[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

Article Types

Countries / Regions

Search Results (200)

Search Parameters:
Keywords = dynamin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 6780 KiB  
Article
AIBP Protects Müller Glial Cells Against Oxidative Stress-Induced Mitochondrial Dysfunction and Reduces Retinal Neuroinflammation
by Seunghwan Choi, Soo-Ho Choi, Tonking Bastola, Keun-Young Kim, Sungsik Park, Robert N. Weinreb, Yury I. Miller and Won-Kyu Ju
Antioxidants 2024, 13(10), 1252; https://doi.org/10.3390/antiox13101252 - 17 Oct 2024
Abstract
Glaucoma, an optic neuropathy with the loss of retinal ganglion cells (RGCs), is a leading cause of irreversible vision loss. Oxidative stress and mitochondrial dysfunction have a significant role in triggering glia-driven neuroinflammation and subsequent glaucomatous RGC degeneration in the context of glaucoma. [...] Read more.
Glaucoma, an optic neuropathy with the loss of retinal ganglion cells (RGCs), is a leading cause of irreversible vision loss. Oxidative stress and mitochondrial dysfunction have a significant role in triggering glia-driven neuroinflammation and subsequent glaucomatous RGC degeneration in the context of glaucoma. It has previously been shown that apolipoprotein A-I binding protein (APOA1BP or AIBP) has an anti-inflammatory function. Moreover, Apoa1bp−/− mice are characterized by retinal neuroinflammation and RGC loss. In this study, we found that AIBP deficiency exacerbated the oxidative stress-induced disruption of mitochondrial dynamics and function in the retina, leading to a further decline in visual function. Mechanistically, AIBP deficiency-induced oxidative stress triggered a reduction in glycogen synthase kinase 3β and dynamin-related protein 1 phosphorylation, optic atrophy type 1 and mitofusin 1 and 2 expression, and oxidative phosphorylation, as well as the activation of mitogen-activated protein kinase (MAPK) in Müller glia dysfunction, leading to cell death and inflammatory responses. In vivo, the administration of recombinant AIBP (rAIBP) effectively protected the structural and functional integrity of retinal mitochondria under oxidative stress conditions and prevented vision loss. In vitro, incubation with rAIBP safeguarded the structural integrity and bioenergetic performance of mitochondria and concurrently suppressed MAPK activation, apoptotic cell death, and inflammatory response in Müller glia. These findings support the possibility that AIBP promotes RGC survival and restores visual function in glaucomatous mice by ameliorating glia-driven mitochondrial dysfunction and neuroinflammation. Full article
Show Figures

Figure 1

Figure 1
<p>AIBP deficiency exacerbates visual dysfunction induced by oxidative stress. (<b>a</b>) Representative graphs of total recordings of pERG analysis among groups. (<b>b</b>) Quantification analysis of pERG test among groups. <span class="html-italic">N</span> = 8 mice. (<b>c</b>) Quantification analysis of optomotor response among groups. <span class="html-italic">N</span> = 8 mice. (<b>d</b>) Quantification analysis of pVEP tests among groups. <span class="html-italic">N</span> = 8 mice. (<b>e</b>) TLR4 and IL-1β immunohistochemistry in retina. Representative images show TLR4- and IL-1β-positive Müller glial cells in retina. Note that quantification analysis showed significant increase in IL-1β immunoreactive intensity under oxidative stress with AIBP deficiency compared with oxidative stress alone. <span class="html-italic">N</span> = 10 sections from middle area of retina from 3 mice. Images were taken with 20X magnification. Scale bar: 20 μm. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. pERG, pattern electroretinogram; PQ, paraquat; pVEP, pattern visual evoked potential; WT, wild-type.</p>
Full article ">Figure 2
<p>AIBP deficiency intensifies impairment of retinal mitochondrial dynamics, OXPHOS activity, and mitochondrial biogenesis induced by oxidative stress. (<b>a</b>) Total DRP1, phospho-DRP S616, and phospho-DRP1 S637 expression in retina. <span class="html-italic">N</span> = 3 mice. (<b>b</b>) OPA1, MFN1, and MFN2 expression in retina. <span class="html-italic">N</span> = 3 to 6 retinas from 3 mice. (<b>c</b>) AIBP, PGC-1α, and TFAM expression in retina. <span class="html-italic">N</span> = 3 mice. (<b>d</b>) OXPHOS complex expression in retina. <span class="html-italic">N</span> = 3 retinas from mice. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. PQ, paraquat; WT, wild-type.</p>
Full article ">Figure 3
<p>Oxidative stress worsens structural and functional impairment of mitochondria in Müller glia cells lacking AIBP. (<b>a</b>) Oligomycin A, FCCP and rotenone were sequentially added at indicated time point. Basal respiration indicates starting basal OCR and value which was set to 100%. Maximum respiration represents ratio between FCCP uncoupled OCR and basal OCR. (<b>b</b>) Quantitative analyses of basal, maximal, and ATP-linked respiration and spare respiratory capacity in rMC-1 cells. <span class="html-italic">N</span> = 8 replicated wells. (<b>c</b>) Quantitative analysis of MMP and mitochondrial ROS. <span class="html-italic">N</span> = 3 independent experiments in rMC-1 cells. (<b>d</b>) AIBP, total DRP1, phospho-DRP S616, and phospho-DRP1 S637 expression in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>e</b>) Oligomycin A, FCCP and rotenone were sequentially added at indicated time point. Basal respiration indicates starting basal OCR and value which was set to 100%. Maximum respiration represents ratio between FCCP uncoupled OCR and basal OCR. (<b>f</b>) Quantitative analyses of basal, maximal, and ATP-linked respiration and spare respiratory capacity in rMC-1 cells. <span class="html-italic">N</span> = 8 replicated wells. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. PQ, paraquat; FCCP, carbonyl cyanide p-trifluoromethoxyphenylhydrazone; OCR, oxygen consumption rate.</p>
Full article ">Figure 4
<p>Oxidative stress exacerbates MAPK activation and apoptotic cell death and inflammatory response in Müller glia cells lacking AIBP. (<b>a</b>) p38, phospho-p38 (pp38), ERK1/2, phospho-ERK1/2 (pERK1/2) expression in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>b</b>) caspase-1, cleaved caspase-1, caspase-3, and cleaved caspase-3 expression in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>c</b>,<b>d</b>) Representative images show cleaved caspase-3-positive rMC-1 cells in the retina. Note that quantification analysis showed a significant increase in cleaved caspase-3 immunoreactive intensity in rMC-1 cells under oxidative stress with AIBP knockdown compared with control rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>e</b>) Quantitative real-time PCR analysis of <span class="html-italic">Nlrp3</span>, <span class="html-italic">Il-1β</span>, <span class="html-italic">Il-6</span>, and <span class="html-italic">Tnfα</span> mRNA expression in rMC-1 cells. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. PQ, paraquat.</p>
Full article ">Figure 5
<p>Administration of rAIBP prevents visual dysfunction, restores mitochondrial dynamics, and enhances OXPHOS activity in retina. (<b>a</b>) Representative graphs of total recordings of pERG analysis among groups. (<b>b</b>) Quantification analysis of pERG test among groups. <span class="html-italic">N</span> = 6 mice. (<b>c</b>) Quantification analysis of optomotor response among groups. <span class="html-italic">N</span> = 6 mice. (<b>d</b>) Quantification analysis of pVEP tests among groups. <span class="html-italic">N</span> = 6 mice. (<b>e</b>) OPA1, total DRP1, phospho-DRP S616, and phospho-DRP1 S637 expression in retina. <span class="html-italic">N</span> = 3 mice. (<b>f</b>) OXPHOS complex expression in retina. <span class="html-italic">N</span> = 3 mice. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. BSA, bovine serum albumin; PQ, paraquat; pERG, pattern electroretinogram; pVEP, pattern visual evoked potential.</p>
Full article ">Figure 6
<p>Administration of rAIBP reduced TLR4-associated lipid rafts in Müller glia exposed to oxidative stress. (<b>a</b>) Representative images of TLR4 (green)-LR (red) immunoreactivity (red). Scale bar: 10 μm. (<b>b</b>) Quantitative fluorescent intensity of TLR4-LR immunoreactivity in rMC-1 cells. a. <span class="html-italic">N</span> = 3 independent experiments. Scale bar: 10 μm. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. LR, lipid raft; PQ, paraquat; BSA, bovine serum albumin.</p>
Full article ">Figure 7
<p>Administration of rAIBP preserves mitochondrial function and dynamics in Müller glia exposed to oxidative stress. (<b>a</b>) Quantitative analysis of MMP and mitochondrial ROS in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>b</b>) Total DRP1, phospho-DRP S616, and phospho-DRP1 S637 expression in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>c</b>) Oligomycin A, FCCP and rotenone were sequentially added at indicated time point. Basal respiration indicates starting basal OCR and value which was set to 100%. Maximum respiration represents ratio between FCCP uncoupled OCR and basal OCR. (<b>d</b>) Quantitative analyses of basal, maximal, and ATP-linked respiration and spare respiratory capacity in rMC-1 cells. <span class="html-italic">N</span> = 4 replicated wells. (<b>e</b>) Glucose, oligomycin A and 2DG were sequentially added at indicated time point. (<b>f</b>) Quantitative analyses of glycolysis, glycolytic capacity, and glycolytic reserve in rMC-1 cells. <span class="html-italic">N</span> = 5 replicated wells. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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. BSA, bovine serum albumin; PQ, paraquat; FCCP, carbonyl cyanide p-trifluoromethoxyphenylhydrazone; OCR, oxygen consumption rate; ECAR, extracellular acidification rate.</p>
Full article ">Figure 8
<p>Administration of rAIBP inhibits MAPK activation, apoptotic cell death, and inflammatory response in Müller glia exposed to oxidative stress. (<b>a</b>) p38, phospho-p38 9 (pp38), ERK1/2, phospho-ERK1/2 (pERK1/2) expression in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>b</b>) caspase-1, cleaved caspase-1, caspase-3, and cleaved caspase-3 expression in rMC-1 cells. <span class="html-italic">N</span> = 3 independent experiments. (<b>c</b>) Representative images show cleaved caspase-3-positive rMC-1 cells. (<b>d</b>) Note that quantification analysis showed significant decrease in cleaved caspase-3 immunoreactive intensity in rMC-1 cells with rAIBP treatment compared with BSA-treated cells under oxidative stress. <span class="html-italic">N</span> = 3 independent experiments. (<b>e</b>) Quantitative real-time PCR analysis of <span class="html-italic">Nlrp3</span>, <span class="html-italic">Il-1β</span>, <span class="html-italic">Il-6</span>, and <span class="html-italic">Tnfα</span> mRNA expression in rMC-1 cells. Error bars represent SEM. Statistical significance was determined using one-way ANOVA test. * <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; 0.0001. PQ, paraquat; BSA, bovine serum albumin.</p>
Full article ">
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 - 15 Sep 2024
Viewed by 560
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 ">
22 pages, 13849 KiB  
Article
Kinetic Landscape of Single Virus-like Particles Highlights the Efficacy of SARS-CoV-2 Internalization
by Aleksandar Atemin, Aneliya Ivanova, Wiley Peppel, Rumen Stamatov, Rodrigo Gallegos, Haley Durden, Sonya Uzunova, Michael D. Vershinin, Saveez Saffarian and Stoyno S. Stoynov
Viruses 2024, 16(8), 1341; https://doi.org/10.3390/v16081341 - 22 Aug 2024
Viewed by 3685
Abstract
The efficiency of virus internalization into target cells is a major determinant of infectivity. SARS-CoV-2 internalization occurs via S-protein-mediated cell binding followed either by direct fusion with the plasma membrane or endocytosis and subsequent fusion with the endosomal membrane. Despite the crucial role [...] Read more.
The efficiency of virus internalization into target cells is a major determinant of infectivity. SARS-CoV-2 internalization occurs via S-protein-mediated cell binding followed either by direct fusion with the plasma membrane or endocytosis and subsequent fusion with the endosomal membrane. Despite the crucial role of virus internalization, the precise kinetics of the processes involved remains elusive. We developed a pipeline, which combines live-cell microscopy and advanced image analysis, for measuring the rates of multiple internalization-associated molecular events of single SARS-CoV-2-virus-like particles (VLPs), including endosome ingression and pH change. Our live-cell imaging experiments demonstrate that only a few minutes after binding to the plasma membrane, VLPs ingress into RAP5-negative endosomes via dynamin-dependent scission. Less than two minutes later, VLP speed increases in parallel with a pH drop below 5, yet these two events are not interrelated. By co-imaging fluorescently labeled nucleocapsid proteins, we show that nucleocapsid release occurs with similar kinetics to VLP acidification. Neither Omicron mutations nor abrogation of the S protein polybasic cleavage site affected the rate of VLP internalization, indicating that they do not confer any significant advantages or disadvantages during this process. Finally, we observe that VLP internalization occurs two to three times faster in VeroE6 than in A549 cells, which may contribute to the greater susceptibility of the former cell line to SARS-CoV-2 infection. Taken together, our precise measurements of the kinetics of VLP internalization-associated processes shed light on their contribution to the effectiveness of SARS-CoV-2 propagation in cells. Full article
(This article belongs to the Special Issue Emerging Concepts in SARS-CoV-2 Biology and Pathology 2.0)
Show Figures

Figure 1

Figure 1
<p>Visualization of SARS-CoV-2 VLPs’ binding and internalization into host cells. (<b>A</b>) Binding and internalization of SARS-CoV-2 VLP<sup>Wu</sup>:M<sup>Ch</sup> into U2OS cells overexpressing ACE2-Neon green and TMPRSS2. The montage shows VLPs binding to filopodia first and then migrating into the cell body. (<b>B</b>) U2OS cells expressing mNeonGreen (to visualize cell volume) treated with VLP<sup>Wu</sup>:M<sup>Ch</sup>. (<b>C</b>) The positions of VLP<sup>Wu</sup>:M<sup>Ch</sup> in the transparent 3D volume of a U2OS cell (viewed from the top). (<b>D</b>) The cell from (<b>C</b>) shown from the side. (<b>E</b>) Same as (<b>C</b>) except VLP<sup>Wu</sup>:M<sup>Ch</sup> are shown in a volumized U2OS cell to highlight internalized VLPs (viewed from the top). (<b>F</b>) Graph showing the speed of the VLP (2) from (<b>A</b>); each point in the graph represents the speed of the particle over time extracted from two consecutive images in 1D—vertically (Z), 2D (XY), and 3D (XYZ). The kymograph represents the change in intensity of the particle and its position in Z as described in detail in <a href="#app1-viruses-16-01341" class="html-app">Figure S1</a>. The time scales of the graph and kymographs are aligned. (<b>G</b>) Same as (<b>F</b>), but for the particle in (<b>B</b>). (<b>H</b>) Cells treated with Tubulin 610 conjugated to cabazitaxel, showing the transport of VLP<sup>Wu</sup>:M<sup>Ch</sup> via microtubules. (<b>I</b>) VLP<sup>Wu</sup>:M<sup>Ch</sup> treated with a Recombinant Anti-SARS-CoV-2 S1 antibody. The VLPs aggregate and are unable to internalize into cells.</p>
Full article ">Figure 2
<p>Dynamics of dynamin recruitment to the site of SARS-CoV-2 VLP<sup>Wu</sup>:M<sup>Ch</sup> binding in Vero E6 cells. (<b>A</b>) A representative image of SARS-CoV-2 VLP<sup>Wu</sup>:M<sup>Ch</sup> added to Vero E6 cells and dynamin recruitment. The montage shows consecutive images from the area shown by the white square; recruitment of dynamin is observed at 5:45 min (white arrow). (<b>B</b>) The top graph represents the speed of the VLP shown in (<b>A</b>); the bottom graph demonstrates the dynamics of the intensity of both channels; the kymograph represents the change in intensity and position of the particle. (<b>C</b>) Distribution of the number of dynamin recruitment events to bound VLP<sup>Wu</sup>:M<sup>Ch</sup>, n = 41. (<b>D</b>) Distribution of time intervals between VLP<sup>Wu</sup>:M<sup>Ch</sup> binding and the first recruitment of dynamin, n = 37. (<b>E</b>) Distribution of the time intervals between VLP<sup>Wu</sup>:M<sup>Ch</sup> binding and VLP speed increase, n = 18. (<b>F</b>) Cells treated with Dynole 34-2, which inhibits vesicle-mediated endocytosis, showing the inability of VLPs to enter cells. (<b>G</b>) Cells expressing GFP-tagged Rab-5 showing lack of co-localization of the Rab5-positive vesicles with the VLP<sup>Wu</sup>:M<sup>Ch</sup>. (<b>H</b>) Cells treated with LysoTracker showing co-localization of the VLP<sup>Wu</sup>:M<sup>Ch</sup> with acidic vesicles (lysosomes or late endosomes).</p>
Full article ">Figure 3
<p>Dynamics of SARS-CoV-2 VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R binding, pH decrease, and speed increase in A549 and Vero E6 cells. (<b>A</b>) Percentage of VLPs in which only M-pHluorin intensity decreases (blue), the intensities of M-pHluorin and M-mCherry decrease simultaneously (orange), or neither decreases (gray) for 100 min after addition of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R in VeroE6 cells with or without ACE2 and TMPRSS2 overexpression. (<b>B</b>) Same experiment as (<b>A</b>) in A549 cells. (<b>C</b>) Comparison of the M-pHluorin intensity decrease during internalization of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R in A549 and Vero E6 cells. The average intensity of pHluorin is represented as a function of time where the individual VLPs were aligned to the start of VLP pHluorin decrease (0 min). The average M-mCherry intensity of the same particles is also presented. Errors bars represent the standard deviation. For A549 n = 55, for Vero E6 n = 93. (<b>D</b>) Comparison of the dynamics of pH decrease during internalization of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R in A549 and Vero E6 cells. The pH decrease is calculated based on the measured M-pHluorin intensity decrease. The average pH of VLPs is represented as a function of time where individual VLPs are aligned to the start of VLP pHluorin decrease (0 min). Error bars represent the standard deviation. For A549 n = 55, for VeroE6 n = 93. (<b>E</b>) The average speed of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R in A549 and Vero E6 cells measured based on the tracked M-mCherry signal. The average speed of VLPs was calculated after alignment of the individual VLP speeds to the start of the VLP pHluorin signal decrease (0 min). Error bars represent the standard deviation. For A549 n = 55, for Vero E6 n = 93. (<b>F</b>) Distribution of time intervals between VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R binding and the start of pHluorin intensity decrease in A549 cells or Vero E6 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.01. For A549 n = 55, for Vero E6 n = 93. (<b>G</b>) Distribution of time intervals between VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R binding and the start of VLP speed increase in A549 and Vero E6 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.01. For A549 n = 55, for Vero E6 n = 93. (<b>H</b>) Distribution of time intervals between the start of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R pHluorin intensity decrease and the start of VLP speed increase in A549 and Vero E6 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. For A549 n = 55, for Vero E6 n = 93. (<b>I</b>) Distribution of the estimated pH of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, calculated based on the pHluorin signal at the moment when VLP speed started to increase in A549 and Vero E6 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. For A549 n = 55, for Vero E6 n = 93. (<b>J</b>) Representative time-lapse images (top), corresponding VLP speed and intensity graphs (middle), and kymographs (merged, M-mCherry, and M-pHluorin) in all dimensions (bottom) for a single VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R undergoing internalization in an VeroE6 cell. In this example, the speed increases in parallel with pHluorin signal decrease. (<b>K</b>) Same as (<b>J</b>) in a A549 cell.</p>
Full article ">Figure 4
<p>Comparison of VLP binding, acidification, and speed increase dynamics between VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, and VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R during internalization in A549 cells. (<b>A</b>) Percentage of VLPs in which only the M-pHluorin intensity decreases (blue), M-pHluorin and M-mCherry intensities decrease simultaneously (orange), or neither decreases (gray). (<b>B</b>) Comparison of pHluorin intensity decrease during the internalization of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R and VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R in A549 cells. The average intensity of pHluorin is represented as a function of time where the individual VLPs were aligned to the start of VLP pHluorin decrease (0 min). The average M-mCherry intensity is also presented. Error bars represent the standard deviation. For VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55, for VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 48, for VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 62. (<b>C</b>) Representative VLP speed and intensity graphs (top) and kymographs (merged, M-mCherry, and M-pHluorin) in all dimensions (bottom) for a single VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R undergoing internalization in an A549 cell. In the example, the speed increases in parallel with pHluorin signal decrease. (<b>D</b>) Same as (<b>C</b>) but for VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R. (<b>E</b>) Distribution of time intervals between VLP binding and start of pHluorin intensity decrease for individual VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, and VLP<sup>del-1</sup>: M<sup>Ch</sup>M<sup>pH</sup>R during internalization in A549 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55, for VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 48, for VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 62. (<b>F</b>) Distribution of time intervals between VLP binding and start of speed increase for individual VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, and VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R during internalization in A549 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. For A549-WT n = 55, for A549-Omi n = 48, for A549-del1 n = 62. (<b>G</b>) Distribution of time intervals between VLP intensity decrease and start of speed increase for individual VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, and VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R during internalization in A549 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. For VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55, for VLP<sup>Omi</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 48, for VLP<sup>del-1</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 62.</p>
Full article ">Figure 5
<p>Comparison of VLP binding, acidification, and speed increase dynamics between VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R, VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R, and VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R during internalization in A549 cells. (<b>A</b>) Percentage of VLPs in which only the signal intensity of N-EGFP (middle and right) or M-pHluorin (left) decreases (blue), the intensities of both N-EGFP/M-pHluorin and M-mCherry decrease simultaneously (orange), and neither decreases (gray). VLPs were tracked for 100 min after addition. (<b>B</b>) Comparison of N-EGFP intensity decrease rate during VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup> R and VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R internalization in A549 cells. The average intensity of N-EGFP is presented as a function of time where the individual VLPs were aligned to the start of N-EGFP signal decrease (0 min). The average M-mCherry intensity is also presented. Error bars represent the standard deviation. For A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 17, for A549 VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup> n = 34. (<b>C</b>) Comparison of M-pHluorin and N-EGFP intensity decrease rates during internalization of VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R and VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R, respectively, in A549 cells. The average intensities of N-EGFP and M-pHluorin are presented as a function of time where the individual VLPs were aligned to the start of VLP N-EGFP or M-pHluorin decrease (0 min). The average M-mCherry intensity of the same particles is also presented. Error bars represent the standard deviation. For A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55, for A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 17. (<b>D</b>) Distribution of time intervals between VLP binding and the start of pHluorin/N-EGFP intensity decrease for individual VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R, and VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R during internalization in A549 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. For A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 17, for VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 34, and for VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55. (<b>E</b>) Distribution of time intervals between VLP binding and start of speed increase for individual VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R, and VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R during internalization in A549 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01. For A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 17, for A549 VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 34, and for A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55. (<b>F</b>) Distribution of time intervals between start of pHluorin/N-EGFP intensity decrease and start of speed increase for individual VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R, VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R, and VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R during internalization in A549 cells. Two-tailed Student’s <span class="html-italic">t</span>-test; NS <span class="html-italic">p</span> &gt; 0.01; * <span class="html-italic">p</span> &lt; 0.01. For A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 17, for A549 VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R n = 34, and for A549 VLP<sup>Wu</sup>:M<sup>Ch</sup>M<sup>pH</sup>R n = 55. (<b>G</b>) Schematic of the major internalization-associated events through time for wild-type and mutant VLPs in VeroE6 and A549 cells. (<b>H</b>) Representative VLP speed and intensity graphs (top) and kymographs (merged, M-mCherry, and N-EGFP) in all dimensions (bottom) for a single VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R undergoing internalization in an A549 cell. In this example, the speed increases in parallel with pHluorin signal decrease. (<b>I</b>) Same as (<b>H</b>) for VLP<sup>Omi</sup>:M<sup>Ch</sup>N<sup>E</sup>R.</p>
Full article ">Figure 6
<p>Tracking of nucleocapsid release following VLP internalization. (<b>A</b>) Time-lapse images of N-EGFP release during internalization of VLP<sup>Wu</sup>:M<sup>Ch</sup>N<sup>E</sup>R in A549 cells where the separation of the nucleocapsid (N-EGFP) from the VLP membrane (M-mCherry) can be observed. (<b>B</b>) Changes in N-EGFP and M-mCherry during nucleocapsid release for the above VLP (tracked based on the N-EGFP signal). (<b>C</b>) VLP speed profile during nucleocapsid release measured based on M-mCherry movement for the same VLP. (<b>D</b>) VLP speed profile during nucleocapsid release measured based on N-EGFP movement for the same VLP. Note the speed increase during nucleocapsid release, which is missing in (<b>C</b>). (<b>E</b>) Kymogram in all dimensions measured based on M-mCherry tracking. (<b>F</b>) Kymogram in all dimensions based on N-EGFP tracking. Note the change in movement along the Z-axis during nucleocapsid release, which is missing in (<b>E</b>).</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 1465
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 1030
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 ">
16 pages, 7113 KiB  
Article
Differential Effects of Three Medium-Chain Fatty Acids on Mitochondrial Quality Control and Skeletal Muscle Maturation
by Ryoichi Nishida, Shota Nukaga, Isao Kawahara, Yoshihiro Miyagawa, Kei Goto, Chie Nakashima, Yi Luo, Takamitsu Sasaki, Kiyomu Fujii, Hitoshi Ohmori, Ruiko Ogata, Shiori Mori, Rina Fujiwara-Tani and Hiroki Kuniyasu
Antioxidants 2024, 13(7), 821; https://doi.org/10.3390/antiox13070821 - 9 Jul 2024
Viewed by 1070
Abstract
Nutritional interventions are one focus of sarcopenia treatment. As medium-chain fatty acids (MCFAs) are oxidized in the mitochondria and produce energy through oxidative phosphorylation (OXPHOS), they are key parts of nutritional interventions. We investigated the in vitro effects of three types of MCFA, [...] Read more.
Nutritional interventions are one focus of sarcopenia treatment. As medium-chain fatty acids (MCFAs) are oxidized in the mitochondria and produce energy through oxidative phosphorylation (OXPHOS), they are key parts of nutritional interventions. We investigated the in vitro effects of three types of MCFA, caprylic acid (C8), capric acid (C10), and lauric acid (C12), in skeletal muscle cells. Compared with C10 and C12, C8 promoted mitophagy through the phosphatase and tensin homolog (PTEN)-induced kinase 1-Parkin pathway and increased the expression of peroxisome proliferator-activated receptor gamma coactivator 1-α and dynamin-related protein 1 to reduce mitochondrial oxidative stress and promote OXPHOS. Furthermore, the expression of myogenic differentiation 1 and myosin heavy chain increased in myotubes, thus promoting muscle differentiation and maturation. These results suggest that C8 improves mitochondrial quality and promotes skeletal muscle maturation; in contrast, C10 and C12 poorly promoted mitochondrial quality control and oxidative stress and suppressed energy production. Future animal experiments are required to establish the usefulness of C8 for nutritional interventions for sarcopenia. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of medium-chain fatty acid (MCFA) treatment on the skeletal muscle differentiation of C2C12 myotube cells. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with MCFAs (48 h). Then, the cells were examined for skeletal muscle differentiation. (<b>A</b>) Effects of MCFAs on cell proliferation. (<b>B</b>) Effects of MCFA treatments on MyoD expression levels assessed using reverse transcription–polymerase chain reaction (RT-PCR). (<b>C</b>) Semi-quantification of MyoD expression levels. (<b>D</b>) Fluorescence immunostaining of MyoD. (<b>E</b>) MyoD-positive cell rate. (<b>F</b>) MYH protein expression levels. (<b>G</b>) Semi-quantification of MYH expression levels. (<b>H</b>) MYH fluorescence immunostaining. (<b>I</b>) Semi-quantification of MYH-positive cell areas. (<b>J</b>) Fused cell ratio to all cells. Scale bar, 50 μm; error bar, standard deviation from three independent trials. Statistical differences were calculated using analysis of variance with Bonferroni correction. CTRL, control; C8, caprylic acid; C10, capric acid; C12, lauric acid; L, low; H, high; MyoD, myogenic differentiation 1; MYH, myosin heavy chain.</p>
Full article ">Figure 1 Cont.
<p>Effects of medium-chain fatty acid (MCFA) treatment on the skeletal muscle differentiation of C2C12 myotube cells. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with MCFAs (48 h). Then, the cells were examined for skeletal muscle differentiation. (<b>A</b>) Effects of MCFAs on cell proliferation. (<b>B</b>) Effects of MCFA treatments on MyoD expression levels assessed using reverse transcription–polymerase chain reaction (RT-PCR). (<b>C</b>) Semi-quantification of MyoD expression levels. (<b>D</b>) Fluorescence immunostaining of MyoD. (<b>E</b>) MyoD-positive cell rate. (<b>F</b>) MYH protein expression levels. (<b>G</b>) Semi-quantification of MYH expression levels. (<b>H</b>) MYH fluorescence immunostaining. (<b>I</b>) Semi-quantification of MYH-positive cell areas. (<b>J</b>) Fused cell ratio to all cells. Scale bar, 50 μm; error bar, standard deviation from three independent trials. Statistical differences were calculated using analysis of variance with Bonferroni correction. CTRL, control; C8, caprylic acid; C10, capric acid; C12, lauric acid; L, low; H, high; MyoD, myogenic differentiation 1; MYH, myosin heavy chain.</p>
Full article ">Figure 2
<p>Effects of different medium-chain fatty acids (MCFAs) on mitochondrial quality control in C2C12 myotubes. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with MCFAs (48 h). Then, mitochondrial quality control was examined. (<b>A</b>) Mitochondrial volume. (<b>B</b>) Quantification of the mitochondrial volume. (<b>C</b>) PGC1α mRNA expression. The lower panel is the semi-quantification. (<b>D</b>) Mitochondrial protein expression. (<b>E</b>) Semi-quantification of panel (<b>C</b>). (<b>F</b>) Fluorescence staining for mitophagy. (<b>G</b>) Semi-quantification of panel (<b>F</b>). (<b>H</b>) Mitochondrial superoxide levels. (<b>I</b>) Semi-quantitative analysis of Panel G. Scale bar, 50 μm; error bar, standard deviation from three independent trials. Statistical differences were calculated using analysis of variance (ANOVA) with Bonferroni correction. CTRL, control; C8, caprylic acid; C10, capric acid; C12, lauric acid; PGC1α, peroxisome proliferator-activated receptor gamma coactivator 1-α; LETM1, leucine zipper and EF-hand-containing transmembrane protein 1; TOM20, translocase of the outer mitochondrial membrane 20; DRP1, dynamin-related protein 1; MFN2, mitofusin 2; PINK1, PTEN-induced kinase 1; LC3, microtubule-associated protein 1A/1B-light chain 3.</p>
Full article ">Figure 2 Cont.
<p>Effects of different medium-chain fatty acids (MCFAs) on mitochondrial quality control in C2C12 myotubes. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with MCFAs (48 h). Then, mitochondrial quality control was examined. (<b>A</b>) Mitochondrial volume. (<b>B</b>) Quantification of the mitochondrial volume. (<b>C</b>) PGC1α mRNA expression. The lower panel is the semi-quantification. (<b>D</b>) Mitochondrial protein expression. (<b>E</b>) Semi-quantification of panel (<b>C</b>). (<b>F</b>) Fluorescence staining for mitophagy. (<b>G</b>) Semi-quantification of panel (<b>F</b>). (<b>H</b>) Mitochondrial superoxide levels. (<b>I</b>) Semi-quantitative analysis of Panel G. Scale bar, 50 μm; error bar, standard deviation from three independent trials. Statistical differences were calculated using analysis of variance (ANOVA) with Bonferroni correction. CTRL, control; C8, caprylic acid; C10, capric acid; C12, lauric acid; PGC1α, peroxisome proliferator-activated receptor gamma coactivator 1-α; LETM1, leucine zipper and EF-hand-containing transmembrane protein 1; TOM20, translocase of the outer mitochondrial membrane 20; DRP1, dynamin-related protein 1; MFN2, mitofusin 2; PINK1, PTEN-induced kinase 1; LC3, microtubule-associated protein 1A/1B-light chain 3.</p>
Full article ">Figure 3
<p>Effects of medium-chain fatty acids (MCFAs) on mitochondrial respiratory function in C2C12 myotubes. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with MCFAs (48 h). Then, the cells were subjected to respiration flux analyses. (<b>A</b>,<b>E</b>,<b>I</b>) Flux analysis. (<b>B</b>,<b>F</b>,<b>J</b>) Energy metabolic parameters calculated from the flux analysis. (<b>C</b>,<b>G</b>,<b>K</b>) ECAR analysis. (<b>D</b>,<b>H</b>,<b>L</b>) Energy metabolism phenotypes. Error bars: standard deviation from three independent trials. CTRL, control; C8, caprylic acid; C10, capric acid; C12, lauric acid; OCR, oxygen consumption rates; ECAR, extracellular acidification rate.</p>
Full article ">Figure 4
<p>Effects of autophagy enhancement by C8 on skeletal muscle differentiation. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with C8 (48 h). We then examined the effects of C8 on skeletal muscle differentiation when autophagy was inhibited by chloroquine (CQ). (<b>A</b>) MyoD expression. (<b>B</b>) Semi-quantification of panel (<b>A</b>). (<b>C</b>) Fluorescence immunostaining for MyoD. (<b>D</b>) MyoD-positive rate. (<b>E</b>) MYH levels. (<b>F</b>) Semi-quantification of MYH protein content normalized to cell number. (<b>G</b>) MYH fluorescent fluorescence. (<b>H</b>) MYH-positive cell area. (<b>I</b>) Fused cell ratio. Scale bar, 50 μm; error bar, standard deviation from three independent trials. Statistical differences were calculated using analysis of variance (ANOVA) with Bonferroni correction. C8, caprylic acid; CQ, chloroquine; MyoD, myogenic differentiation 1; MYH, myosin heavy chain; CTRL, control; C8, caprylic acid; Hoechst, Hoechst 33342 dye.</p>
Full article ">Figure 5
<p>Effects of C8-induced autophagy enhancement on mitochondrial turnover. C2C12 cells were induced to myotube differentiation for 48 h with pretreatment with C8 (48 h). We then investigated the effects of C8 on myotube mitochondria when autophagy was inhibited using chloroquine (CQ). (<b>A</b>) Mitophagy fluorescence staining. (<b>B</b>) Semi-quantification of panel (<b>A</b>). (<b>C</b>) Changes in mitochondria-related protein levels. (<b>D</b>) Semi-quantification of panel (<b>C</b>). (<b>E</b>) Flux analysis. (<b>F</b>) Energy metabolic parameters calculated from the flux analysis. (<b>G</b>) Mitochondrial superoxide levels. (<b>H</b>) Semi-quantitative analysis of panel G. Scale bar, 50 μm; error bar, standard deviation from three independent trials. Statistical differences were calculated using analysis of variance (ANOVA) with Bonferroni correction. C8, caprylic acid; CQ, chloroquine; MYH, myosin heavy chain; CTRL, control; C8, caprylic acid; PGC1α, peroxisome proliferator-activated receptor gamma coactivator 1-α; TOM20, translocase of the outer mitochondrial membrane 20; DRP1, dynamin-related protein 1; MFN2, mitofusin 2; PINK1, PTEN-induced kinase 1; LC3, microtubule-associated protein 1A/1B-light chain 3.</p>
Full article ">
10 pages, 602 KiB  
Communication
Discovery of Pathogenic Variants Associated with Idiopathic Recurrent Pregnancy Loss Using Whole-Exome Sequencing
by Jeong Yong Lee, JaeWoo Moon, Hae-Jin Hu, Chang Soo Ryu, Eun Ju Ko, Eun Hee Ahn, Young Ran Kim, Ji Hyang Kim and Nam Keun Kim
Int. J. Mol. Sci. 2024, 25(10), 5447; https://doi.org/10.3390/ijms25105447 - 17 May 2024
Viewed by 1206
Abstract
Idiopathic recurrent pregnancy loss (RPL) is defined as at least two pregnancy losses before 20 weeks of gestation. Approximately 5% of pregnant couples experience idiopathic RPL, which is a heterogeneous disease with various causes including hormonal, chromosomal, and intrauterine abnormalities. Although how pregnancy [...] Read more.
Idiopathic recurrent pregnancy loss (RPL) is defined as at least two pregnancy losses before 20 weeks of gestation. Approximately 5% of pregnant couples experience idiopathic RPL, which is a heterogeneous disease with various causes including hormonal, chromosomal, and intrauterine abnormalities. Although how pregnancy loss occurs is still unknown, numerous biological factors are associated with the incidence of pregnancy loss, including genetic variants. Whole-exome sequencing (WES) was conducted on blood samples from 56 Korean patients with RPL and 40 healthy controls. The WES data were aligned by means of bioinformatic analysis, and the detected variants were annotated using machine learning tools to predict the pathogenicity of protein alterations. Each indicated variant was confirmed using Sanger sequencing. A replication study was also conducted in 112 patients and 114 controls. The Variant Effect Scoring Tool, Combined Annotation Dependent Depletion tool, Sorting Intolerant from Tolerant annotation tool, and various databases detected 10 potential variants previously associated with spontaneous abortion genes in patients by means of a bioinformatic analysis of WES data. Several variants were detected in more than one patient. Interestingly, several of the detected genes were functionally clustered, including some with a secretory function (mucin 4; MUC4; rs200737893 G>A and hyaluronan-binding protein 2; HABP2; rs542838125 G>T), in which growth arrest-specific 2 Like 2 (GAS2L2; rs140842796 C>T) and dynamin 2 (DNM2; rs763894364 G>A) are functionally associated with cell protrusion and the cytoskeleton. ATP Binding Cassette Subfamily C Member 6 (ABCC6) was the only gene with two variants. HABP2 (rs542838125 G>T), MUC4 (rs200737893 G>A), and GAS2L2 (rs140842796 C>T) were detected in only the patient group in the replication study. The combination of WES and machine learning tools is a useful method to detect potential variants associated with RPL. Using bioinformatic tools, we found 10 potential variants in 9 genes. WES data from patients are needed to better understand the causes of RPL. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Whole-exome sequencing analysis pipeline. An abortion-associated, matching gene list was filtered using the Disgenet and ClinVar databases and the results were compared with those in the control sample to select patient-specific variants. Combined Annotation Dependent Depletion: CADD; Variant Effect Scoring Tool: VEST; Sorting Intolerant from Tolerant: SIFT.</p>
Full article ">
17 pages, 6248 KiB  
Article
Inhibition of Drp1–Filamin Protein Complex Prevents Hepatic Lipid Droplet Accumulation by Increasing Mitochondria–Lipid Droplet Contact
by Kohei Ariyoshi, Kazuhiro Nishiyama, Yuri Kato, Xinya Mi, Tomoya Ito, Yasu-Taka Azuma, Akiyuki Nishimura and Motohiro Nishida
Int. J. Mol. Sci. 2024, 25(10), 5446; https://doi.org/10.3390/ijms25105446 - 17 May 2024
Cited by 1 | Viewed by 3766
Abstract
Lipid droplet (LD) accumulation in hepatocytes is one of the major symptoms associated with fatty liver disease. Mitochondria play a key role in catabolizing fatty acids for energy production through β-oxidation. The interplay between mitochondria and LD assumes a crucial role in lipid [...] Read more.
Lipid droplet (LD) accumulation in hepatocytes is one of the major symptoms associated with fatty liver disease. Mitochondria play a key role in catabolizing fatty acids for energy production through β-oxidation. The interplay between mitochondria and LD assumes a crucial role in lipid metabolism, while it is obscure how mitochondrial morphology affects systemic lipid metabolism in the liver. We previously reported that cilnidipine, an already existing anti-hypertensive drug, can prevent pathological mitochondrial fission by inhibiting protein–protein interaction between dynamin-related protein 1 (Drp1) and filamin, an actin-binding protein. Here, we found that cilnidipine and its new dihydropyridine (DHP) derivative, 1,4-DHP, which lacks Ca2+ channel-blocking action of cilnidipine, prevent the palmitic acid-induced Drp1–filamin interaction, LD accumulation and cytotoxicity of human hepatic HepG2 cells. Cilnidipine and 1,4-DHP also suppressed the LD accumulation accompanied by reducing mitochondrial contact with LD in obese model and high-fat diet-fed mouse livers. These results propose that targeting the Drp1–filamin interaction become a new strategy for the prevention or treatment of fatty liver disease. Full article
Show Figures

Figure 1

Figure 1
<p>Treatment with cilnidipine reduces LDs in HepG2. (<b>A</b>–<b>C</b>) Effects of cilnidipine on palmitic acid (PA)-induced LDs. (<b>A</b>) Representative imaging of HepG2 cells treated with 30 μM of PA with or without cilnidipine (1 μM). (<b>B</b>,<b>C</b>) Number (<b>B</b>) and area (<b>C</b>) of LDs for each cell. (<b>D</b>) Representative images of contact between mitochondria (green) and LDs (blue). (<b>E</b>,<b>F</b>) Quantitative results of (<b>D</b>). (<b>E</b>) Size of LDs in each cell (n = 57–108 cells) and (<b>F</b>) area of mitochondria–LD contact (area shown by light blue). (<b>G</b>,<b>H</b>) Effects of cilnidipine on PA-induced (<b>G</b>) cytotoxicity and (<b>H</b>) viability. Cell death was induced by exposure to PA for 24 h. (<b>I</b>) Representative images of the Duolink proximity ligation assay (PLA) between Drp1 and FLNA. PLA signals are shown as white spots (yellow arrowhead) counterstained with phalloidin (green) and DAPI (blue). (<b>J</b>) Number of PLA signals for each cell with or without cilnidipine treatment (&gt;150 cells). Data are means ± SEM (n = 3 in each group). Significance was determined using one-way ANOVA followed by Tukey’s comparison test. Scale bars: 10 μm (<b>A</b>,<b>D</b>,<b>I</b>).</p>
Full article ">Figure 2
<p>Treatment with cilnidipine improves liver injury and LDs in ob/ob mice. (<b>A</b>–<b>D</b>) Effect of cilnidipine on plasma levels of ALT (<b>A</b>), AST (<b>B</b>), TCHO (<b>C</b>) and TG (<b>D</b>) (n = 5 mice in each group). (<b>E</b>) Representative TEM images of mouse livers. (<b>F</b>) Quantitative result of LD areas using ImageJ (Version 1.54g) n = 30 cells in each group). (<b>G</b>) Representative TEM images of mitochondria–LD contact (shown by red line). Individual mitochondria and LD are marked in blue line. (<b>H</b>) Quantitative result of mitochondria–LD contact using ImageJ (n = 30 cells in each group). Data are means ± SEM. Significance was determined using one-way ANOVA followed by Tukey’s comparison test. Scale bars: 5 μm (<b>E</b>) and 1 μm (<b>G</b>).</p>
Full article ">Figure 3
<p>Treatment with cilnidipine improves HFD-induced LD accumulation. (<b>A</b>–<b>D</b>) Effect of cilnidipine on plasma levels of ALT (<b>A</b>), AST (<b>B</b>) TCHO (<b>C</b>) and TG (<b>D</b>) (n = 10 mice in each group). (<b>E</b>) Representative TEM images of livers. Scale bar: 5 μm. (<b>F</b>) Quantitative result of LD formation using ImageJ (n = 30 cells in each group). Data are means ± SEM. Significance was determined using student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>1,4-DHP reduces PA-induced LD accumulation in HepG2. (<b>A</b>) Chemical structures of cilnidipine and 1,4-DHP. (<b>B</b>–<b>D</b>) Effects of 1,4-DHP on PA-induced LDs. (<b>B</b>) Representative imaging of HepG2 cells treated with 30 μM of PA with or without 1,4-DHP in HepG2. Number (<b>C</b>) and area (<b>D</b>) of LDs for each cell. (<b>E</b>) Representative images of contact between mitochondria (green) and LDs (blue). (<b>F</b>,<b>G</b>) Quantitative results of (<b>E</b>). (<b>F</b>) Size of LDs in each cell (n = 63–108 cells) and (<b>G</b>) the area of mitochondria–LD contact (shown by merged color (light blue)). (<b>H</b>) Representative PLA images between Drp1 and FLNA. PLA signals are shown as white spots (yellow arrowhead) counterstained with phalloidin (green) and DAPI (blue). (<b>I</b>) Number of PLA signals for each cell with or without 1,4-DHP treatment (&gt;150 cells). Data are means ± SEM (n = 3–5 in each group). Significance was determined using one-way ANOVA followed by Tukey’s comparison test. Scale bars: 10 μm (<b>B</b>,<b>E</b>,<b>H</b>).</p>
Full article ">Figure 5
<p>Treatment with 1,4-DHP improves liver injury and LDs in ob/ob mice fed HFD. (<b>A</b>–<b>D</b>) Effect of cilnidipine or 1,4-DHP on plasma levels of ALT (<b>A</b>), AST (<b>B</b>), TCHO (<b>C</b>) and TG (<b>D</b>) (n = 5 mice in each group). (<b>E</b>) Representative TEM images of livers. (<b>F</b>) Quantitative data of steatosis using ImageJ. (<b>G</b>) Representative TEM images of mitochondria–LD contact. (<b>H</b>) Quantitative data of mitochondria–LD contact using ImageJ (n = 30 cells in each group). Data are means ± SEM (n = 30 cells in each group). Significance was determined using one-way ANOVA followed by Tukey’s comparison test. Scale bars: 5 μm (<b>E</b>) and 1 μm (<b>G</b>).</p>
Full article ">Figure 6
<p>Knockdown of Drp1 and FLNA suppresses LD formation in HepG2 cells. (<b>A</b>–<b>C</b>) Effect of Drp1 siRNA knockdown on LD accumulation of HepG2 cells exposed to PA (30 μM) for 24 h. (<b>A</b>) Representative images of HepG2 cells treated with or without Drp1 siRNA. Average number (<b>B</b>) and area (<b>C</b>) of LDs in each cell (n = 65–115 cells per experiment). (<b>D</b>–<b>F</b>) Effect of FLNA siRNA on PA-induced LD formation. (<b>D</b>) Representative images of HepG2 cells treated with or without FLNA siRNA. Average number (<b>E</b>) and area (<b>F</b>) of LDs in each cell (n = 70–129 cells per experiment). NC: negative control. Data are means ± SEM (n = 3 in each group). Significance was determined using one-way ANOVA followed by Tukey’s comparison test. Scale bars: 10 μm (<b>A</b>,<b>D</b>).</p>
Full article ">Figure 7
<p>Cilnidipine improves PA-induced respiratory dysfunction of HepG2 cells. (<b>A</b>) Experimental scheme of FAO assay. (<b>B</b>) Oxygen consumption rate (OCR) of HepG2. Cells were pretreated with 200 μM PA or BSA for 24 h with or without 1 μM cilnidipine in substrate-limited D-MEM media supplemented; cells were changed in FAO assay medium and incubated in non-CO<sub>2</sub> free incubator. Cells were stimulated with PA or BSA just before OCR measurement. Group names were defined by the order in which reagents were added. OL: oligomycin, FCCP: carbonyl cyanide <span class="html-italic">p</span>-(trifluoromethoxy) phenylhydrazone, ROT: Rotenone, ANT: antimycin. (<b>C</b>–<b>F</b>) Average basal respiration (<b>C</b>), maximal respiration (<b>D</b>), ATP production (<b>E</b>) and spare capacity (<b>F</b>). Data are means ± SEM (n = 3 in each group). Significance was determined using one-way ANOVA followed by Tukey’s comparison test.</p>
Full article ">
19 pages, 11203 KiB  
Article
Lemon Peel Water Extract: A Novel Material for Retinal Health, Protecting Retinal Pigment Epithelial Cells against Dynamin-Related Protein 1-Mediated Mitochondrial Fission by Blocking ROS-Stimulated Mitogen-Activated Protein Kinase/Extracellular Signal-Regulated Kinase Pathway
by Shang-Chun Tsou, Chen-Ju Chuang, Inga Wang, Tzu-Chun Chen, Jui-Hsuan Yeh, Chin-Lin Hsu, Yu-Chien Hung, Ming-Chung Lee, Yuan-Yen Chang and Hui-Wen Lin
Antioxidants 2024, 13(5), 538; https://doi.org/10.3390/antiox13050538 - 27 Apr 2024
Viewed by 1854
Abstract
Previous studies showed that NaIO3 can induce oxidative stress-mediated retinal pigment epithelium (RPE) damage to simulate age-related macular degeneration (AMD). Lemon peel is rich in antioxidants and components that can penetrate the blood–retinal barrier, but their role in retinal oxidative damage remains [...] Read more.
Previous studies showed that NaIO3 can induce oxidative stress-mediated retinal pigment epithelium (RPE) damage to simulate age-related macular degeneration (AMD). Lemon peel is rich in antioxidants and components that can penetrate the blood–retinal barrier, but their role in retinal oxidative damage remains unexplored. Here, we explore the protection of lemon peel ultrasonic-assisted water extract (LUWE), containing large amounts of flavonoids and polyphenols, against NaIO3-induced retinal degeneration. We initially demonstrated that LUWE, orally administered, prevented retinal distortion and thinning on the inner and outer nuclei layers, downregulating cleaved caspase-3 protein expression in RPE cells in NaIO3-induced mice. The effect of LUWE was achieved through the suppression of apoptosis and the associated proteins, such as cleaved PARP and cleaved caspase-3, as suggested by NaIO3-induced ARPE-19 cell models. This is because LUWE reduced reactive oxygen species-mediated mitochondrial fission via regulating p-Drp-1 and Fis1 expression. We further confirmed that LUWE suppresses the expression of p-MEK-1/2 and p-ERK-1/2 in NaIO3-induced ARPE-19 cells, thereby providing the protection described above, which was confirmed using PD98059 and U0126. These results indicated that LUWE prevents mitochondrial oxidative stress-mediated RPE damage via the MEK/ERK pathway. Elucidation of the molecular mechanism may provide a new protective strategy against retinal degeneration. Full article
Show Figures

Figure 1

Figure 1
<p>The 3D-HPLC chromatographic fingerprint analysis of LUWE. The chemical compositions of the LUWE samples were analyzed using high-performance liquid chromatography (HPLC) with a photodiode array detector (PDA). Twenty-four polyphenolic or flavonoid compounds were detected and are displayed in <a href="#antioxidants-13-00538-t002" class="html-table">Table 2</a> according to their corresponding peak numbers. AU = arbitrary perfusion units.</p>
Full article ">Figure 2
<p>Effects of LUWE on retinal histological changes in Balb/c mice induced by NaIO<sub>3</sub>. All mice were sacrificed after 7 days of intravenous injection. Mouse eyeballs were (<b>A</b>) stained with hematoxylin and eosin to quantify the thickness of the (<b>B</b>) retina, (<b>C</b>) ONL, (<b>D</b>) INL, or (<b>E</b>) immunohistochemically stained with cleaved caspase-3 antibody. Scale bar = 100 µm. The red arrows marked irregular deformations at the base of the retina, while the yellow arrows indicated migrated cell nucleus. All data on thickness were expressed as mean ± standard deviation (<span class="html-italic">n</span> = 6). Different letters (a,b) superscripted in the statistical chart indicate that there is statistical significance between groups (<span class="html-italic">p</span> &lt; 0.05); on the contrary, being marked with the same letter indicates that there is no statistical significance.</p>
Full article ">Figure 3
<p>Protection of LUWE in NaIO<sub>3</sub>-induced ARPE-19 cell apoptosis. ARPE-19 cells were (<b>A</b>) treated with different doses of LUWE, or (<b>B</b>) pretreated with LUWE as designed for 1.5 h, and then co-cultured with 6 mM NaIO<sub>3</sub> for 24 h to detect cell viability. (<b>C</b>) The cell morphology was recorded (bar = 10 µm), and (<b>D</b>) annexin V/PI staining was performed. The percentage of apoptotic cells is the sum of early apoptotic cells (Q5-LR: annexin V+/PI−) and late apoptotic cells (Q5-UR: annexin V+/PI+). (<b>E</b>) Cell lysates were collected for analysis of caspase-3, cleaved caspase-3, and cleaved PARP protein expression by Western blotting. Protein expression is represented by the fold of the mock group. All data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Different letters (a–d) superscripted in the statistical chart indicate that there is statistical significance between groups (<span class="html-italic">p</span> &lt; 0.05); on the contrary, being marked with the same letter indicates that there is no statistical significance.</p>
Full article ">Figure 4
<p>Effects of LUWE on the mitochondrial ROS level and MAPK proteins induced by NaIO<sub>3</sub>. After treatment according to the experimental design, ARPE-19 cells in each group (<b>A</b>) were analyzed for mitochondrial ROS levels using MitoSOX Red staining or were collected for analyzing (<b>B</b>) the p-ERK, p-p38, p-JNK2, and (<b>C</b>) p-MEK protein expression by Western blotting. MitoSOX Red in each group was quantified as a percentage compared with the fluorescent values of the mock group. Protein expression was represented by the fold of the mock group. All data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Different letters (a–d) superscripted in the statistical chart indicate that there is statistical significance between groups (<span class="html-italic">p</span> &lt; 0.05); on the contrary, being marked with the same letter indicates that there is no statistical significance.</p>
Full article ">Figure 5
<p>Effects of LUWE, U0126, and PD98059 on NaIO<sub>3</sub>-induced ARPE-19 cell apoptosis. ARPE-19 cells were pretreated with 10 µM U0126 or 20 µM PD98059 mixed with 2.5 mg/ mL LUWE or not for 1.5 h, and then cotreated with NaIO<sub>3</sub> for 24 h. (<b>A</b>) The cell viability was analyzed using the CCK-8 reagent, and (<b>B</b>) the ratio of apoptotic cells was determined by annexin V/PI staining. Cell lysates were collected for analysis of (<b>C</b>) p-MEK-1/2, p-ERK-1/2, (<b>D</b>) cleaved caspase-9 PARP, and cleaved PARP protein expression by Western blotting. Protein expression was represented by the fold of the mock group. All data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Different letters (a–d) superscripted in the statistical chart indicate that there is statistical significance between groups (<span class="html-italic">p</span> &lt; 0.05); on the contrary, being marked with the same letter indicates that there is no statistical significance.</p>
Full article ">Figure 6
<p>Effects of LUWE, U0126, and PD98059 on NaIO<sub>3</sub>-induced mitochondrial homeostasis. ARPE-19 cells were pretreated with 10 µM U0126 or 20 µM PD98059 mixed with 2.5 mg/mL LUWE or not for 1.5 h, and then cotreated with sodium iodate for 24 h. (<b>A</b>) Cell lysates were collected for analysis of p-Drp-1, Fis1, and cytochrome c protein expression by Western blotting. Protein expression was represented by the fold of the mock group. (<b>B</b>) The isolated mitochondria from cells in each group were analyzed for particle size using forward scatter (FSC-A) with flow cytometry. All data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Different letters (a–d) superscripted in the statistical chart indicate that there is statistical significance between groups (<span class="html-italic">p</span> &lt; 0.05); on the contrary, being marked with the same letter indicates that there is no statistical significance.</p>
Full article ">Figure 7
<p>Protection of LUWE against sodium iodate-induced apoptosis in RPE cells. LUWE mitigates mitochondrial ROS-mediated MEK/ERK signaling, consequently reducing mitochondrial fission regulated by Drp-1 and Fis1. As a result, LUWE suppresses the expression of cytochrome c protein, leading to a reduction in RPE cell apoptosis regulated through the caspase-9/caspase-3/PARP cascade. LUWE protected against sodium iodate-induced RPE cell death, thereby reducing retinal thinning and distortion.</p>
Full article ">
13 pages, 726 KiB  
Review
OMA1-Mediated Mitochondrial Dynamics Balance Organellar Homeostasis Upstream of Cellular Stress Responses
by Robert Gilkerson, Harpreet Kaur, Omar Carrillo and Isaiah Ramos
Int. J. Mol. Sci. 2024, 25(8), 4566; https://doi.org/10.3390/ijms25084566 - 22 Apr 2024
Viewed by 1583
Abstract
In response to cellular metabolic and signaling cues, the mitochondrial network employs distinct sets of membrane-shaping factors to dynamically modulate organellar structures through a balance of fission and fusion. While these organellar dynamics mediate mitochondrial structure/function homeostasis, they also directly impact critical cell-wide [...] Read more.
In response to cellular metabolic and signaling cues, the mitochondrial network employs distinct sets of membrane-shaping factors to dynamically modulate organellar structures through a balance of fission and fusion. While these organellar dynamics mediate mitochondrial structure/function homeostasis, they also directly impact critical cell-wide signaling pathways such as apoptosis, autophagy, and the integrated stress response (ISR). Mitochondrial fission is driven by the recruitment of the cytosolic dynamin-related protein-1 (DRP1), while fusion is carried out by mitofusins 1 and 2 (in the outer membrane) and optic atrophy-1 (OPA1) in the inner membrane. This dynamic balance is highly sensitive to cellular stress; when the transmembrane potential across the inner membrane (Δψm) is lost, fusion-active OPA1 is cleaved by the overlapping activity with m-AAA protease-1 (OMA1 metalloprotease, disrupting mitochondrial fusion and leaving dynamin-related protein-1 (DRP1)-mediated fission unopposed, thus causing the collapse of the mitochondrial network to a fragmented state. OMA1 is a unique regulator of stress-sensitive homeostatic mitochondrial balance, acting as a key upstream sensor capable of priming the cell for apoptosis, autophagy, or ISR signaling cascades. Recent evidence indicates that higher-order macromolecular associations within the mitochondrial inner membrane allow these specialized domains to mediate crucial organellar functionalities. Full article
(This article belongs to the Special Issue Mitochondria as a Cellular Hub in Neurological Disorders 2.0)
Show Figures

Figure 1

Figure 1
<p><b>Mitochondria balance between fission and fusion.</b> Mitochondrial fission, or division, is carried out by the recruitment of the cytosolic GTPase DRP1. Conversely, the fusion of the inner membrane is carried out by OPA1. Under cellular stress, OMA1 metalloprotease cleaves fusion-active OPA1, leaving fission unopposed. SHSY5Y cells visualized by anti-TOM20 immunolabeling, Nikon AX confocal.</p>
Full article ">Figure 2
<p><b>Schematic of OMA1 and OPA1 macromolecular associations within the inner membrane.</b> In addition to the regulated maintenance of oxidative phosphorylation complexes into supercomplexes and dimeric assemblies of the F<sub>1</sub>F<sub>0</sub> ATP synthase, increasing evidence indicates that other critical mitochondrial functionalities may be carried out by domains of interacting factors within the inner membrane. Prohibitin-associated SLP2 interacts with both the mitochondrial ribosome and inner-membrane proteases PARL and YME1L as part of the SPY complex (<b>left</b>), while OMA1 interacts with the MICOS complex as part of contact sites between the outer and inner membrane (<b>center</b>). Independently, OPA1 associates with MICOS and SLC25 to mediate crista junction formation (<b>right</b>).</p>
Full article ">
23 pages, 11243 KiB  
Article
The CaMK Family Differentially Promotes Necroptosis and Mouse Cardiac Graft Injury and Rejection
by Haitao Lu, Jifu Jiang, Jeffery Min, Xuyan Huang, Patrick McLeod, Weihua Liu, Aaron Haig, Lakshman Gunaratnam, Anthony M. Jevnikar and Zhu-Xu Zhang
Int. J. Mol. Sci. 2024, 25(8), 4428; https://doi.org/10.3390/ijms25084428 - 17 Apr 2024
Viewed by 1049
Abstract
Organ transplantation is associated with various forms of programmed cell death which can accelerate transplant injury and rejection. Targeting cell death in donor organs may represent a novel strategy for preventing allograft injury. We have previously demonstrated that necroptosis plays a key role [...] Read more.
Organ transplantation is associated with various forms of programmed cell death which can accelerate transplant injury and rejection. Targeting cell death in donor organs may represent a novel strategy for preventing allograft injury. We have previously demonstrated that necroptosis plays a key role in promoting transplant injury. Recently, we have found that mitochondria function is linked to necroptosis. However, it remains unknown how necroptosis signaling pathways regulate mitochondrial function during necroptosis. In this study, we investigated the receptor-interacting protein kinase 3 (RIPK3) mediated mitochondrial dysfunction and necroptosis. We demonstrate that the calmodulin-dependent protein kinase (CaMK) family members CaMK1, 2, and 4 form a complex with RIPK3 in mouse cardiac endothelial cells, to promote trans-phosphorylation during necroptosis. CaMK1 and 4 directly activated the dynamin-related protein-1 (Drp1), while CaMK2 indirectly activated Drp1 via the phosphoglycerate mutase 5 (PGAM5). The inhibition of CaMKs restored mitochondrial function and effectively prevented endothelial cell death. CaMKs inhibition inhibited activation of CaMKs and Drp1, and cell death and heart tissue injury (n = 6/group, p < 0.01) in a murine model of cardiac transplantation. Importantly, the inhibition of CaMKs greatly prolonged heart graft survival (n = 8/group, p < 0.01). In conclusion, CaMK family members orchestrate cell death in two different pathways and may be potential therapeutic targets in preventing cell death and transplant injury. Full article
Show Figures

Figure 1

Figure 1
<p>The CaMK family participates in MVEC necroptosis. (<b>A</b>) Cells (20 × 10<sup>3</sup>/well) were seeded in quadruplicates in a 96-well plate. Cell death was induced by TNFα (T, 20 ng/mL) with Smac mimetic BV6 (S, 2 μM). Apoptosis was inhibited by caspase-8 inhibitor z-IETD (I, 30 μM). Necroptosis was inhibited by RIPK1 inhibitor Nec-1s (N, 10 μM). CaMK was inhibited by KN93 (K, 20 μg/mL). Cell death was detected by SYTOX Green uptake into the dead cell from 0 to 24 h by IncuCyte Image system (Essen Bioscience, Ann Arbor, MI, USA). (<b>B</b>) SYTOX uptake was quantified at 24 h. Data are shown as mean ± standard deviation (SD) of quadruplicates and representative of three independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001. <span class="html-italic">t</span>-test. (<b>C</b>) Expression of CaMK1, CaMK2, and CaMK4 was quantified by real time PCR after cell death induction for 4 h. β-actin was used as endogenous control for mRNA expression. Data are shown as mean ± SD of three independent experiments. Western blot analysis of CaMK1 and p-CaMK1 (<b>D</b>,<b>E</b>), CaMK2 and p-CaMK2 (<b>F</b>,<b>G</b>), and CaMK4 and p-CaMK4 (<b>H</b>,<b>I</b>). Cells were collected for Western blot analysis 5 h after cell death induction. GAPDH was used as loading control. Images were quantified by densitometry (ImageJ 1.54g). Relative ratio = phosphorylated protein/total protein. Data are shown as the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 1 Cont.
<p>The CaMK family participates in MVEC necroptosis. (<b>A</b>) Cells (20 × 10<sup>3</sup>/well) were seeded in quadruplicates in a 96-well plate. Cell death was induced by TNFα (T, 20 ng/mL) with Smac mimetic BV6 (S, 2 μM). Apoptosis was inhibited by caspase-8 inhibitor z-IETD (I, 30 μM). Necroptosis was inhibited by RIPK1 inhibitor Nec-1s (N, 10 μM). CaMK was inhibited by KN93 (K, 20 μg/mL). Cell death was detected by SYTOX Green uptake into the dead cell from 0 to 24 h by IncuCyte Image system (Essen Bioscience, Ann Arbor, MI, USA). (<b>B</b>) SYTOX uptake was quantified at 24 h. Data are shown as mean ± standard deviation (SD) of quadruplicates and representative of three independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001. <span class="html-italic">t</span>-test. (<b>C</b>) Expression of CaMK1, CaMK2, and CaMK4 was quantified by real time PCR after cell death induction for 4 h. β-actin was used as endogenous control for mRNA expression. Data are shown as mean ± SD of three independent experiments. Western blot analysis of CaMK1 and p-CaMK1 (<b>D</b>,<b>E</b>), CaMK2 and p-CaMK2 (<b>F</b>,<b>G</b>), and CaMK4 and p-CaMK4 (<b>H</b>,<b>I</b>). Cells were collected for Western blot analysis 5 h after cell death induction. GAPDH was used as loading control. Images were quantified by densitometry (ImageJ 1.54g). Relative ratio = phosphorylated protein/total protein. Data are shown as the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 2
<p>CaMKs participate in necroptosis. CaMK2δ silencing in MVECs was confirmed by PCR 16 h after siRNA treatment (<b>A</b>) and Western blot (<b>B</b>) 24 h after siRNA treatment. Untreated cells (UT) or vehicle control (VC, transfection reagent) treated cells were used as controls. GAPDH from the same blot was used as the loading control. Data are pooled and represent three independent experiments. (<b>C</b>) CaMK2δ-siRNA or VC treated cells were harvested after 24 h and subjected to the cell death assay. SYTOX uptake/cell death was monitored by IncuCyte Image system. Data are shown as mean ± SD of quadruplicates and represent three independent experiments. siRNA-induced silencing of CaMK1 (<b>D</b>,<b>E</b>) and CaMK4 (<b>F</b>,<b>G</b>) in MVECs was confirmed by PCR and Western blot analysis. (<b>H</b>) CaMK1, CaMK4, or CaMK1+4 siRNAs or vehicle control (VC, EndoFectin) treated cells were harvested after 24 h and subjected to the cell death assay. SYTOX uptake was monitored for 24 h by IncuCyte Image system. Data are shown as mean ± SD of quadruplicates at 24 h and represent three independent experiments. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span>≤0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 2 Cont.
<p>CaMKs participate in necroptosis. CaMK2δ silencing in MVECs was confirmed by PCR 16 h after siRNA treatment (<b>A</b>) and Western blot (<b>B</b>) 24 h after siRNA treatment. Untreated cells (UT) or vehicle control (VC, transfection reagent) treated cells were used as controls. GAPDH from the same blot was used as the loading control. Data are pooled and represent three independent experiments. (<b>C</b>) CaMK2δ-siRNA or VC treated cells were harvested after 24 h and subjected to the cell death assay. SYTOX uptake/cell death was monitored by IncuCyte Image system. Data are shown as mean ± SD of quadruplicates and represent three independent experiments. siRNA-induced silencing of CaMK1 (<b>D</b>,<b>E</b>) and CaMK4 (<b>F</b>,<b>G</b>) in MVECs was confirmed by PCR and Western blot analysis. (<b>H</b>) CaMK1, CaMK4, or CaMK1+4 siRNAs or vehicle control (VC, EndoFectin) treated cells were harvested after 24 h and subjected to the cell death assay. SYTOX uptake was monitored for 24 h by IncuCyte Image system. Data are shown as mean ± SD of quadruplicates at 24 h and represent three independent experiments. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span>≤0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 3
<p>RIPK3 and CaMKs form a complex during necroptosis. Cells were induced to undergo necroptosis, as shown in <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a>, and collected after 4 h. (<b>A</b>) Untreated and vehicle control treated cells were used as co-immunoprecipitation controls. Cell lysates were immunoprecipitated with CaMK1, CaMK4, and RIPK3 antibodies, respectively, and followed by Western blot analysis to detect CaMK2. (<b>B</b>) Cell lysates were immunoprecipitated with CaMK1, CaMK2, and CaMK4 antibodies, respectively. Rabbit IgG were used as control. The immunoprecipitants were used to detect RIPK3 in Western blot analysis. (<b>C</b>) Cell lysates were immunoprecipitated with CaMK1 and CaMK4 antibodies. The immunoprecipitants were used to detect CaMK2 in Western blot analysis. (<b>D</b>) CaMK1 siRNA or vehicle treated cells were harvested after 24 h for cell death induction and then collected after 4 h. Cell lysates were immunoprecipitated with CaMK2 antibody or rabbit IgG. The immunoprecipitants were used to detect CaMK4 in Western blot analysis. Data (<b>B</b>–<b>D</b>) represent three independent experiments. (<b>E</b>) Cell death was induced as described in <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a>. ATP level was quantified by the CellTiter-Glo<sup>®</sup> Luminescent Cell Viability kit. Data are shown as mean ± SD of three independent experiments. (<b>F</b>) Mitochondria were probed by MitoTracker. Fluorescent intensity was automatically quantified by the IncuCyte System. Data are shown as mean ± SD of three independent experiments. *** <span class="html-italic">p</span> ≤ 0.001; **** <span class="html-italic">p</span> ≤ 0.0001, 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 3 Cont.
<p>RIPK3 and CaMKs form a complex during necroptosis. Cells were induced to undergo necroptosis, as shown in <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a>, and collected after 4 h. (<b>A</b>) Untreated and vehicle control treated cells were used as co-immunoprecipitation controls. Cell lysates were immunoprecipitated with CaMK1, CaMK4, and RIPK3 antibodies, respectively, and followed by Western blot analysis to detect CaMK2. (<b>B</b>) Cell lysates were immunoprecipitated with CaMK1, CaMK2, and CaMK4 antibodies, respectively. Rabbit IgG were used as control. The immunoprecipitants were used to detect RIPK3 in Western blot analysis. (<b>C</b>) Cell lysates were immunoprecipitated with CaMK1 and CaMK4 antibodies. The immunoprecipitants were used to detect CaMK2 in Western blot analysis. (<b>D</b>) CaMK1 siRNA or vehicle treated cells were harvested after 24 h for cell death induction and then collected after 4 h. Cell lysates were immunoprecipitated with CaMK2 antibody or rabbit IgG. The immunoprecipitants were used to detect CaMK4 in Western blot analysis. Data (<b>B</b>–<b>D</b>) represent three independent experiments. (<b>E</b>) Cell death was induced as described in <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a>. ATP level was quantified by the CellTiter-Glo<sup>®</sup> Luminescent Cell Viability kit. Data are shown as mean ± SD of three independent experiments. (<b>F</b>) Mitochondria were probed by MitoTracker. Fluorescent intensity was automatically quantified by the IncuCyte System. Data are shown as mean ± SD of three independent experiments. *** <span class="html-italic">p</span> ≤ 0.001; **** <span class="html-italic">p</span> ≤ 0.0001, 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 4
<p>CaMKs are responsible for Drp1 activation. (<b>A</b>) Cell death was induced as described in <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a>. Drp1 inhibitor Midivi-1 (50 μM) was added. Cell death was detected by IncuCyte Image system. SYTOX uptake was quantified at 24 h. Data are shown as mean ± SD of quadruplicates and represent three independent experiments. (<b>B</b>) Drp1siRNA or vehicle-treated cells were harvested after 24 h and subjected to the cell death assay. Data are shown as mean ± SD of quadruplicates and represent three independent experiments. **** <span class="html-italic">p</span> ≤ 0.0001, 1-way ANOVA; Tukey’s multiple comparisons. (<b>C</b>) p-Drp1 Western blot. Cell death was induced as described in <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a>. Drp1 inhibitor Midivi-1 or CaMKs inhibitor KN93 was added. Cells were collected for Western blot analysis 4 h after cell death induction. Untreated (UT) cells were used as controls. (<b>D</b>) Images were quantified by ImageJ. Relative Ratio of protein level = p-Drp1/Total Drp1. Data are shown as mean ± SD of three independent experiments. (<b>E</b>) CaMK1, CaMK2, CaMK4, CaMK1+4, and CaKM1+2+4 siRNAs-treated cells were harvested after 24 h for the cell death assay. Cells were collected after 4 h for Western blot analysis of p-Drp1 (S616). (<b>F</b>) Images were quantified by ImageJ. GAPDH was used to normalize protein levels. Data are shown as mean ± SD of three independent experiments. **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 5
<p>CaMK2 indirectly binds to Drp1 via PGAM5 while CaMK1 and CaMK4 directly bind to Drp1 without PGAM5. (<b>A</b>) Cells death was induced as <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a> and cells were collected 4 h after. Cell lysates were immunoprecipitated with CaMK1, CaMK2, and CaMK4 antibodies. Rabbit IgG was used as isotype control. The immunoprecipitants were used to detect PGAM5 in Western blot analysis. (<b>B</b>) Untreated and vehicle control treated cells were used as co-immunoprecipitation controls. Cell lysates were immunoprecipitated with CaMK1, CaMK2, CaMK4 and PGAM5 antibodies, respectively, and followed by Western blot analysis to detect Drp1. (<b>C</b>) PGAM5 siRNA- or vehicle control (VC)-treated cells were harvested after 24 h for the cell death assay. Four hours after, cell lysates were immunoprecipitated with anti-CaMK2 or control IgG. The immunoprecipitants were used to detect Drp1 in Western blot analysis. (<b>D</b>) CaMKs siRNAs- or vehicle control-treated cells were harvested after 24 h and used in the cell death assay. Cells were collected after 4 h and cell lysates were immunoprecipitated with PGAM5 antibody or rabbit IgG. Immunoprecipitants were used to detect Drp1 in Western blot. (<b>E</b>) Images were quantified by ImageJ. The relative level of Drp1 was calculated against the Drp1 level of necroptotic cells (TSI treated) in vehicle control (VC) group. Data are shown as mean ± SD of three independent experiments. **** <span class="html-italic">p</span> ≤ 0.0001; <span class="html-italic">t</span>-test. (<b>F</b>) PGAM5 siRNA- or VC-treated cells were harvested after 24 h and used in the cell death assay. Cells were collected after 4 h and cell lysates were immunoprecipitated with CaMK1 antibody or control IgG. The immunoprecipitants were used for anti-Drp1 detection in Western blot analysis. (<b>G</b>) CaMK1, CaMK2, or PGAM5 siRNA-treated cells were harvested after 24 h for the cell death assay. Cells were collected after 4 h and immunoprecipitated with CaMK4 antibody or control IgG. The immunoprecipitants were used for anti-Drp1 detection in Western blot analysis. (<b>H</b>) Images were quantified by ImageJ. The relative level of Drp1 was calculated against the Drp1 level of necroptotic cells (TSI treated) in vehicle control (VC) group. Data are shown as mean ± SD of three independent experiments. **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 5 Cont.
<p>CaMK2 indirectly binds to Drp1 via PGAM5 while CaMK1 and CaMK4 directly bind to Drp1 without PGAM5. (<b>A</b>) Cells death was induced as <a href="#ijms-25-04428-f001" class="html-fig">Figure 1</a> and cells were collected 4 h after. Cell lysates were immunoprecipitated with CaMK1, CaMK2, and CaMK4 antibodies. Rabbit IgG was used as isotype control. The immunoprecipitants were used to detect PGAM5 in Western blot analysis. (<b>B</b>) Untreated and vehicle control treated cells were used as co-immunoprecipitation controls. Cell lysates were immunoprecipitated with CaMK1, CaMK2, CaMK4 and PGAM5 antibodies, respectively, and followed by Western blot analysis to detect Drp1. (<b>C</b>) PGAM5 siRNA- or vehicle control (VC)-treated cells were harvested after 24 h for the cell death assay. Four hours after, cell lysates were immunoprecipitated with anti-CaMK2 or control IgG. The immunoprecipitants were used to detect Drp1 in Western blot analysis. (<b>D</b>) CaMKs siRNAs- or vehicle control-treated cells were harvested after 24 h and used in the cell death assay. Cells were collected after 4 h and cell lysates were immunoprecipitated with PGAM5 antibody or rabbit IgG. Immunoprecipitants were used to detect Drp1 in Western blot. (<b>E</b>) Images were quantified by ImageJ. The relative level of Drp1 was calculated against the Drp1 level of necroptotic cells (TSI treated) in vehicle control (VC) group. Data are shown as mean ± SD of three independent experiments. **** <span class="html-italic">p</span> ≤ 0.0001; <span class="html-italic">t</span>-test. (<b>F</b>) PGAM5 siRNA- or VC-treated cells were harvested after 24 h and used in the cell death assay. Cells were collected after 4 h and cell lysates were immunoprecipitated with CaMK1 antibody or control IgG. The immunoprecipitants were used for anti-Drp1 detection in Western blot analysis. (<b>G</b>) CaMK1, CaMK2, or PGAM5 siRNA-treated cells were harvested after 24 h for the cell death assay. Cells were collected after 4 h and immunoprecipitated with CaMK4 antibody or control IgG. The immunoprecipitants were used for anti-Drp1 detection in Western blot analysis. (<b>H</b>) Images were quantified by ImageJ. The relative level of Drp1 was calculated against the Drp1 level of necroptotic cells (TSI treated) in vehicle control (VC) group. Data are shown as mean ± SD of three independent experiments. **** <span class="html-italic">p</span> ≤ 0.0001; 1-way ANOVA; Tukey’s multiple comparisons.</p>
Full article ">Figure 6
<p>p-CaMK1, p-CaMK2, p-CaMK4, and p-Drp1 (S616) increases were inhibited by KN93 in the graft post heart transplantation. (<b>A</b>) B6-to-BALB/c heart transplantation and KN93 injection was performed as detailed in the Methods. The grafts (n = 3) were collected after 3 days for Western blot analysis by CaMK1, p-CaMK1, CaMK2, p-CaMK2, CaMK4, p-CaMK4, Drp1 and p-Drp1 (S616) antibodies, respectively. (<b>B</b>–<b>E</b>) Images were quantified by ImageJ. Relative ratio of protein = phosphorylated protein/total protein. Data are shown as mean ± SD of 3 transplants. **** <span class="html-italic">p</span> ≤ 0.0001; Student <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>Inhibition of CaMKs attenuates heart transplant acute injury. (<b>A</b>) B6-to-BALB/c heart transplantation and KN93 or saline injection were performed as detailed in <a href="#sec4-ijms-25-04428" class="html-sec">Section 4</a>. Grafts (n = 6/group) were collected after 3 days for H&amp;E staining. Images were taken under 200 times magnification. (<b>B</b>) Graft injuries were scored for lymphocyte infiltration, infarction, and PMN infiltration from 0 to 5 in a blinded fashion. Scores were averaged as mean ± SD of 6 grafts. (<b>C</b>) Grafts (n = 6) were used for immunohistochemistry with anti-CD45, and positive areas (brown color) are indicated by red arrows. Images were taken under 200 times magnification. (<b>D</b>) Positive areas of each graft were automatically counted in six connected random areas under 200 times magnification by Image J and averaged in a double-blinded manner. (<b>E</b>) Grafts (n = 6) were assessed by TUNEL. B6 naive hearts were used as control. Brown color indicates TUNEL positive cells as indicated by red arrows. Images are at 200 times magnification. (<b>F</b>) Necroptosis in the graft was detected by p-MLKL immunohistochemistry. Images are at 200 times magnification. Positive cells are indicated by red arrows. (<b>G</b>) TUNEL positive areas were quantified as above. (<b>H</b>) p-MLKL positive areas were quantified as above. ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001. <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7 Cont.
<p>Inhibition of CaMKs attenuates heart transplant acute injury. (<b>A</b>) B6-to-BALB/c heart transplantation and KN93 or saline injection were performed as detailed in <a href="#sec4-ijms-25-04428" class="html-sec">Section 4</a>. Grafts (n = 6/group) were collected after 3 days for H&amp;E staining. Images were taken under 200 times magnification. (<b>B</b>) Graft injuries were scored for lymphocyte infiltration, infarction, and PMN infiltration from 0 to 5 in a blinded fashion. Scores were averaged as mean ± SD of 6 grafts. (<b>C</b>) Grafts (n = 6) were used for immunohistochemistry with anti-CD45, and positive areas (brown color) are indicated by red arrows. Images were taken under 200 times magnification. (<b>D</b>) Positive areas of each graft were automatically counted in six connected random areas under 200 times magnification by Image J and averaged in a double-blinded manner. (<b>E</b>) Grafts (n = 6) were assessed by TUNEL. B6 naive hearts were used as control. Brown color indicates TUNEL positive cells as indicated by red arrows. Images are at 200 times magnification. (<b>F</b>) Necroptosis in the graft was detected by p-MLKL immunohistochemistry. Images are at 200 times magnification. Positive cells are indicated by red arrows. (<b>G</b>) TUNEL positive areas were quantified as above. (<b>H</b>) p-MLKL positive areas were quantified as above. ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001. <span class="html-italic">t</span>-test.</p>
Full article ">Figure 8
<p>Inhibition of CaMKs attenuates heart transplant chronic injury and rejection. (<b>A</b>) B6-to-BALB/c heart transplantation and anti-CD154 injection are detailed in <a href="#sec4-ijms-25-04428" class="html-sec">Section 4</a>. KN93 or saline was injected on day 1, 2, and 3 followed by every 48 h until 21 days post transplantation. Recipient mice (n = 4/group) were euthanized, and the grafts were collected for H&amp;E and elastin-trichrome staining. Images were taken under 200 times magnification. Representative images are shown. (<b>B</b>) Graft injuries were quantified blindly by a pathologist. Scores were averaged as mean ± SD of 4 grafts. ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001, <span class="html-italic">t</span>-test. Grafts were assessed by immunohistochemistry for anti-CD3 (<b>C</b>), anti-IgG (<b>E</b>) and anti-Foxp3 (<b>G</b>) and positive staining areas (brown color) are indicated by red arrows. Images were taken under 200 times magnification. Positive areas anti-CD3 (<b>D</b>), anti-IgG (<b>F</b>) and anti-FoxP3 (<b>H</b>) of each graft were automatically counted in six connected random areas under 200 times magnification by Image J and averaged in a double-blinded fashion. **** <span class="html-italic">p</span> ≤ 0.0001. <span class="html-italic">t</span>-test. (<b>I</b>) B6-to-BALB/c heart transplantation using KN93 or saline administration was performed as above. Graft survival was monitored daily. Cessation of beating is considered as rejection. n = 8 per group, ** <span class="html-italic">p</span> = 0.004. Log Rank test.</p>
Full article ">Figure 8 Cont.
<p>Inhibition of CaMKs attenuates heart transplant chronic injury and rejection. (<b>A</b>) B6-to-BALB/c heart transplantation and anti-CD154 injection are detailed in <a href="#sec4-ijms-25-04428" class="html-sec">Section 4</a>. KN93 or saline was injected on day 1, 2, and 3 followed by every 48 h until 21 days post transplantation. Recipient mice (n = 4/group) were euthanized, and the grafts were collected for H&amp;E and elastin-trichrome staining. Images were taken under 200 times magnification. Representative images are shown. (<b>B</b>) Graft injuries were quantified blindly by a pathologist. Scores were averaged as mean ± SD of 4 grafts. ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001, <span class="html-italic">t</span>-test. Grafts were assessed by immunohistochemistry for anti-CD3 (<b>C</b>), anti-IgG (<b>E</b>) and anti-Foxp3 (<b>G</b>) and positive staining areas (brown color) are indicated by red arrows. Images were taken under 200 times magnification. Positive areas anti-CD3 (<b>D</b>), anti-IgG (<b>F</b>) and anti-FoxP3 (<b>H</b>) of each graft were automatically counted in six connected random areas under 200 times magnification by Image J and averaged in a double-blinded fashion. **** <span class="html-italic">p</span> ≤ 0.0001. <span class="html-italic">t</span>-test. (<b>I</b>) B6-to-BALB/c heart transplantation using KN93 or saline administration was performed as above. Graft survival was monitored daily. Cessation of beating is considered as rejection. n = 8 per group, ** <span class="html-italic">p</span> = 0.004. Log Rank test.</p>
Full article ">
16 pages, 3310 KiB  
Article
Altered Glycolysis, Mitochondrial Biogenesis, Autophagy and Apoptosis in Peritoneal Endometriosis in Adolescents
by Elena P. Khashchenko, Mikhail Yu. Vysokikh, Maria V. Marey, Ksenia O. Sidorova, Ludmila A. Manukhova, Natalya N. Shkavro, Elena V. Uvarova, Vladimir D. Chuprynin, Timur Kh. Fatkhudinov, Leila V. Adamyan and Gennady T. Sukhikh
Int. J. Mol. Sci. 2024, 25(8), 4238; https://doi.org/10.3390/ijms25084238 - 11 Apr 2024
Cited by 1 | Viewed by 1432
Abstract
Energy metabolism plays a pivotal role in the pathogenesis of endometriosis. For the initial stages of the disease in adolescents, this aspect remains unexplored. The objective of this paper was to analyze the association of cellular and endosomal profiles of markers of glycolysis, [...] Read more.
Energy metabolism plays a pivotal role in the pathogenesis of endometriosis. For the initial stages of the disease in adolescents, this aspect remains unexplored. The objective of this paper was to analyze the association of cellular and endosomal profiles of markers of glycolysis, mitochondrial biogenesis, apoptosis, autophagy and estrogen signaling in peritoneal endometriosis (PE) in adolescents. We included 60 girls aged 13–17 years in a case–control study: 45 with laparoscopically confirmed PE (main group) and 15 with paramesonephric cysts (comparison group). Samples of plasma and peritoneal fluid exosomes, endometrioid foci and non-affected peritoneum were tested for estrogen receptor (Erα/β), hexokinase (Hex2), pyruvate dehydrogenase kinase (PDK1), glucose transporter (Glut1), monocarboxylate transporters (MCT1 and MCT2), optic atrophy 1 (OPA1, mitochondrial fusion protein), dynamin-related protein 1 (DRP1, mitochondrial fission protein), Bax, Bcl2, Beclin1, Bnip3, P38 mitogen-activated protein kinase (MAPK), hypoxia-inducible factor 1 (Hif-1α), mitochondrial voltage-dependent anion channel (VDAC) and transforming growth factor (TGFβ) proteins as markers of estrogen signaling, glycolysis rates, mitochondrial biogenesis and damage, apoptosis and autophagy (Western-Blot and PCR). The analysis identified higher levels of molecules associated with proliferation (ERβ), glycolysis (MCT2, PDK1, Glut1, Hex2, TGFβ and Hif-1α), mitochondrial biogenesis (OPA1, DRP1) and autophagy (P38, Beclin1 and Bnip3) and decreased levels of apoptosis markers (Bcl2/Bax) in endometrioid foci compared to non-affected peritoneum and that in the comparison group (p < 0.05). Patients with PE had altered profiles of ERβ in plasma and peritoneal fluid exosomes and higher levels of Glut1, MCT2 and Bnip3 in plasma exosomes (p < 0.05). The results of the differential expression profiles indicate microenvironment modification, mitochondrial biogenesis, estrogen reception activation and glycolytic switch along with apoptosis suppression in peritoneal endometrioid foci already in adolescents. Full article
(This article belongs to the Special Issue Molecular Research in Gynecological Diseases)
Show Figures

Figure 1

Figure 1
<p>Comparative analysis of estrogen receptor levels in exosomes and in tissues of patients in the studied groups. Levels of exosomal ERβ in peripheral blood (<b>A</b>) and peritoneal fluid (<b>B</b>) in patients with endometriosis were significantly higher than those in the comparison group. (<b>C</b>) ERβ and ERα protein levels (left and right panels, respectively) in the peritoneal tissues of patients with endometriosis were higher in endometrioid foci vs. intact peritoneum. Data of relative protein level quantification are presented as boxes with median, interquartile range and min–max values. Representative Western blots are presented on panels below graphs. See also corresponding Ponceau-stained images in the <a href="#app1-ijms-25-04238" class="html-app">Figure S1A–C of Supplemental Materials</a>, original Western Blot in <a href="#app1-ijms-25-04238" class="html-app">Figure S7</a>.</p>
Full article ">Figure 2
<p>Estimation of energy metabolism markers level in peritoneal tissues. (<b>A</b>) Expression of Glut1 in peritoneal tissue biopsies according to RT-PCR data (see <a href="#sec4-ijms-25-04238" class="html-sec">Section 4</a> for details). (<b>B</b>) Western blot analysis for glycolysis markers Glut1, Hex2, Hif1α, MCT1, MCT2, PDK1 and also VDAC1 and TGFβ. Data of relative protein level quantification and gene expression are presented as boxes with median, interquartile range and min–max values. Representative corresponding Western blots are presented below graphs on panels. See also Ponceau-stained images in <a href="#app1-ijms-25-04238" class="html-app">Figure S2 of Supplemental Materials</a>, original Western Blot in <a href="#app1-ijms-25-04238" class="html-app">Figure S7</a>.</p>
Full article ">Figure 3
<p>Comparative analysis of glucose and lactate transporter proteins in blood exosomes of patients of studied groups. Data of relative protein level quantification are presented as boxes with median, interquartile range and min–max values. Representative Western blots are presented below corresponding graphs on panels. See also Ponceau-stained images in <a href="#app1-ijms-25-04238" class="html-app">Figure S3 of Supplemental Materials</a>, original Western Blot in <a href="#app1-ijms-25-04238" class="html-app">Figure S7</a>.</p>
Full article ">Figure 4
<p>Mitochondrial fission and fusion markers DRP1 and OPA1 levels are changed in patients with EMS. Reliable decrease in DRP1 transcripts (<b>A</b>) and Drp1 protein (<b>B</b>) levels in the foci biopsies of patients with EMS and increase in blood exosomes (<b>C</b>). (<b>D</b>) Relative expression of OPA1 protein increased for EMS patients compared to the intact peritoneum biopsies of patients of the comparison group. Relative data of protein level quantification and gene expression are presented as boxes with median, interquartile range and min–max values. Representative Western blots are presented below graphs. See also corresponding Ponceau-stained images in <a href="#app1-ijms-25-04238" class="html-app">Figure S4A,B of Supplemental Materials</a>, original Western Blot in <a href="#app1-ijms-25-04238" class="html-app">Figure S7</a>.</p>
Full article ">Figure 5
<p>Expression of apoptotic markers in biopsies of peritoneal tissues. (<b>A</b>) Increased expression of the anti-apoptotic <span class="html-italic">Bcl2</span> and decreased expression of pro-apoptotic <span class="html-italic">Bax</span> in the foci vs. intact peritoneum of both groups. <span class="html-italic">Bcl-2/Bax</span> transcripts ratio is reliably high in EMS foci vs. intact peritoneum of the comparison group. (<b>B</b>) Levels of Bcl2 and Bax proteins in the peritoneal tissues. Data of relative protein level and gene expression are presented as boxes with median, interquartile range and min–max values. Representative Western blots are presented below graphs. See corresponding Ponceau-stained images in <a href="#app1-ijms-25-04238" class="html-app">Figure S5 of Supplemental Materials</a>, original Western Blot in <a href="#app1-ijms-25-04238" class="html-app">Figure S7</a>.</p>
Full article ">Figure 6
<p>Levels of autophagy markers in peritoneal tissue biopsies and blood exosomes of patients with EMS. (<b>A</b>) Decreased level of p38 protein and increased level of Beclin1 protein in EMS foci. (<b>B</b>) Decreased exosomal level of BNIP3 in the blood of patients with EMS vs. comparison group. Data of relative proteins level quantification are presented as boxes with median, interquartile range and min–max values. Representative Western blots are presented below graphs. See corresponding Ponceau-stained images in <a href="#app1-ijms-25-04238" class="html-app">Figure S6A,B of Supplemental Materials</a>, original Western Blot in <a href="#app1-ijms-25-04238" class="html-app">Figure S7</a>.</p>
Full article ">Figure 7
<p>Activation of aerobic glycolysis and mitochondrial biogenesis under the conditions of increased estrogen reception in the pathogenesis of peritoneal endometriosis in adolescents. Expression levels for the key effectors of glycolysis, from the uptake of glucose by membrane transporters of the GLUT family (notably GLUT1) and its phosphorylation to glucose-6-phosphate by hexokinase (Hex2), are increased. The enhanced phosphorylation of pyruvate dehydrogenase (PDH) by its kinase (PDK1) limits pyruvate conversion to acetyl-CoA, thereby disconnecting glycolysis from tricarboxylic acid cycle (TCA) in mitochondria and reinforcing pyruvate conversion to lactate by lactate dehydrogenase A (LDHA). Lactate transport out of the cell through its transporters MCT is also increased. At the same time, a decrease in the activity of oxidative phosphorylation (TCA) in mitochondria is associated with less electron leakage and reactive oxygen species (ROS) formation and, accordingly, with the control of oxidative stress. Mitochondrial biogenesis is also increased, along with stabilization of mitochondrial cristae and enhanced apoptosis resistance (fusion marker OPA1 and fission marker DRP1). Contacts with EPR (glucose-regulated protein (GRP75) facilitate mitochondria-associated ER membrane (MAM) formation) and Ca<sup>2+</sup> influx (VDAC) reinforce activation of cholesterol synthesis and steroidogenesis pathways. The implementation of ER-β signals leads to changes (p38 MAPK kinase cascade) in the expression of nuclear (p53) and mitochondrial genes, particularly those responsible for protection against apoptosis (the ratio of Bcl2 to Bax increases) and autophagy activation (Beclin, Bnip). HIF-1α and TGFβ signaling has a positive feedback loop with glycolytic switch and associated cell reprogramming.</p>
Full article ">
24 pages, 3974 KiB  
Article
The Effect of Cold-Water Swimming on Energy Metabolism, Dynamics, and Mitochondrial Biogenesis in the Muscles of Aging Rats
by Mateusz Bosiacki, Maciej Tarnowski, Kamila Misiakiewicz-Has and Anna Lubkowska
Int. J. Mol. Sci. 2024, 25(7), 4055; https://doi.org/10.3390/ijms25074055 - 5 Apr 2024
Viewed by 1337
Abstract
Our study aimed to explore the potential positive effects of cold water exercise on mitochondrial biogenesis and muscle energy metabolism in aging rats. The study involved 32 male and 32 female rats aged 15 months, randomly assigned to control sedentary animals, animals training [...] Read more.
Our study aimed to explore the potential positive effects of cold water exercise on mitochondrial biogenesis and muscle energy metabolism in aging rats. The study involved 32 male and 32 female rats aged 15 months, randomly assigned to control sedentary animals, animals training in cold water at 5 ± 2 °C, or animals training in water at thermal comfort temperature (36 ± 2 °C). The rats underwent swimming training for nine weeks, gradually increasing the duration of the sessions from 2 min to 4 min per day, five days a week. The results demonstrated that swimming in thermally comfortable water improved the energy metabolism of aging rat muscles (increased metabolic rates expressed as increased ATP, ADP concentration, TAN (total adenine nucleotide) and AEC (adenylate energy charge value)) and increased mRNA and protein expression of fusion regulatory proteins. Similarly, cold-water swimming improved muscle energy metabolism in aging rats, as shown by an increase in muscle energy metabolites and enhanced mitochondrial biogenesis and dynamics. It can be concluded that the additive effect of daily activity in cold water influenced both an increase in the rate of energy metabolism in the muscles of the studied animals and an intensification of mitochondrial biogenesis and dynamics (related to fusion and fragmentation processes). Daily activity in warm water also resulted in an increase in the rate of energy metabolism in muscles, but at the same time did not cause significant changes in mitochondrial dynamics. Full article
(This article belongs to the Section Molecular Endocrinology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>ATP, ADP, AMP, and Ado in the muscles of male (<b>A</b>) and female (<b>B</b>) rats from the control group and the experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 1 Cont.
<p>ATP, ADP, AMP, and Ado in the muscles of male (<b>A</b>) and female (<b>B</b>) rats from the control group and the experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 2
<p>TAN and AEC in the muscles of male rats (<b>A</b>) and female rats (<b>B</b>) from the control group and experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 3
<p>Expression of PGC-1α mRNA (<b>I</b>) and ELISA method protein analysis (<b>II</b>) in the muscles of male rats (<b>A</b>) and female rats (<b>B</b>) from the control group and experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 4
<p>Expression of Mfn1 mRNA (<b>I</b>) and ELISA method protein analysis (<b>II</b>) in the muscles of male rats (<b>A</b>) and female rats (<b>B</b>) from the control group and experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## is <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 5
<p>Expression of Mfn2 mRNA (<b>I</b>) and ELISA method protein analysis (<b>II</b>) in the muscles of male rats (<b>A</b>) and female rats (<b>B</b>) from the control group and experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), # is <span class="html-italic">p</span> &lt; 0.05 level of significance compared to the 5 °C group (Mann–Whitney U test), ## is <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 6
<p>Representative microphotography showing immunoexpression of Mfn1 in muscles of male and female rats from the control group and experimental groups. The immunopositive reaction’s area appears as brown colored precipitates within muscle fiber. The color intensity of the precipitate indicates the level of immunoexpression of Mfn1 detected using IHC reaction. Scale bar 20 μm (objective magnification ×100).</p>
Full article ">Figure 7
<p>Representative microphotography showing immunoexpression of Mfn2 in the skeletal muscles of male and female rats from the control group and experimental groups. The immunopositive reaction’s area appears as brown colored precipitates within muscle fiber. The color intensity of the precipitate indicates the level of immunoexpression of Mfn2 detected using IHC reaction. Scale bar 20 μm (objective magnification ×100).</p>
Full article ">Figure 8
<p>Expression of Opa1 mRNA (<b>I</b>) and ELISA method protein analysis (<b>II</b>) in the muscles of male rats (<b>A</b>) and female rats (<b>B</b>) from the control group and experimental groups. The results are presented as means and standard deviations. ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## is <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">Figure 9
<p>Expression of Drp1 mRNA (<b>I</b>) and ELISA method protein analysis (<b>II</b>) in the muscles of male rats (<b>A</b>) and female rats (<b>B</b>) from the control group and experimental groups. The results are presented as means and standard deviations; ** <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the control group (Mann–Whitney U test), ## is <span class="html-italic">p</span> &lt; 0.005 level of significance compared to the 5 °C group (Mann–Whitney U test).</p>
Full article ">
19 pages, 4503 KiB  
Article
Oral Antiviral Defense: Saliva- and Beverage-like Hypotonicity Dynamically Regulate Formation of Membraneless Biomolecular Condensates of Antiviral Human MxA in Oral Epithelial Cells
by Pravin B. Sehgal, Huijuan Yuan, Anthony Centone and Susan V. DiSenso-Browne
Cells 2024, 13(7), 590; https://doi.org/10.3390/cells13070590 - 28 Mar 2024
Viewed by 1153
Abstract
The oral mucosa represents a defensive barrier between the external environment and the rest of the body. Oral mucosal cells are constantly bathed in hypotonic saliva (normally one-third tonicity compared to plasma) and are repeatedly exposed to environmental stresses of tonicity, temperature, and [...] Read more.
The oral mucosa represents a defensive barrier between the external environment and the rest of the body. Oral mucosal cells are constantly bathed in hypotonic saliva (normally one-third tonicity compared to plasma) and are repeatedly exposed to environmental stresses of tonicity, temperature, and pH by the drinks we imbibe (e.g., hypotonic: water, tea, and coffee; hypertonic: assorted fruit juices, and red wines). In the mouth, the broad-spectrum antiviral mediator MxA (a dynamin-family large GTPase) is constitutively expressed in healthy periodontal tissues and induced by Type III interferons (e.g., IFN-λ1/IL-29). Endogenously induced human MxA and exogenously expressed human GFP-MxA formed membraneless biomolecular condensates in the cytoplasm of oral carcinoma cells (OECM1 cell line). These condensates likely represent storage granules in equilibrium with antivirally active dispersed MxA. Remarkably, cytoplasmic MxA condensates were exquisitely sensitive sensors of hypotonicity—the condensates in oral epithelium disassembled within 1–2 min of exposure of cells to saliva-like one-third hypotonicity, and spontaneously reassembled in the next 4–7 min. Water, tea, and coffee enhanced this disassembly. Fluorescence changes in OECM1 cells preloaded with calcein-AM (a reporter of cytosolic “macromolecular crowding”) confirmed that this process involved macromolecular uncrowding and subsequent recrowding secondary to changes in cell volume. However, hypertonicity had little effect on MxA condensates. The spontaneous reassembly of GFP-MxA condensates in oral epithelial cells, even under continuous saliva-like hypotonicity, was slowed by the protein-phosphatase-inhibitor cyclosporin A (CsA) and by the K-channel-blocker tetraethylammonium chloride (TEA); this is suggestive of the involvement of the volume-sensitive WNK kinase-protein phosphatase (PTP)-K-Cl cotransporter (KCC) pathway in the regulated volume decrease (RVD) during condensate reassembly in oral cells. The present study identifies a novel subcellular consequence of hypotonic stress in oral epithelial cells, in terms of the rapid and dynamic changes in the structure of one class of phase-separated biomolecular condensates in the cytoplasm—the antiviral MxA condensates. More generally, the data raise the possibility that hypotonicity-driven stresses likely affect other intracellular functions involving liquid–liquid phase separation (LLPS) in cells of the oral mucosa. Full article
(This article belongs to the Section Cellular Immunology)
Show Figures

Figure 1

Figure 1
<p>IFN-λ1 (IL-29), but not IFN-α2, induces expression of MxA in oral epithelial cells (OECM1 cell line). (<b>A</b>,<b>B</b>) 35 mm cultures of OECM1 and A549 cells were left untreated or treated with human IFN-α2 or human IFN-λ1 (50 ng/mL for 2 days) followed by fixation (4% PFA for 1 h at 4 °C) and immunofluorescence imaging for MxA; (<b>C</b>) OECM1 cultures treated with IFN-λ1 for 2 days were exposed to 5% 1,6-hexanediol in PBS, or left untreated, and then fixed and imaged for MxA; (<b>D</b>) Western blot of extracts (30 µg/lane) prepared from parallel plates as in (<b>A</b>); (<b>E</b>) OECM1 cultures were transfected with pGFP-MxA expression vector and imaged in PBS 2 days later, followed by treatment with 5%-Hex and imaging of the same cells 5 min later. All scale bars = 10 µm.</p>
Full article ">Figure 2
<p>Single-cell antiviral phenotype (protected against VSV; white arrows) of OECM1 cells expressing GFP-MxA mainly in condensates (<b>A</b>) or mainly in the dispersed phase (<b>B</b>). Cultures in 35 mm plates were transiently transfected with pGFP-MxA vector, and two days later were challenged with VSV (moi &gt; 10 pfu/cell) and fixed 24 h after the start of infection [<a href="#B31-cells-13-00590" class="html-bibr">31</a>]. VSV replication was assessed by immunostaining for the VSV nucleocapsid (N) protein (in red) [<a href="#B19-cells-13-00590" class="html-bibr">19</a>,<a href="#B24-cells-13-00590" class="html-bibr">24</a>]. White arrows point to GFP-containing cells with reduced VSV-N. Scale bars = 20 µm. (<b>C</b>) Schematic highlighting the dynamic equilibrium between GFP-MxA in condensed vs. dispersed phase (note from Figure 4B, and additional Figures below, that even in cells with visually “mainly” condensed GFP-MxA, there is 15–25% of GFP-MxA in the dispersed phase, which can be still antivirally active as in (<b>B</b>) above).</p>
Full article ">Figure 3
<p>Reversible osmosensing by GFP-MxA condensates in oral epithelial cells—focus on beverage-like hypotonicity. (<b>A</b>) Sequential live-cell imaging of the same OECM1 cells expressing GFP-MxA condensates 2 days after transient transfection first in isotonic PBS (300 mOsm) and then after shifting to hypotonic ELB (40 mOsm). (<b>B</b>) the same cells as in (<b>A</b>) were sequentially imaged after further shifting back to isotonic PBS (300 mOsm). White arrows, formations of vacuole-like dilations [<a href="#B29-cells-13-00590" class="html-bibr">29</a>] prior to condensate formation. (<b>C</b>) IFN-λ1 (50 ng/mL for 2 days) treated OECM1 cultures were fixed after washing with PBS, or after 5 min in ELB, or after 5 min in ELB, and then returned back to PBS for 5 min. Cultures were fixed using 4% PFA and immunostained for MxA. All scale bars = 10 µm.</p>
Full article ">Figure 4
<p>Spontaneously reversible osmosensing by GFP-MxA condensates in oral epithelial cells-focus on saliva-like hypotonicity. (<b>A</b>) Sequential live-cell imaging of the same OECM1 cell expressing GFP-MxA condensates 2 days after transient transfection first in full-culture medium (330 mOsm) and then after shifting to hypotonic of one-third tonicity (110 mOsm; full medium diluted 1:2 with water) for the next 8–10 min. Scale bar = 10 µm. (<b>B</b>) Quantitation of GFP-MxA in condensates on a % per-cell basis in the images shown in (<b>A</b>). This quantitation was carried out using the small object subtract Filter in Image J [<a href="#B20-cells-13-00590" class="html-bibr">20</a>,<a href="#B24-cells-13-00590" class="html-bibr">24</a>].</p>
Full article ">Figure 5
<p>Rapid disassembly of GFP-MxA condensates by drinking water, tea, and coffee. OECM1 cultures expressing GFP-MxA condensates (2 days after transfection) were first imaged in full medium (approx. 330 mOsm) and then shifted to one-third tonicity saliva-like medium (100 mOsm) for 1–2.4 h to allow completion of the disassembly and reassembly cycle as shown in <a href="#cells-13-00590-f004" class="html-fig">Figure 4</a>. Single live cells in the respective cultures were then imaged and the imaging continued upon shifting the cultures to drinking water (<b>A</b>), tea (<b>B</b>), coffee (<b>C</b>). Scale bars = 10 µm.</p>
Full article ">Figure 6
<p>Testing a biophysical basis for hypotonicity sensing by GFP-MxA condensates in oral epithelial cells using calcein quenching as a reporter for macromolecular crowding. (<b>A</b>) OECM1 cells were preloaded with calcein-AM (2 µM in PBS) for 15 min, washed 4× with PBS and then imaged. The culture was then shifted to hypotonic ELB for 2 min and imaged immediately using the same fluorescence settings. The culture was then shifted to isotonic PBS for 9 min and cells imaged. Areas within white dashed boxes are shown at higher magnification in the lower panels. Scale bar = 10 µm. (<b>B</b>) Calcein fluorescence on a per-cell basis (in arbitrary units) is depicted. ****, <span class="html-italic">p</span> &lt; 0.0001; ns, not significant; VLD, vacuole-like dilatations; n, number of cells quantitated.</p>
Full article ">Figure 7
<p>Testing a biochemical basis for hypotonicity sensing by GFP-MxA condensates in oral epithelial cells. (<b>A</b>,<b>B</b>), OECM1 cells were either exposed to CsA (25 µM) or DMSO alone for 20 min in full-culture medium, and then shifted to one-third tonicity medium in the continued presence of CsA. Live-cell imaging was carried out as indicated. Scale bar = 10 µm. (<b>C</b>) Quantitation of % GFP-MxA per cell in condensates (in the same cells shown in (<b>A</b>,<b>B</b>)).</p>
Full article ">Figure 8
<p>Testing a biochemical basis for hypotonicity sensing by GFP-MxA condensates in oral epithelial cells. (<b>A</b>,<b>B</b>) OECM1 cells were either kept in full-culture medium or exposed to TEA (20 mM) in full-culture medium for 20 min. Cultures were then shifted to one-third tonicity medium in the continued presence of TEA. Live-cell imaging was carried out as indicated. Scale bar = 10 µm. (<b>C</b>) Quantitation of % GFP-MxA per cell in condensates (in the same cells shown in (<b>A</b>,<b>B</b>)).</p>
Full article ">Figure 9
<p><b>Hypothesis:</b> Overview of the biophysical and biochemical mechanisms possibly involved in the cell-volume-driven dynamic regulation of the formation, disassembly, and reassembly of MxA condensates in oral epithelial cells subjected to saliva- and beverage-like hypotonicity. 2-DG, 2-deoxyglucose; KCC, potassium-chloride cotransporter channels 1–4; MLO, membraneless organelle, TEA, tetraethylammonium chloride; RVD, regulated volume decrease; VLD, vacuoele-like dilatations; WNK kinase, “With no lysine” kinase family members 1–4.</p>
Full article ">
15 pages, 8124 KiB  
Article
Silybin Alleviated Hepatic Injury by Regulating Redox Balance, Inflammatory Response, and Mitochondrial Function in Weaned Piglets under Paraquat-Induced Oxidative Stress
by Long Cai, Dongxu Ming, Wenning Chen, Ying Zhao, Yanpin Li, Wenjuan Sun, Yu Pi, Xianren Jiang and Xilong Li
Antioxidants 2024, 13(3), 324; https://doi.org/10.3390/antiox13030324 - 6 Mar 2024
Cited by 5 | Viewed by 1864
Abstract
Silybin (Si) is the main element of silymarin isolated from the seeds of Silybum marianum L. Gaernt., which has superior antioxidant properties. However, the protective role of Si in maintaining liver health under oxidative stress remains ambiguous. This study aimed to investigate the [...] Read more.
Silybin (Si) is the main element of silymarin isolated from the seeds of Silybum marianum L. Gaernt., which has superior antioxidant properties. However, the protective role of Si in maintaining liver health under oxidative stress remains ambiguous. This study aimed to investigate the underlying mechanism of the beneficial effect of dietary Si against hepatic oxidative injury induced by paraquat (PQ) in weaned piglets. A total of 24 piglets were randomly allocated to four treatments with six replicates per treatment and 1 piglet per replicate: the control group; Si group; PQ group; and Si + PQ group. Piglets in the control group and PQ group were given a basal diet, while piglets in the Si and Si + PQ groups were given a Si-supplemented diet. On the 18th day, the pigs in the PQ treatment group received an intraperitoneal injection of PQ, and the others were intraperitoneally injected with the same volume of saline. All piglets were sacrificed on day 21 for plasma and liver sample collection. The results showed that dietary Si supplementation mitigated PQ-induced liver damage, as proven by the reduction in liver pathological changes and plasma activity of alanine transaminase and aspartate transaminase. Si also improved superoxide dismutase and glutathione peroxidase activities and total antioxidant capacity, as well as decreased malondialdehyde and hydrogen peroxide concentration in the liver, which were closely related to the activation of the nuclear factor-erythroid 2-related factor 2 signaling pathway. Meanwhile, Si reduced tumor necrosis factor-α and interleukin-8 production and their transcript levels as well as abrogated the overactivation of nuclear factor-κB induced by PQ. Importantly, Si improved mitochondrial function by maintaining mitochondrial energetics and mitochondrial dynamics, which was indicated by the elevated activity of mitochondrial complexes I and V and adenosine triphosphate content, decreased expression of dynamin 1 protein, and increased expression of mitofusin 2 protein. Moreover, Si inhibited excessive hepatic apoptosis by regulating the B-cell lymphoma-2 (Bcl-2)/Bcl-2-associated-X-protein signaling pathway. Taken together, these results indicated that Si potentially mitigated PQ-induced hepatic oxidative insults by improving antioxidant capacity and mitochondrial function and inhibiting inflammation and cell apoptosis in weaned piglets. Full article
(This article belongs to the Special Issue Novel Antioxidants for Animal Nutrition)
Show Figures

Figure 1

Figure 1
<p>Dietary silybin supplementation alleviated liver injury induced by paraquat in piglets. (<b>A</b>) Representative H&amp;E staining of liver tissues (captured at 100× or 400× magnification). (<b>B</b>) Histopathology scores of liver tissues. (<b>C</b>–<b>E</b>) Plasma activities of AST, ALT, and ALP. Ctrl, piglets were given a basal diet and were challenged with saline; Si, piglets were given a silybin-supplemented diet and were challenged with saline; PQ, piglets were given a basal diet and were challenged with paraquat; Si + PQ, piglets were given a silybin-supplemented diet and were challenged with paraquat; AST, aspartate transaminase; ALT, alanine transaminase; ALP, alkaline phosphatase. Data are expressed as mean ± standard error (n = 6). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Dietary silybin administration ameliorated PQ-induced oxidative stress in the liver of piglets. The activities of CAT (<b>A</b>), SOD (<b>B</b>), and GSH-Px (<b>C</b>) and the level of T-AOC (<b>D</b>), MDA content (<b>E</b>), and H<sub>2</sub>O<sub>2</sub> level (<b>F</b>) in the liver. Ctrl, piglets were given a basal diet and were challenged with saline; Si, piglets were given a silybin-supplemented diet and were challenged with saline; PQ, piglets were given a basal diet and were challenged with paraquat; Si + PQ, piglets were given a silybin-supplemented diet and were challenged with paraquat; CAT, catalase; SOD, superoxide dismutase; GSH-Px, glutathione peroxidase; T-AOC, the total antioxidant capacity; MDA, malondialdehyde; H<sub>2</sub>O<sub>2</sub>, hydrogen peroxide. Data are expressed as mean ± standard error (n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Silybin enhanced hepatic antioxidant capacity by activating the Nrf2 signaling pathway. The expression of Nrf2 signaling pathway genes including <span class="html-italic">CAT</span> (<b>A</b>), <span class="html-italic">SOD1</span> (<b>B</b>), <span class="html-italic">GPX1</span> (<b>C</b>), and <span class="html-italic">GPX4</span> (<b>D</b>). (<b>E</b>–<b>G</b>) The relative protein expression level of Nrf2 and Keap1. Ctrl, piglets were given a basal diet and were challenged with saline; Si, piglets were given a silybin-supplemented diet and were challenged with saline; PQ, piglets were given a basal diet and were challenged with paraquat; Si + PQ, piglets were given a silybin-supplemented diet and were challenged with paraquat; <span class="html-italic">CAT</span>, catalase; <span class="html-italic">SOD</span>, superoxide dismutase; <span class="html-italic">GPx</span>, glutathione peroxidase; <span class="html-italic">Nrf2</span>, nuclear factor-erythroid 2-related factor 2; <span class="html-italic">Keap1</span>, kelch-like ECH-associated protein l. Data are expressed as mean ± standard error (n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Dietary silybin supplementation alleviated hepatic inflammation induced by paraquat in piglets. The concentration of TNF-α (<b>A</b>), IL-6 (<b>B</b>), IL-8 (<b>C</b>), and IL-10 (<b>D</b>). The mRNA expression level of <span class="html-italic">TNF-α</span> (<b>E</b>), <span class="html-italic">IL-6</span> (<b>F</b>), <span class="html-italic">IL-8</span> (<b>G</b>), and <span class="html-italic">IL-10</span> (<b>H</b>). (<b>I</b>–<b>K</b>) The relative protein expression level of P-NF-κB and NF-κB. Ctrl, piglets were given a basal diet and were challenged with saline; Si, piglets were given a silybin-supplemented diet and were challenged with saline; PQ, piglets were given a basal diet and were challenged with paraquat; Si + PQ, piglets were given a silybin-supplemented diet and were challenged with paraquat; TNF-α, tumor necrosis factor-α; IL, interleukin; NF-κB, nuclear factor-kB; P-NF-κB, phosphorylated NF-κB. Data are expressed as mean ± standard error (n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Dietary silybin supplementation protected against PQ-induced mitochondrial dysfunction. (<b>A</b>,<b>B</b>) The activities of COX Ⅰ and COX Ⅴ in the liver. (<b>C</b>) The level of ATP. (<b>D</b>) The expression of mitochondrial respiratory chain protein complex genes including <span class="html-italic">NDUFS2</span>, <span class="html-italic">NDUFV2</span>, <span class="html-italic">SDHA</span>, <span class="html-italic">UQCRB</span>, and <span class="html-italic">ATP5H</span>. (<b>E</b>–<b>G</b>) The relative protein expression level of Drp1 and Mfn2. Ctrl, piglets were given a basal diet and were challenged with saline; Si, piglets were given a silybin-supplemented diet and were challenged with saline; PQ, piglets were given a basal diet and were challenged with paraquat; Si + PQ, piglets were given a silybin-supplemented diet and were challenged with paraquat; COX, mitochondrial complex; ATP, adenosine triphosphate. <span class="html-italic">NDUFS2</span>, NADH ubiquinone oxidoreductase core subunit S2; <span class="html-italic">NDUFV2</span>, NADH ubiquinone oxidoreductase core subunit V2; <span class="html-italic">SDHA</span>, succinate dehydrogenase complex flavoprotein subunit A; <span class="html-italic">UQCRB</span>, ubiquinol-cytochrome c reductase binding protein; <span class="html-italic">ATP5H</span>, ATP synthase subunit d; Drp1, dynamin 1; Mfn2, mitofusin 2. Data are expressed as mean ± standard error (n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Dietary silybin addition suppressed hepatocyte apoptosis. (<b>A</b>,<b>B</b>) The activities of caspase 3 and caspase 9 in the liver tissues. (<b>C</b>) Representative image of Western blot. The protein expression levels of Cleaved caspase 3 (<b>D</b>), Bcl-2 (<b>E</b>), Bax (<b>F</b>), and the ratio of Bcl-2 to Bax (<b>G</b>). Ctrl, piglets were given a basal diet and were challenged with saline; Si, piglets were given a silybin-supplemented diet and were challenged with saline; PQ, piglets were given a basal diet and were challenged with paraquat; Si + PQ, piglets were given a silybin-supplemented diet and were challenged with paraquat; Bcl-2, B-cell lymphoma-2; Bax, Bcl-2-associated-X-protein; Bcl-2/Bax, the ratio of Bcl-2 to Bax. Data are expressed as mean ± standard error (n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
Back to TopTop