WAT tissues. The data are expressed as mean ± SEM. <span class="html-italic">n</span> = 6–7 per group. Statistical significance was analyzed by one-way ANOVA followed by Uncorrected Fisher’s LSD using GraphPad Prism 9. The data is expressed as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, and # 0.05 &lt; <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 10
<p>The effect of PSE on TNFα gene and protein expression in different tissues of mice. The data are expressed as mean ± SEM. <span class="html-italic">n</span> = 6–7 per group. Statistical analysis was carried out by one-way ANOVA followed by Uncorrected Fisher’s LSD with GraphPad Prism 9. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005, and **** <span class="html-italic">p</span> &lt; 0.00005.</p>
Full article ">Figure 11
<p>PSE administration in mice leads to alteration in gene and protein expression of IBA1 in various tissues. The data are expressed as mean ± SEM. <span class="html-italic">n</span> = 6–7 per group. The data were analyzed by one-way ANOVA followed by Uncorrected Fisher’s LSD using GraphPad Prism 9. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005, **** <span class="html-italic">p</span> &lt; 0.00005, and # 0.05 &lt; <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">
20 pages, 534 KiB  
Article
An Energy-Efficient Logistic Drone Routing Method Considering Dynamic Drone Speed and Payload
by Kunpeng Wu, Shaofeng Lu, Haoqin Chen, Minling Feng and Zenghao Lu
Sustainability 2024, 16(12), 4995; https://doi.org/10.3390/su16124995 - 12 Jun 2024
Viewed by 699
Abstract
Unmanned aerial vehicles (UAVs), or drones, are recognized for their potential to improve efficiency in last-mile delivery. Unlike the vehicle routing problem, drone route design is challenging due to several operational signatures, such as speed optimization, multi-trip operation, and energy consumption estimation. Drone [...] Read more.
Unmanned aerial vehicles (UAVs), or drones, are recognized for their potential to improve efficiency in last-mile delivery. Unlike the vehicle routing problem, drone route design is challenging due to several operational signatures, such as speed optimization, multi-trip operation, and energy consumption estimation. Drone energy consumption is a nonlinear function of both speed and payload. Moreover, the high speed of drones can significantly curtail the drone range, thereby limiting the efficiency of drone delivery systems. This paper addresses the trade-off between speed and flight range in a multi-trip drone routing problem with variable flight speeds (DRP–VFS). We propose a new model to specifically consider energy constraints using a nonlinear energy consumption model and treat drone speeds as decision variables. The DRP–VFS is initially formulated using mixed-integer linear programming (MILP) to minimize energy consumption. To solve large-scale instances, we propose a three-phase adaptive large neighborhood search (ALNS) algorithm and compare its performance with a commercial MIP solver. The experimental results demonstrate that the proposed method is capable of effectively identifying suboptimal solutions in practical scenarios. Furthermore, results indicate that operating drones at variable speeds leads to about 21% energy savings compared to fixed speeds, with advantages in cost savings and range extension. Full article
(This article belongs to the Section Sustainable Engineering and Science)
Show Figures

Figure 1

Figure 1
<p>The required power, energy consumption per unit distance, endurance, and range of a drone versus speed and payload are recalculated based on the drone energy model proposed by [<a href="#B17-sustainability-16-04995" class="html-bibr">17</a>]. The circles in subfigure (<b>a</b>) and (<b>c</b>) indicate minimum required power and maximum endurance under fixed battery capacity for different speed and payload, respectively. The triangles in subfigure (<b>b</b>) and (<b>d</b>) indicate minimum energy consumption and maximum flight range under fixed battery capacity for different speed and payload, respectively.</p>
Full article ">Figure 2
<p>Example of ALNS heuristic on a DRP–VFS solution. The numbers indicate customers to be served by drones. The red, green and blue lines represent three different drone routes. Focusing on the red route, subfigure (<b>a</b>) shows that a drone sequentially serves customers <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>9</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>11</mn> <mo>,</mo> <mn>12</mn> <mo>]</mo> </mrow> </semantics></math> at the begining. Subfigure (<b>b</b>) shows the customer 9 is removed from the red route during destroy operation. Subfigure (<b>c</b>) shows the customer 1 and 8 are added into the red route during repair operation.</p>
Full article ">
16 pages, 1307 KiB  
Article
Medication Risks and Their Association with Patient-Reported Outcomes in Inpatients with Cancer
by Maximilian Günther, Markus Schuler, Leopold Hentschel, Hanna Salm, Marie-Therese Schmitz and Ulrich Jaehde
Cancers 2024, 16(11), 2110; https://doi.org/10.3390/cancers16112110 - 31 May 2024
Viewed by 688
Abstract
Background: We aimed to assess medication risks and determine factors influencing the health-related quality of life (HRQOL) in cancer inpatients. Methods: A retrospective analysis was conducted to identify drug-related problems (DRPs) based on medication reviews, including patient-reported outcomes (PROs). Multiple linear regression analyses [...] Read more.
Background: We aimed to assess medication risks and determine factors influencing the health-related quality of life (HRQOL) in cancer inpatients. Methods: A retrospective analysis was conducted to identify drug-related problems (DRPs) based on medication reviews, including patient-reported outcomes (PROs). Multiple linear regression analyses were performed to identify sociodemographic, disease-related, and drug therapy-related factors influencing changes from hospital admission to discharge in the scales of the EORTC QLQ-C30 questionnaire. Results: A total of 162 inpatients with various hematological and solid cancer diseases was analyzed. Patients received a mean of 11.6 drugs and 92.6% of patients exhibited polymedication resulting in a mean of 4.0 DRPs per patient. Based on PRO data, 21.5% of DRPs were identified. Multiple linear regression models described the variance of the changes in global HRQOL and physical function in a weak-to-moderate way. While drug therapy-related factors had no influence, relapse status and duration of hospital stay were identified as significant covariates for global HRQOL and physical function, respectively. Conclusion: This analysis describes underlying DRPs in a German cancer inpatient population. PROs provided valuable information for performing medication reviews. The multiple linear regression models for global HRQOL and physical function provided explanations for changes during hospital stay. Full article
Show Figures

Figure 1

Figure 1
<p>Study framework.</p>
Full article ">Figure 2
<p>Time-points of documentation of the clinical trial. Visit 1 was only performed for patients with a hospital stay of at least seven days.</p>
Full article ">Figure 3
<p>Drug classes categorized by Anatomical Therapeutical Chemical (ATC) code level 1 (n = 1884).</p>
Full article ">Figure 4
<p>Histograms of the distribution of the score changes from baseline to hospital discharge in global health-related quality of life ((<b>A</b>) n = 153) and physical function ((<b>B</b>) n = 152).</p>
Full article ">
70 pages, 4695 KiB  
Review
Numerical Dissipation Control in High-Order Methods for Compressible Turbulence: Recent Development
by H. C. Yee and Björn Sjögreen
Fluids 2024, 9(6), 127; https://doi.org/10.3390/fluids9060127 - 29 May 2024
Viewed by 673
Abstract
This comprehensive overview presents our continued efforts in high-order finite difference method (FDM) development for adaptive numerical dissipation control in the long-time integration of direct numerical simulation (DNS), large eddy simulation (LES), and implicit LES (ILES) computations of compressible turbulence for gas dynamics [...] Read more.
This comprehensive overview presents our continued efforts in high-order finite difference method (FDM) development for adaptive numerical dissipation control in the long-time integration of direct numerical simulation (DNS), large eddy simulation (LES), and implicit LES (ILES) computations of compressible turbulence for gas dynamics and MHD. The focus is on turbulence with shock wave numerical simulations using the adaptive blending of high-order structure-preserving non-dissipative methods (classical central, Padé (compact), and dispersion relation-preserving (DRP)) with high-order shock-capturing methods in such a way that high-order shock-capturing methods are active only in the vicinity of shock/shear waves, and high-gradient and spurious high-frequency oscillation regions guided via flow sensors. Any efficient and high-resolution high-order shock-capturing methods are good candidates for the blending of methods procedure. Typically, the adaptive blending of more than one method falls under two camps: hybrid methods and nonlinear filter methods. They are applicable to unstructured finite volume, finite element, discontinuous Galerkin, and spectral element methods. This work represents the culmination of over 20 years of high-order FDM developments and hands-on experience by the authors and collaborators in adaptive numerical dissipation control using the “high order nonlinear filter approach”. Extensions of these FDM versions to curvilinear nonuniform, freestream-preserving moving grids and time-varying deforming grids were also developed. By examining the construction of these two approaches using the high-order multistage type of temporal discretization, the nonlinear filter approach is made more efficient and less CPU-intensive while obtaining similar accuracy. A representative variety of test cases that compare the various blending of high-order methods with standalone standard methods is illustrated. Due to the fact that our nonlinear filter methods are not well known in compressible turbulence with shock waves, the intent of this comprehensive overview is for general audiences who are not familiar with our nonlinear filter methods. For readers interested in the implementation of our methods into their computer code, it is hoped that the long overview will be helpful. Full article
(This article belongs to the Special Issue Next-Generation Methods for Turbulent Flows)
Show Figures

Figure 1

Figure 1
<p>2D Alf<math display="inline"><semantics> <mrow> <mover> <mi mathvariant="normal">v</mi> <mo>́</mo> </mover> </mrow> </semantics></math>en waves’ MHD simulation problem setup.</p>
Full article ">Figure 2
<p>2D Alf<math display="inline"><semantics> <mrow> <mover> <mi mathvariant="normal">v</mi> <mo>́</mo> </mover> </mrow> </semantics></math>en waves MHD simulation: Maximum error time evolution comparison. The comparisons include six methods. Between Padé vs. classical central and Padé vs. classical central on the entropy split form of the inviscid flux derivative with two different entropy split parameters, <math display="inline"><semantics> <mi>β</mi> </semantics></math>. “Ce” denotes the eighth-order central, and “Co” denotes the eighth-order Padé. <math display="inline"><semantics> <mi>β</mi> </semantics></math> denotes the magnetic field vector. “CeES” denotes the eighth-order central, and “CoES” denotes the eighth-order Padé in the entropy-splitting form of the inviscid flux derivative.</p>
Full article ">Figure 3
<p>Nonlinear filter procedure.</p>
Full article ">Figure 4
<p>Nonlinear filter procedure with local flow sensors.</p>
Full article ">Figure 5
<p>Accuracy and CPU comparison between a hybrid method and a nonlinear filter method using the same non-dissipative high-order linear spatial discretization blending with the same shock-capturing method.</p>
Full article ">Figure 6
<p>3D isotropic turbulence with shocklets problem setup.</p>
Full article ">Figure 7
<p>3D isotropic turbulence with shocklets: comparison between the standard seventh-order WENO method (WENO7) and our two seventh-order nonlinear filter methods (C08Econs + WENO7fi and C08Dsplit+WENO7fi) using a very coarse <math display="inline"><semantics> <msup> <mn>64</mn> <mn>3</mn> </msup> </semantics></math> grid with the filtered DNS computation on a <math display="inline"><semantics> <msup> <mn>256</mn> <mn>3</mn> </msup> </semantics></math> grid. Kinetic energy (<b>top left</b>)), enstrophy (<b>top right</b>), temperature variance (<b>bottom left</b>), and dilatation (<b>bottom right</b>). C08Econs denotes the eighth-order classical central applied to the entropy split form, and C08Dsplit denotes the eighth-order classical central applied to the Ducros et al. split form of the Euler inviscid flux derivative.</p>
Full article ">Figure 8
<p>Schematic of Strang operator splitting on the homogeneous portion and the source term portion of the governing equations.</p>
Full article ">Figure 9
<p>Subcell resolution method in solving equations containing nonlinear source terms.</p>
Full article ">Figure 10
<p>Subcell resolution using three staggered steps.</p>
Full article ">Figure 11
<p>A 1D C-J detonation problem setup.</p>
Full article ">Figure 12
<p>1D C-J detonation comparison: Pressure (<b>left</b>) and mass fraction of unburnt gas (<b>right</b>). Comparison among four high-order shock-capturing methods for a C-J detonation problem. One reaction, Arrhenius model, using 50 uniformly distributed grid points.</p>
Full article ">Figure 13
<p>A 2D C-J detonation problem setup.</p>
Full article ">Figure 14
<p>2D C-J detonation comparison: 1D cross-section of density at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>1.7</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>7</mn> </mrow> </msup> </mrow> </semantics></math> via the same four methods as the 1D case on a uniform coarse grid of <math display="inline"><semantics> <mrow> <mn>200</mn> <mo>×</mo> <mn>40</mn> </mrow> </semantics></math>. The CFL = <math display="inline"><semantics> <mrow> <mn>0.05</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.5825</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>10</mn> </msup> </mrow> </semantics></math>. The right figure is a close-up of the vicinity of the discontinuity. The reference solution is achieved via WENO5 using <math display="inline"><semantics> <mrow> <mn>4000</mn> <mo>×</mo> <mn>800</mn> </mrow> </semantics></math> uniform grid points.</p>
Full article ">Figure 15
<p>1D and 2D C-J detonation CPU comparison among methods.</p>
Full article ">Figure 16
<p>The EAST experiment.</p>
Full article ">Figure 17
<p>One-dimensional 13-species problem setup related to the EAST experiment.</p>
Full article ">Figure 18
<p>3D 13-species problem setup related to the EAST experiment.</p>
Full article ">Figure 19
<p>Temperature comparison using three grids and comparison among five high-order shock-capturing methods for 1D 13-species chemical reacting flows.</p>
Full article ">Figure 20
<p>Temperature and pressure evolution of the 2D 13-species chemical reacting flows.</p>
Full article ">Figure 21
<p>Temperature comparison among three high-order shock-capturing methods for the same 2D 13-species chemical reacting flows.</p>
Full article ">Figure 22
<p>3D supersonic shock–turbulence Interaction test case.</p>
Full article ">Figure 23
<p>Instantaneous velocity field, <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math> (<b>top</b>) and <math display="inline"><semantics> <msub> <mi>u</mi> <mi>y</mi> </msub> </semantics></math> (<b>bottom</b>), obtained with DNS on grid of <math display="inline"><semantics> <mrow> <mn>1553</mn> <mo>×</mo> <msup> <mn>256</mn> <mn>2</mn> </msup> </mrow> </semantics></math> points. Slice <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">s</mi> <mi mathvariant="normal">t</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 24
<p>2D isentropic vortex convection problem setup.</p>
Full article ">Figure 25
<p>2D inviscid isentropic vortex convection: comparison of maximum norm error vs. time for different <math display="inline"><semantics> <mi>β</mi> </semantics></math> by eighth-order ES (<b>top</b>) and ESSW (<b>bottom</b>) using <math display="inline"><semantics> <msup> <mn>101</mn> <mn>2</mn> </msup> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <msup> <mn>201</mn> <mn>2</mn> </msup> </semantics></math> (<b>right</b>) grid points.</p>
Full article ">Figure 26
<p>2D inviscid isentropic vortex convection: comparison of maximum norm error, mass conservation error, entropy errors, and kinetic energy by <math display="inline"><semantics> <msub> <mi>E</mi> <mi>H</mi> </msub> </semantics></math> vs. time for the nine eighth-order methods using fine <math display="inline"><semantics> <msup> <mn>201</mn> <mn>2</mn> </msup> </semantics></math> grid points and <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 27
<p>2D inviscid isentropic vortex convection: final end time of <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1440</mn> </mrow> </semantics></math> comparison of maximum norm error, entropy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>H</mi> </msub> </semantics></math>, entropy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>L</mi> </msub> </semantics></math>, and kinetic energy vs. time for the eight eighth-order methods using <math display="inline"><semantics> <msup> <mn>100</mn> <mn>2</mn> </msup> </semantics></math> grid points and <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 28
<p>3D inviscid Taylor–Green vortex using <math display="inline"><semantics> <msup> <mn>64</mn> <mn>3</mn> </msup> </semantics></math> grid points: comparison of kinetic energy (<b>top left</b>), enstrophy (<b>top right</b>), entropy (<b>bottom left</b>), and entropy (close-up, <b>bottom right</b>) vs. time for the nine eighth-order methods using <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 29
<p>3D inviscid Taylor–Green vortex using <math display="inline"><semantics> <msup> <mn>64</mn> <mn>3</mn> </msup> </semantics></math> grid points: comparison of mass conservation (<b>top left</b>), x-momentum conservation (<b>top right</b>), y-momentum conservation (<b>bottom left</b>), and z-momentum conservation (<b>bottom right</b>) vs. time for the nine eighth-order methods using <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 30
<p>Brio and Wu MHD shock-tube problem setup.</p>
Full article ">Figure 31
<p>Brio and Wu MHD shock-tube: comparison among seven structure-preserving eighth-order central base schemes using the same WENO7 as the filter.</p>
Full article ">Figure 32
<p>Maximum norm error vs. time. Comparison of new and old ES schemes for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2.5</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>25.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 33
<p>Wave configuration of the Brio–Wu–Riemann problem.</p>
Full article ">Figure 34
<p>Brio–Wu–Riemann problem: Density computed via old and new entropy split schemes with <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mn>2.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>25.5</mn> </mrow> </semantics></math>. Bottom figure is close-up of top figure.</p>
Full article ">Figure 35
<p>Brio–Wu–Riemann problem. <math display="inline"><semantics> <msub> <mi>B</mi> <mn>2</mn> </msub> </semantics></math> computed via old and new entropy split schemes with <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mn>2.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>25.5</mn> </mrow> </semantics></math>. Middle and bottom figures are close-ups of top figure.</p>
Full article ">Figure 36
<p>Brio–Wu–Riemann problem. Pressure computed via old and new entropy split schemes with <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mn>2.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>25.5</mn> </mrow> </semantics></math>. Bottom figure is close-up of top figure.</p>
Full article ">Figure 37
<p>Solution at time 1.8, density (<b>top</b>), pressure (<b>middle</b>), and <span class="html-italic">y</span>-magnetic field (<b>bottom</b>).</p>
Full article ">Figure 38
<p>Solution at time 1.8, close-ups of oscillatory region for density (<b>top</b>), pressure (<b>middle</b>), and <span class="html-italic">y</span>-magnetic field (<b>bottom</b>).</p>
Full article ">Figure 39
<p>Conserved quantity vs. time for ESnew (red), ESSWnew (black), and ECH (blue). From top to bottom, density, <span class="html-italic">x</span>-momentum, and total energy.</p>
Full article ">
20 pages, 4584 KiB  
Article
Liraglutide Pretreatment Does Not Improve Acute Doxorubicin-Induced Cardiotoxicity in Rats
by Carolina R. Tonon, Marina G. Monte, Paola S. Balin, Anderson S. S. Fujimori, Ana Paula D. Ribeiro, Natália F. Ferreira, Nayane M. Vieira, Ronny P. Cabral, Marina P. Okoshi, Katashi Okoshi, Leonardo A. M. Zornoff, Marcos F. Minicucci, Sergio A. R. Paiva, Mariana J. Gomes and Bertha F. Polegato
Int. J. Mol. Sci. 2024, 25(11), 5833; https://doi.org/10.3390/ijms25115833 - 27 May 2024
Viewed by 850
Abstract
Doxorubicin is an effective drug for cancer treatment; however, cardiotoxicity limits its use. Cardiotoxicity pathophysiology is multifactorial. GLP-1 analogues have been shown to reduce oxidative stress and inflammation. In this study, we evaluated the effect of pretreatment with liraglutide on doxorubicin-induced acute cardiotoxicity. [...] Read more.
Doxorubicin is an effective drug for cancer treatment; however, cardiotoxicity limits its use. Cardiotoxicity pathophysiology is multifactorial. GLP-1 analogues have been shown to reduce oxidative stress and inflammation. In this study, we evaluated the effect of pretreatment with liraglutide on doxorubicin-induced acute cardiotoxicity. A total of 60 male Wistar rats were allocated into four groups: Control (C), Doxorubicin (D), Liraglutide (L), and Doxorubicin + Liraglutide (DL). L and DL received subcutaneous injection of liraglutide 0.6 mg/kg daily, while C and D received saline for 2 weeks. Afterwards, D and DL received a single intraperitoneal injection of doxorubicin 20 mg/kg; C and L received an injection of saline. Forty-eight hours after doxorubicin administration, the rats were subjected to echocardiogram, isolated heart functional study, and euthanasia. Liraglutide-treated rats ingested significantly less food and gained less body weight than animals that did not receive the drug. Rats lost weight after doxorubicin injection. At echocardiogram and isolated heart study, doxorubicin-treated rats had systolic and diastolic function impairment. Myocardial catalase activity was statistically higher in doxorubicin-treated rats. Myocardial protein expression of tumor necrosis factor alpha (TNF-α), phosphorylated nuclear factor-κB (p-NFκB), troponin T, and B-cell lymphoma 2 (Bcl-2) was significantly lower, and the total NFκB/p-NFκB ratio and TLR-4 higher in doxorubicin-treated rats. Myocardial expression of OPA-1, MFN-2, DRP-1, and topoisomerase 2β did not differ between groups (p > 0.05). In conclusion, doxorubicin-induced cardiotoxicity is accompanied by decreased Bcl-2 and phosphorylated NFκB and increased catalase activity and TLR-4 expression. Liraglutide failed to improve acute doxorubicin-induced cardiotoxicity in rats. Full article
(This article belongs to the Special Issue Heart Failure Mechanisms and Treatment Advances)
Show Figures

Figure 1

Figure 1
<p>Body weight and food ingestion. The area under the curve (AUC) was compared between the groups in the periods D1–D7, D7–D12, and D12–D14 for body weight and D1–D9 and D9–D14 for food ingestion. C: control (<span class="html-italic">n</span> = 15); L: liraglutide (<span class="html-italic">n</span> = 15); D: doxorubicin (<span class="html-italic">n</span> = 15); DL: doxorubicin + liraglutide (<span class="html-italic">n</span> = 15). Data are expressed as means ± standard error. Generalized linear model (GLM); pD: <span class="html-italic">p</span>-value for doxorubicin effect; pL: <span class="html-italic">p</span>-value for liraglutide effect; pDxL: <span class="html-italic">p</span>-value for the interaction between doxorubicin and liraglutide.</p>
Full article ">Figure 2
<p>Echocardiographic structural and functional variables. C: control (<span class="html-italic">n</span> = 15); L: liraglutide (<span class="html-italic">n</span> = 15); D: doxorubicin (<span class="html-italic">n</span> = 15); DL: doxorubicin + liraglutide (<span class="html-italic">n</span> = 15); LVDD: left ventricle diastolic diameter; LAD: left atrial diameter; Aorta: aorta dimension; E: peak velocity of early ventricular filling; A: peak velocity of transmitral flow during atrial contraction; PWSV: posterior wall shortening velocity; IVRT: isovolumetric relaxation time; E’: average of mitral ring displacement of the lateral and septal walls during initial diastole in tissue Doppler image (TDI); A’: average of mitral ring displacement of the lateral and septal wall during late diastole in TDI; S’: average of mitral ring displacement of the lateral and septal wall during systole in TDI. Data are expressed as means ± standard error. Generalized linear model (GLM) or ANCOVA; * <span class="html-italic">p</span> &lt; 0.05 for dox factor; % <span class="html-italic">p</span> &lt; 0.05 vs. C; &amp; <span class="html-italic">p</span> &lt; 0.05 vs. L; # <span class="html-italic">p</span> &lt; 0.05 vs. D.</p>
Full article ">Figure 3
<p>Isolated heart preparation. C: control (<span class="html-italic">n</span> = 5); L: liraglutide (<span class="html-italic">n</span> = 6); D: doxorubicin (<span class="html-italic">n</span> = 6); DL: doxorubicin + liraglutide (<span class="html-italic">n</span> = 5). DP: maximum developed systolic pressure; +dP/dt: maximum left ventricular (LV) pressure development rate; −dP/dt: maximum LV pressure decrease rate. Data are expressed as means ± standard error. Generalized linear model (GLM). * <span class="html-italic">p</span> &lt; 0.05 for dox factor.</p>
Full article ">Figure 4
<p>Diastolic pressure–volume curve and stress–strain relationship (AUC). C: control; D: doxorubicin; L: liraglutide; DL: doxorubicin + liraglutide. Data are expressed as means. Generalized linear model (GLM); pD: <span class="html-italic">p</span>-value for doxorubicin effect; <span class="html-italic">p</span>-value for liraglutide effect &gt; 0.05; <span class="html-italic">p</span>-value for the interaction between doxorubicin and liraglutide &gt; 0.05.</p>
Full article ">Figure 5
<p>Myocardial histology. Representative longitudinal sections of the myocardium stained with hematoxylin and eosin. C: control (<span class="html-italic">n</span> = 7); L: liraglutide (<span class="html-italic">n</span> = 8); D: doxorubicin (<span class="html-italic">n</span> = 8); DL: doxorubicin + liraglutide (<span class="html-italic">n</span> = 8).</p>
Full article ">Figure 6
<p>In situ oxidative stress evaluation. C: control (<span class="html-italic">n</span> = 5); L: liraglutide (<span class="html-italic">n</span> = 4); D: doxorubicin (<span class="html-italic">n</span> = 5); DL: doxorubicin + liraglutide (<span class="html-italic">n</span> = 5). (<b>A</b>) Representative myocardial sections stained with dihydroethidium to identify reactive oxygen species in the nuclei. The analysis was performed under a fluorescence microscope at 40× magnification. (<b>B</b>) Integrated density of DHE staining. Data are expressed as means ± standard error.</p>
Full article ">Figure 7
<p>Protein expression by Western blot. Sample size: 5–7. TNF-α: tumor necrosis factor alpha; NFκB: nuclear factor kappa B; p-NFκB: phosphorylated nuclear factor kappa B; TLR-4: toll-like receptor type 4; BCL-2: B2 cell lymphoma; GAPDH: glyceraldehyde 3-phosphate dehydrogenase; C: control; D: doxorubicin; L: liraglutide; DL: doxorubicin + liraglutide; AU: arbitrary unit. Proteins were normalized by GAPDH, except for NFκB and p-NFκB, which were normalized by endogenous total protein using Ponceau staining. IL-10 and TLR4 were performed in the same membrane. Data are expressed as means ± standard error. Generalized linear model (GLM); * <span class="html-italic">p</span> &lt; 0.05 for dox factor.</p>
Full article ">Figure 8
<p>Experimental design. C: control; D: doxorubicin; L: liraglutide; DL: doxorubicin + liraglutide; SC: subcutaneous; IP: intraperitoneal; Dox: doxorubicin; WB: Western blot.</p>
Full article ">
21 pages, 10469 KiB  
Article
Textures and Chemical Compositions of Magnetite from Zhibo Submarine Volcanic Iron Oxide Deposit, Xinjiang, China: Implications for Re-Equilibration Processes
by Yang Wu, Ping Shen, Haoxuan Feng, Changhao Li, Jiayu Zhao, Yaoqing Luo and Wenguang Li
Minerals 2024, 14(6), 548; https://doi.org/10.3390/min14060548 - 26 May 2024
Viewed by 556
Abstract
The Awulale Iron Metallogenic Belt (AIMB) has developed many medium–large iron deposits, of which the Zhibo iron deposit is selected as the research object in this paper. The Zhibo deposit’s ore primarily consists of magnetite as the main mineral, accompanied by extensive epidotization. [...] Read more.
The Awulale Iron Metallogenic Belt (AIMB) has developed many medium–large iron deposits, of which the Zhibo iron deposit is selected as the research object in this paper. The Zhibo deposit’s ore primarily consists of magnetite as the main mineral, accompanied by extensive epidotization. The mineral assemblage includes diopside, albite, actinolite, epidote, chlorite, K-feldspar, quartz, calcite, chalcopyrite, and pyrite. Magnetite is classified into two groups based on sulfide content and mineral assemblage (MagI for sulfide-free and MagII for sulfide-rich ores). Two-stage mineralization of magnetite has been identified based on mineral assemblages and paragenesis, including the magmatic stage MagI and hydrothermal stage MagII. Mag I shows inhomogeneous backscattered electron (BSE) textures and consists of BSE-light and -dark domains (Mag I-L and MagI-D). Seven subtypes of magnetite have been recognized in this deposit. MagI-L and MagI-D have formed in the magmatic stage and show BSE images in light and dark colors, respectively. MagI-L is anhedral to subhedral and is inclusion-free. MagI-D has mainly replaced MagI-L along fractures and contains inclusions and pores. MagII has formed in the hydrothermal stage and is characterized by coupled dissolution–reprecipitation (DRP) textures. It can be divided into five sub-generations, that is, MagII-1, MagII-2, MagII-3, MagII-L, and MagII-D. MagII-1, MagII-2, and MagII-3 comprise the core–mantle–rim texture, while MagII-L and MagII-D comprise the core–rim texture. MagII-1 is BSE-light and is enriched with inclusions and pores. MagII-2 has partly replaced MagII-1 and exhibits oscillatory zoning under BSE imaging. It also contains inclusions. BSE-light MagII-3 occurs as overgrowth along MagII-2 margins and is inclusion-free. MagI magnetite is enriched with V, Cr, and Ni, whereas MagII is enriched with W, Ta, Nb, Sr, Sb, Sn, Y, Zr, Mg, Al, and Ti, indicating a decreased temperature of magnetite formation. MagI-L crystallizes from the original magma, while MagI-D is formed from the residual magma enriched with incompatible elements. MagII crystallizes from later multiple hydrothermal activities through the dissolution of early magnetite and the re-precipitation of later magnetite or from MagI-D which has later undergone a hydrothermal overprinting process. According to the texture and chemical composition of magnetite from the Zhibo deposit, we suggest that the Zhibo iron deposit was formed from the initial magmatic origin and then underwent a hydrothermal overprinting process. Full article
Show Figures

Figure 1

Figure 1
<p>Geological maps of (<b>a</b>) the western Tianshan metallogenic belt (modified from [<a href="#B8-minerals-14-00548" class="html-bibr">8</a>]) and (<b>b</b>) the eastern part of the AIMB (modified from [<a href="#B9-minerals-14-00548" class="html-bibr">9</a>]).</p>
Full article ">Figure 2
<p>A geological map of (<b>a</b>) the Zhibo iron deposit, (<b>b</b>) the eastern mining section, and (<b>c</b>) simplified geological sections along the C-C’ line (the figures are modified from [<a href="#B56-minerals-14-00548" class="html-bibr">56</a>]).</p>
Full article ">Figure 3
<p>Hand specimens and photomicrographs of wall rocks from Zhibo deposit. (<b>a</b>) Gray andesite is located in upper part of orebodies. (<b>b</b>) Grey–green andesite containing plagioclase phenocrysts. (<b>c</b>) Hand sample of granodiorite. (<b>d</b>) Plagioclase phenocrysts in andesite (cross-polarized light). (<b>e</b>) Andesite contains plagioclase and clinopyroxene phenocrysts (cross-polarized light). (<b>f</b>) Clinopyroxene replaced by epidote and magnetite is present in andesite (plane-polarized light). (<b>g</b>) Plagioclase and clinopyroxene phenocrysts and magnetite are present in andesite, accompanying epidotization alteration (cross-polarized light). Mag = magnetite; Pl = plagioclase; Ep = epidote; Cpx = clinopyroxene.</p>
Full article ">Figure 4
<p>Photographs of outcrops and hand specimens from the Zhibo iron deposit. (<b>a</b>) The ore body is in contact with the epidote alteration; (<b>b</b>) a massive ore is intercalated with andesite; (<b>c</b>) a massive magnetite ore; (<b>d</b>) a massive magnetite ore crosscut by an epidote–quartz–calcite vein; (<b>e</b>) a massive magnetite ore with dendritic magnetite inclusions; (<b>f</b>) a disseminated magnetite ore; (<b>g</b>) a complex breccia magnetite ore; (<b>h</b>) a complex breccia magnetite ore; (<b>i</b>) a banded magnetite ore; and (<b>j</b>) magnetite lava with flow texture. Act = actinolite; Cal = calcite; Di = diopside; Ep = epidote; Mag = magnetite; Py = pyrite; Qz = quartz.</p>
Full article ">Figure 5
<p>Photomicrographs from the Zhibo iron deposit. (<b>a</b>) Euhedral magnetite, with actinolite and albite filling the interstices between the magnetite crystals (plane-polarized light); (<b>b</b>) volcanic rock is replaced by an alteration assemblage of diopside and albite, with magnetite (plane-polarized light); (<b>c</b>) pyrite replaces magnetite (reflected light); (<b>d</b>) platy magnetite with granules of pyrite (reflected light); (<b>e</b>) subhedral magnetite grows with K-feldspar, epidote, and chlorite; (<b>f</b>) calcite is associated with quartz and chlorite (cross-polarized light). Ab = albite; Act = actinolite; Cal = calcite; Ccp = chalcopyrite; Chl = chlorite; Di = diopside; Ep = epidote; Mag = magnetite; Py = pyrite; Qz = quartz.</p>
Full article ">Figure 6
<p>Paragenesis of minerals and alteration in the Zhibo iron deposit (The long axis of the black ellipse indicates the duration, while the short axis indicates the relative amount of mineral formation).</p>
Full article ">Figure 7
<p>SEM-BSE images of magnetite. (<b>a</b>,<b>b</b>) MagI-L is subhedral–euhedral without inclusion. MagI-D replaces MagI-L magnetite along mineral fissures and contains inclusions and voids that vary from the faintly distributed micro- to nanometer scale; (<b>c</b>) MagII-D is dark magnetite with inclusion and voids which has been replaced by MagII-L, which grows in boundary integration contact of MagII-D magnetite; (<b>d</b>) fine-grained aggregated magnetite displays well-defined 120° triple junctions; (<b>e</b>) MagII-D with subhedral pyrite and MagII-L without inclusions; (<b>f</b>) core (MagII-1) is enriched with inclusions and voids. Core magnetite is replaced by MagII-2 with oscillatory zoning, and rim magnetite (MagII-3) grows along MagII-2.</p>
Full article ">Figure 8
<p>Box plots for magnetite EMPA.</p>
Full article ">Figure 9
<p>Radar plots of the upper threshold values, in parts per million, for the suite of important magnetite elements.</p>
Full article ">Figure 10
<p>Plot of Ti + V vs. Ca + Al + Mn (wt%) in magnetite. Reference fields from [<a href="#B37-minerals-14-00548" class="html-bibr">37</a>].</p>
Full article ">Figure 11
<p>Boxplots of Ti/V and (Ti+V)/(Al+Mn) ratios.</p>
Full article ">Figure 12
<p>The concentration of Ti vs. V in magnetite. (<b>a</b>) EMPA; (<b>b</b>) LA-ICP-MS. The red area includes igneous-formed magnetite, while the blue area is defined by hydrothermal magnetite based on the data set of [<a href="#B40-minerals-14-00548" class="html-bibr">40</a>]. The data of MagI are plotted mostly in the overlapping area (green color), with some samples tending toward pure igneous magnetite. The concentration of V vs. Cr in magnetite. (<b>c</b>) EMPA; (<b>d</b>) LA-ICP-MS [<a href="#B45-minerals-14-00548" class="html-bibr">45</a>].</p>
Full article ">
Back to TopTop