[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (10,231)

Search Parameters:
Keywords = wild type

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3320 KiB  
Article
Enhancing Coleoptile Length of Rice Seeds under Submergence through NAL11 Knockout
by Zhe Zhao, Yuelan Xie, Mengqing Tian, Jinzhao Liu, Chun Chen, Jiyong Zhou, Tao Guo and Wuming Xiao
Plants 2024, 13(18), 2593; https://doi.org/10.3390/plants13182593 (registering DOI) - 17 Sep 2024
Abstract
Submergence stress challenges direct seeding in rice cultivation. In this study, we identified a heat shock protein, NAL11, with a DnaJ domain, which can regulate the length of rice coleoptiles under flooded conditions. Through bioinformatics analyses, we identified cis-regulatory elements in [...] Read more.
Submergence stress challenges direct seeding in rice cultivation. In this study, we identified a heat shock protein, NAL11, with a DnaJ domain, which can regulate the length of rice coleoptiles under flooded conditions. Through bioinformatics analyses, we identified cis-regulatory elements in its promoter, making it responsive to abiotic stresses, such as hypoxia or anoxia. Expression of NAL11 was higher in the basal regions of shoots and coleoptiles during flooding. NAL11 knockout triggered the rapid accumulation of abscisic acid (ABA) and reduction of Gibberellin (GA), stimulating rice coleoptile elongation and contributes to flooding stress management. In addition, NAL11 mutants were found to be more sensitive to ABA treatments. Such knockout lines exhibited enhanced cell elongation for coleoptile extension. Quantitative RT-PCR analysis revealed that NAL11 mediated the gluconeogenic pathway, essential for the energy needed in cell expansion. Furthermore, NAL11 mutants reduced the accumulation of reactive oxygen species (ROS) and malondialdehyde under submerged stress, attributed to an improved antioxidant enzyme system compared to the wild-type. In conclusion, our findings underscore the pivotal role of NAL11 knockout in enhancing the tolerance of rice to submergence stress by elucidating its mechanisms. This insight offers a new strategy for improving resilience against flooding in rice cultivation. Full article
Show Figures

Figure 1

Figure 1
<p>Evolutionary tree of NAL11 homologs with Oryza sativa Japonica, Brachypodium distachyon, Sorghun bicolor, Zea mays, Glycine max, Nymphaea colorata, Arabidopsis, and wild rice (Oryza brachyantha, Oryza glaberrima). XP_015645205.1 is the accession number of NAL11 in NCBI, which is highlighted in the red box.</p>
Full article ">Figure 2
<p>Spatiotemporal expression analysis of <span class="html-italic">NAL11</span>. (<b>A</b>) Transcription levels of <span class="html-italic">NAL11</span> in germinating seeds of ZH11 under aerobic and submerged conditions using quantitative reverse transcription polymerase chain reaction (RT-PCR). Gene expression was normalized to that of <span class="html-italic">OsActin</span>, with relative expression levels represented as fold change relative to the expression level of <span class="html-italic">NAL11</span> at 0 h. (<b>B</b>) Schematic structure of the expression vector for pro<span class="html-italic"><sub>NAL11</sub></span>:GUS. (<b>C</b>) β-glucuronidase (GUS) staining of seeds after 24 h, 48 h, and 72 h of submergence stress. Scale Bar, 0.5 cm. (Data are presented as mean ± SD, <span class="html-italic">n</span> = 5; significant differences were determined by two-tailed Student’s <span class="html-italic">t</span>-tests. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns, no significance). (<b>D</b>) 2000 bp promoter sequence of <span class="html-italic">NAL11</span>.</p>
Full article ">Figure 3
<p>Relative expression level of expansin genes in ZH11 and knockout lines at 48 h after submergence. (Data are presented as mean ± SD, <span class="html-italic">n</span> = 5; significant differences were determined by two-tailed Student’s <span class="html-italic">t</span>-tests. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns, no significance).</p>
Full article ">Figure 4
<p><span class="html-italic">NAL11</span> is involved in the sugar and energy pathway. (<b>A</b>) Quantitative RT-PCR analysis of nine α-amylase family genes in seeds of ZH11 and knockout lines at 48 h after submergence treatment. (<b>B</b>) α-AMS activity at different stages. (<b>C</b>) RT-qPCR analysis of sugar and energy metabolism genes at 48 h after submergence treatment, respectively. (Data are presented as mean ± SD, <span class="html-italic">n</span> = 5; significant differences were determined by two-tailed Student’s <span class="html-italic">t</span>-ests. * <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, ns, no significance).</p>
Full article ">Figure 5
<p>Analysis of MDA and H<sub>2</sub>O<sub>2</sub> content, transcript levels of ROS-production gene, <span class="html-italic">OsRbohA</span> and <span class="html-italic">OsRbohE</span> of WT and transgenic plants under normal and submerged conditions. (<b>A</b>) MDA content. (<b>B</b>) H<sub>2</sub>O<sub>2</sub> content. (<b>C</b>) <span class="html-italic">OsRbohA</span>. (<b>D</b>) <span class="html-italic">OsRbohE</span>. (Data are presented as mean ± SD, <span class="html-italic">n</span> = 5; significant differences were determined by two-tailed Student’s <span class="html-italic">t</span>-tests. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns, no significance).</p>
Full article ">Figure 6
<p>Knockout of <span class="html-italic">NAL11</span> affects the levels of ABA, GA and IAA. (<b>A</b>–<b>C</b>) Content of endogenous ABA, GA and IAA at 24 h, 48 h, and 72 h after submergence. (Data are presented as mean ± SD, <span class="html-italic">n</span> = 5; significant differences were determined by two-tailed Student’s <span class="html-italic">t</span>-tests. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns, no significance).</p>
Full article ">Figure 7
<p>Knockout of <span class="html-italic">NAL11</span> shows sensitivity to ABA. (<b>A</b>) Representative images of coleoptile length in response to different concentrations of ABA (0 µM, 0.001 µM, 0.01 µM, 0.1 µM, 1 µM, and 10 µM, respectively) after 4 d of submergence for both WT and knockout lines. Scale bars: 1 cm. (<b>B</b>) Comparison of coleoptile lengths in WT and knockout lines in response to control (H<sub>2</sub>O) and ABA treatments after 4 d of submergence. (Data are presented as mean ± SD, <span class="html-italic">n</span> = 5 biologically independent samples; significant differences were determined by two-tailed Student’s <span class="html-italic">t</span>-tests. ** <span class="html-italic">p</span> &lt; 0.01, ns, no significance).</p>
Full article ">
14 pages, 4036 KiB  
Article
The Role of TGF-β1 and Mutant SMAD4 on Epithelial-Mesenchymal Transition Features in Head and Neck Squamous Cell Carcinoma Cell Lines
by Michael Bette, Laura Reinhardt, Uyanga Gansukh, Li Xiang-Tischhauser, Haifa Meskeh, Pietro Di Fazio, Malte Buchholz, Boris A. Stuck and Robert Mandic
Cancers 2024, 16(18), 3172; https://doi.org/10.3390/cancers16183172 - 16 Sep 2024
Viewed by 249
Abstract
The aim of the present study was to investigate possible differences in the sensitivity of HNSCC cells to known EMT regulators. Three HNSCC cell lines (UM-SCC-1, -3, -22B) and the HaCaT control keratinocyte cell line were exposed to transforming growth factor beta 1 [...] Read more.
The aim of the present study was to investigate possible differences in the sensitivity of HNSCC cells to known EMT regulators. Three HNSCC cell lines (UM-SCC-1, -3, -22B) and the HaCaT control keratinocyte cell line were exposed to transforming growth factor beta 1 (TGF-β1), a known EMT master regulator, and the cellular response was evaluated by real-time cell analysis (RTCA), Western blot, quantitative PCR, flow cytometry, immunocytochemistry, and the wound closure (scratch) assay. Targeted sequencing on 50 cancer-related genes was performed using the Cancer Hotspot Panel v2. Mutant, and wild type SMAD4 cDNA was used to generate recombinant SMAD4 constructs for expression in mammalian cell lines. The most extensive response to TGF-β1, such as cell growth and migration, β-actin expression, or E-cadherin (CDH1) downregulation, was seen in cells with a more epithelial phenotype. Lower response correlated with higher basal p-TGFβ RII (Tyr424) levels, pointing to a possible autocrine pre-activation of these cell lines. Targeted sequencing revealed a homozygous SMAD4 mutation in the UM-SCC-22B cell line. Furthermore, PCR cloning of SMAD4 cDNA from the same cell line revealed an additional SMAD4 transcript with a 14 bp insertion mutation, which gives rise to a truncated SMAD4 protein. Overexpression of this mutant SMAD4 protein in the highly epithelial control cell line HaCaT resulted in upregulation of TGF-β1 and vimentin. Consistent with previous reports, the invasive and metastatic potential of HNSCC tumor cells appears associated with the level of autocrine secretion of EMT regulators such as TGF-β1, and it could be influenced by exogenous EMT cytokines such as those derived from immune cells of the tumor microenvironment. Furthermore, mutant SMAD4 appears to be a significant contributor to the mesenchymal transformation of HNSCC cells. Full article
(This article belongs to the Section Molecular Cancer Biology)
Show Figures

Figure 1

Figure 1
<p>The effect of TGF-β1 on CDH1 surface expression in HNSCC and HaCaT cell lines. (<b>A</b>) Shown are representative flow cytometry histograms evaluating CDH1 surface expression (FL2-A channel for PE fluorescence) in cell lines treated with or without TGF-β1. (<b>B</b>) Comparison of absolute CDH1 geometric mean fluorescence (GMF) levels before and after TGF-β1 treatment. (<b>C</b>) Same data as in (B) depicting relative CDH1-GMF levels. Data represent the mean ± SD (<span class="html-italic">n</span> = 3), with <span class="html-italic">p</span> &lt; 0.05 considered statistically significant. Statistical differences were indicated as *: <span class="html-italic">p</span> &lt; 0.05, ***: <span class="html-italic">p</span> &lt; 0.001, n.s.: not significant (see also <a href="#app1-cancers-16-03172" class="html-app">Supplementary Figure S2</a>).</p>
Full article ">Figure 2
<p>Differential responsiveness of HNSCC and HaCaT cell lines to exogenous TGF-β1. Real-time cell analysis (RTCA) demonstrates major differences in TGF-β1 mediated cell growth (normalized cell index) induction in the tested cell lines (left). Data represent the mean ± SD (<span class="html-italic">n</span> = 3), with <span class="html-italic">p</span> &lt; 0.05 considered statistically significant. Statistical differences were indicated as *: <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 and ****: <span class="html-italic">p</span> &lt; 0.0001. Shown on the right is the protein expression of CDH1 (E-cadherin), pTGFβ RII (Tyr424), β-tubulin, and β-actin. Uncropped Western blots are shown in <a href="#app1-cancers-16-03172" class="html-app">Figure S3</a>.</p>
Full article ">Figure 3
<p>Effect of TGF-β1 on cell motility. Wound closure and relative coastline length were evaluated 24 and 48 hr after TGF-β1 treatment in HaCaT (<b>A</b>), UM-SCC-3 (<b>B</b>), UM-SCC-1 (<b>C</b>), and UM-SCC-22B (<b>D</b>) cell lines. Data represent the mean ± SD (<span class="html-italic">n</span> = 3), with <span class="html-italic">p</span> &lt; 0.05 considered statistically significant. Statistical differences were indicated as *: <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, n.s.: not significant.</p>
Full article ">Figure 4
<p>The effect of TGF-β1 on the cellular localization of CDH1 and β-actin. Confocal microscopy depicting CDH1 (<b>A</b>) and β-actin (<b>B</b>) expression (both red) in HaCaT, UM-SCC-3, UM-SCC-1, and UM-SCC-22B cell lines in the presence or absence of TGF-β1. DAPI (blue) was used for nuclear counterstaining. (No specific signal is seen in cells treated with anti-mouse IgG, see <a href="#app1-cancers-16-03172" class="html-app">Supplementary Figure S4</a>).</p>
Full article ">Figure 5
<p>Influence of <span class="html-italic">SMAD4wt</span> and <span class="html-italic">SMAD4mut</span> overexpression on EMT-related genes. <span class="html-italic">SMAD4wt</span> and <span class="html-italic">SMAD4mut</span> expressing plasmids were transfected in HaCaT (<b>A</b>) and UM-SCC-22B (<b>B</b>) cell lines. Gene expression changes were compared against control values (<b>A</b>,<b>B</b>). (<b>C</b>) Gene expression ratios of UM-SCC-22B compared with HaCaT were evaluated in control cells and after transfection with <span class="html-italic">SMAD4</span>-expressing plasmids. (<b>D</b>) Comparison of basal mRNA expression levels of <span class="html-italic">SMAD4</span>, <span class="html-italic">CDH1</span>, <span class="html-italic">TGF-β1</span>, <span class="html-italic">VIM,</span> and <span class="html-italic">ZEB1</span> between HaCaT and UM-SCC-22B cells (data correspond to the controls (ctrl) as shown in <b>A</b>–<b>C</b>). Data represent the mean ± SD (<span class="html-italic">n</span> = 3–4), with <span class="html-italic">p</span> &lt; 0.05 considered statistically significant. Statistical differences were indicated as *: <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, n.s.: not significant.</p>
Full article ">Figure 5 Cont.
<p>Influence of <span class="html-italic">SMAD4wt</span> and <span class="html-italic">SMAD4mut</span> overexpression on EMT-related genes. <span class="html-italic">SMAD4wt</span> and <span class="html-italic">SMAD4mut</span> expressing plasmids were transfected in HaCaT (<b>A</b>) and UM-SCC-22B (<b>B</b>) cell lines. Gene expression changes were compared against control values (<b>A</b>,<b>B</b>). (<b>C</b>) Gene expression ratios of UM-SCC-22B compared with HaCaT were evaluated in control cells and after transfection with <span class="html-italic">SMAD4</span>-expressing plasmids. (<b>D</b>) Comparison of basal mRNA expression levels of <span class="html-italic">SMAD4</span>, <span class="html-italic">CDH1</span>, <span class="html-italic">TGF-β1</span>, <span class="html-italic">VIM,</span> and <span class="html-italic">ZEB1</span> between HaCaT and UM-SCC-22B cells (data correspond to the controls (ctrl) as shown in <b>A</b>–<b>C</b>). Data represent the mean ± SD (<span class="html-italic">n</span> = 3–4), with <span class="html-italic">p</span> &lt; 0.05 considered statistically significant. Statistical differences were indicated as *: <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, n.s.: not significant.</p>
Full article ">Figure 6
<p>The epithelial or mesenchymal phenotype of HNSCC and HaCaT cells depends on the supply of autocrine and exogenous EMT cytokines.</p>
Full article ">
12 pages, 1009 KiB  
Article
The Systemic Inflammation Response Index Efficiently Discriminates between the Failure Patterns of Patients with Isocitrate Dehydrogenase Wild-Type Glioblastoma Following Radiochemotherapy with FLAIR-Based Gross Tumor Volume Delineation
by Sukran Senyurek, Murat Serhat Aygun, Nulifer Kilic Durankus, Eyub Yasar Akdemir, Duygu Sezen, Erkan Topkan, Yasemin Bolukbasi and Ugur Selek
Brain Sci. 2024, 14(9), 922; https://doi.org/10.3390/brainsci14090922 (registering DOI) - 15 Sep 2024
Viewed by 351
Abstract
Background/Objectives: The objective of this study was to assess the connection between the systemic inflammation response index (SIRI) values and failure patterns of patients with IDH wild-type glioblastoma (GB) who underwent radiotherapy (RT) with FLAIR-based gross tumor volume (GTV) delineation. Methods: Seventy-one patients [...] Read more.
Background/Objectives: The objective of this study was to assess the connection between the systemic inflammation response index (SIRI) values and failure patterns of patients with IDH wild-type glioblastoma (GB) who underwent radiotherapy (RT) with FLAIR-based gross tumor volume (GTV) delineation. Methods: Seventy-one patients who received RT at a dose of 60 Gy to the GTV and 50 Gy to the clinical target volume (CTV) and had documented recurrence were retrospectively analyzed. Each patient’s maximum distance of recurrence (MDR) from the GTV was documented in whichever plane it extended the farthest. The failure patterns were described as intra-GTV, in-CTV/out-GTV, distant, and intra-GTV and distant. For analytical purposes, the failure pattern was categorized into two groups, namely Group 1, intra-GTV or in-CTV/out-GTV, and Group 2, distant or intra-GTV and distant. The SIRI was calculated before surgery and corticosteroid administration. A receiver operating characteristic (ROC) curve analysis was used to determine the optimal SIRI cut-off that distinguishes between the different failure patterns. Results: Failure occurred as follows: intra-GTV in 40 (56.3%), in-CTV/out-GTV in 4 (5.6%), distant in 18 (25.4%), and intra-GTV + distant in 9 (12.7%) patients. The mean MDR was 13.5 mm, and recurrent lesions extended beyond 15 mm in only seven patients. Patients with an SIRI score ≥ 3 demonstrated a significantly higher incidence of Group 1 failure patterns than their counterparts with an SIRI score < 3 (74.3% vs. 50.0%; p = 0.035). Conclusions: The present results show that using the SIRI with a cut-off value of ≥3 significantly predicts failure patterns. Additionally, the margin for the GTV can be safely reduced to 15 mm when using FLAIR-based target delineation in patients with GB. Full article
(This article belongs to the Special Issue Brain Tumors: From Molecular Basis to Therapy)
Show Figures

Figure 1

Figure 1
<p>Examples of failure pattern definition. Intra-GTV (<b>a</b>,<b>b</b>), in-CTV/out-GTV (<b>c</b>), and distant (<b>d</b>). Red line: gross tumor volume (prescribed dose: 60 Gy/30 fr), blue line: clinical target volume (prescribed dose: 50 Gy/30 fr), cyan line: recurrent lesion.</p>
Full article ">Figure 2
<p>Receiver operating characteristic curve analyses outcomes. Area under curve: 71.8%; sensitivity: 71.8%; specificity: 70.3%; J-index: 0.421.</p>
Full article ">
17 pages, 13879 KiB  
Article
Sirt2 Regulates Liver Metabolism in a Sex-Specific Manner
by Alexandra V. Schmidt, Sivakama S. Bharathi, Keaton J. Solo, Joanna Bons, Jacob P. Rose, Birgit Schilling and Eric S. Goetzman
Biomolecules 2024, 14(9), 1160; https://doi.org/10.3390/biom14091160 - 15 Sep 2024
Viewed by 347
Abstract
Sirtuin-2 (Sirt2), an NAD+-dependent lysine deacylase enzyme, has previously been implicated as a regulator of glucose metabolism, but the specific mechanisms remain poorly defined. Here, we observed that Sirt2−/− males, but not females, have decreased body fat, moderate hypoglycemia upon fasting, and perturbed [...] Read more.
Sirtuin-2 (Sirt2), an NAD+-dependent lysine deacylase enzyme, has previously been implicated as a regulator of glucose metabolism, but the specific mechanisms remain poorly defined. Here, we observed that Sirt2−/− males, but not females, have decreased body fat, moderate hypoglycemia upon fasting, and perturbed glucose handling during exercise compared to wild type controls. Conversion of injected lactate, pyruvate, and glycerol boluses into glucose via gluconeogenesis was impaired, but only in males. Primary Sirt2−/− male hepatocytes exhibited reduced glycolysis and reduced mitochondrial respiration. RNAseq and proteomics were used to interrogate the mechanisms behind this liver phenotype. Loss of Sirt2 did not lead to transcriptional dysregulation, as very few genes were altered in the transcriptome. In keeping with this, there were also negligible changes to protein abundance. Site-specific quantification of the hepatic acetylome, however, showed that 13% of all detected acetylated peptides were significantly increased in Sirt2−/− male liver versus wild type, representing putative Sirt2 target sites. Strikingly, none of these putative target sites were hyperacetylated in Sirt2−/− female liver. The target sites in the male liver were distributed across mitochondria (44%), cytoplasm (32%), nucleus (8%), and other compartments (16%). Despite the high number of putative mitochondrial Sirt2 targets, Sirt2 antigen was not detected in purified wild type liver mitochondria, suggesting that Sirt2’s regulation of mitochondrial function occurs from outside the organelle. We conclude that Sirt2 regulates hepatic protein acetylation and metabolism in a sex-specific manner. Full article
(This article belongs to the Special Issue Molecular Mechanisms Underlying Liver Diseases)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Mass spectrometry was used to quantify acetylated peptides from male <span class="html-italic">Sirt2+/+</span> (WT) and <span class="html-italic">Sirt2−/−</span> (KO) livers. Peptides with significant changes in acetylation are rendered in red. Blue lines indicate the mean log2 fold-change. A total of 2452 acetylated peptides were quantified. The mean acetylation level significantly increased in 317 peptides (<span class="html-italic">p</span> &lt; 0.01, FC &gt; 1.5) due to Sirt2 ablation. (<b>b</b>) Mass spectrometry was used to quantify acetylated peptides from female <span class="html-italic">Sirt2+/+</span> and <span class="html-italic">Sirt2−/−</span> livers. There was no appreciable change in acetylation levels across putative Sirt2-targets. (<b>c</b>) Volcano plot of <span class="html-italic">Sirt2+/+</span> female to male ratios at each of the 2452 putative Sirt2 targets, expressed as Log2 FC. (<b>d</b>) Volcano plot of <span class="html-italic">Sirt2−/−</span> female to male ratios at each of the 2452 putative Sirt2 targets, expressed as Log2 FC. (<b>e</b>) The mean Log2 FC in <span class="html-italic">Sirt2+/+</span> female/male acetylation was 4.2; the mean Log2 FC in <span class="html-italic">Sirt2−/−</span> female/male acetylation was 1.3. Quantification and comparison of the Log2 FC ratios using a Student’s two-sided <span class="html-italic">t</span>-test showed that <span class="html-italic">Sirt2−/−</span> males had significant decrease in hyperacetylation in 2452 putative sites compared to wild type counterparts (<span class="html-italic">p</span> &lt; 0.0001). (<b>f</b>) Immunoblotting for Sirt2 in livers from wild type mice indicated no sex difference in the expression of the Sirt2 protein. Note that there are multiple bands due to multiple isoforms of Sirt2. **** <span class="html-italic">p</span> &lt; 0.0001 by Student’s two-sided <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>(<b>a</b>) The 306 putative Sirt2 target sites mapped to proteins spread across multiple cellular compartments but were most abundant in the mitochondria and cytosol. (<b>b</b>) A plot of the Log2 ratio (FC &gt; 1.5) of <span class="html-italic">Sirt2−/−</span> vs. <span class="html-italic">Sirt2+/+</span> by cellular compartment revealed that the mitochondria and cytosol had the largest FCs in lysine acetylation. (<b>c</b>) The Log2 Ratio of putative mitochondrial Sirt2 targets from this study was compared to known Sirt3 targets within the mitochondria. Of the 140 Sirt2 targets represented in this plot, 19% are also Sirt3 targets. (<b>d</b>) Preparations of <span class="html-italic">Sirt2+/+</span> nuclear and cytosolic fractions show that Sirt2 is present in both compartments, confirmed with the cytosolic marker α-tubulin and the nuclear marker, Lamin A/C (1 and 2 denote matched, replicate samples). Sirt2 is present in both compartments, confirming acetylomics data. (<b>e</b>) Sirt2 was not detected in mitochondria prepared from <span class="html-italic">Sirt2+/+</span> hepatic tissue lysates (N = 3). 1–3 denotes replicate samples. (<b>f</b>) Sirt2 was not detected in peroxisomes prepared from <span class="html-italic">Sirt2+/+</span> hepatic tissue lysates (N = 3). 1–3 denotes replicate samples. (<b>g</b>) The ultrapure subcellular fractionation technique used completely separated mitochondria and peroxisomes, confirmed with the peroxisomal marker, ACOX1, and the mitochondrial marker, TIMM23 (N = 3). 1–3 denotes matched, replicate samples. Membranes exposed at the same time, cropped for clarity.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Total body weight and lean body mass were the same between <span class="html-italic">Sirt2−/−</span> (N = 11) and <span class="html-italic">Sirt2+/+</span> (N = 8) males, but <span class="html-italic">Sirt2−/−</span> males had significantly lower fat mass compared to <span class="html-italic">Sirt2+/+</span> males (<span class="html-italic">p</span> &lt; 0.05). (<b>b</b>) Total body weight, fat mass, and lean mass were not significantly different between <span class="html-italic">Sirt2−/−</span> (N = 6) and <span class="html-italic">Sirt2+/+</span> females (N = 8). Body composition was calculated for each mouse by dividing fat mass by lean mass. Only <span class="html-italic">Sirt2−/−</span> males had a significantly reduced fat/lean mass body composition (<span class="html-italic">p</span> &lt; 0.001) compared to controls; female <span class="html-italic">Sirt2−/−</span> did not. (<b>c</b>) Blood glucose levels in <span class="html-italic">Sirt2−/−</span> vs. <span class="html-italic">Sirt2+/+</span> males or females in both a fed and fasted state. In a fed state, there was no difference in blood glucose levels between <span class="html-italic">Sirt2−/−</span> and <span class="html-italic">Sirt2+/+</span> in either males (N = 8) or females (N = 8). Upon fasting, only <span class="html-italic">Sirt2−/−</span> males had significantly reduced blood glucose levels (N = 8; <span class="html-italic">p</span> &lt; 0.05) while <span class="html-italic">Sirt2−/−</span> females showed no significant reduction compared to controls (N = 8). (<b>d</b>) H&amp;E staining of <span class="html-italic">Sirt2−/−</span> vs. <span class="html-italic">Sirt2+/+</span> livers indicated no obvious morphological changes or tissue damage (N = 4). (<b>e</b>) Male <span class="html-italic">Sirt2−/−</span> mice (N = 3) had significantly less triglycerides in hepatic tissue compared to <span class="html-italic">Sirt2+/+</span> males (N = 3; <span class="html-italic">p</span> &lt; 0.05). Comparisons made with a Student’s nonparametric <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>(<b>a</b>) <span class="html-italic">Sirt2−/−</span> and <span class="html-italic">Sirt2+/+</span> males ran the same distance (N = 5) before reaching exhaustion. (<b>b</b>) Blood glucose levels were normal prior to the treadmill assay in male <span class="html-italic">Sirt2−/−</span> and <span class="html-italic">Sirt2+/+</span> mice. Upon exhaustion, male <span class="html-italic">Sirt2−/−</span> mice (N = 5) had significantly lower blood glucose levels than <span class="html-italic">Sirt2+/+</span> counterparts (N = 5; <span class="html-italic">p</span> &lt; 0.01). (<b>c</b>) <span class="html-italic">Sirt2−/−</span> females (N = 6) ran significantly further than <span class="html-italic">Sirt2+/+</span> females (N = 7; <span class="html-italic">p</span> &lt; 0.005) before reaching exhaustion. (<b>d</b>) There was no significant difference in blood glucose levels between <span class="html-italic">Sirt2−/−</span> (N = 6) and <span class="html-italic">Sirt2+/+</span> females (N = 7) post-exhaustion. Comparisons were made using a Student’s nonparametric <span class="html-italic">t</span>-test; ns = no difference; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>(<b>a</b>) Male <span class="html-italic">Sirt2−/−</span> mice had a blunted response to a bolus lactate injection, with significantly decreased glucose production 15-min post-injection (N = 5; <span class="html-italic">p</span> &lt; 0.05). (<b>b</b>) <span class="html-italic">Sirt2−/−</span> females had a non-significantly different response to bolus lactate injection compared to <span class="html-italic">Sirt2+/+</span> females (N = 5). (<b>c</b>) Ldha activity in <span class="html-italic">Sirt2−/−</span> fasted males (N = 3) was significantly reduced compared to <span class="html-italic">Sirt2+/+</span> male mice (N = 3; <span class="html-italic">p</span> &lt; 0.05). Ldha activity in fed <span class="html-italic">Sirt2−/−</span> mice was significantly higher than in <span class="html-italic">Sirt2+/+</span> male mice (N = 3; <span class="html-italic">p</span> &lt; 0.05). (<b>d</b>) Male <span class="html-italic">Sirt2−/−</span> mice (N = 5) had a significantly reduced response to a pyruvate tolerance test than male <span class="html-italic">Sirt2+/+</span> mice (N=) 75 (<span class="html-italic">p</span> &lt; 0.0005), 85 (<span class="html-italic">p</span> &lt; 0.005), and 100 min (<span class="html-italic">p</span> &lt; 0.05) post-injection. (<b>e</b>) <span class="html-italic">Sirt2−/−</span> males (N = 5) had significantly less glucose production at 30-, 60-, and 120-min post-glycerol injection (<span class="html-italic">p</span> &lt; 0.05). All comparisons were made at each time point using a Student’s parametric <span class="html-italic">t</span>-test; * <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.</p>
Full article ">Figure 6
<p>Schematic of glycolytic enzymes (red) and gluconeogenic enzymes (purple) that contain putative Sirt2 targets. Each significantly altered residue (<span class="html-italic">p</span> &lt; 0.01; FC &gt; 1.5) is identified with FC reported. Non-targets are in blue.</p>
Full article ">Figure 7
<p>(<b>a</b>) Hepatocytes isolated from a <span class="html-italic">Sirt2−/−</span> liver showed reduced OCR compared to <span class="html-italic">Sirt2+/+</span> hepatocytes during a mitochondrial stress test using a Seahorse Extracellular Flux Analyzer (<span class="html-italic">p</span> &lt; 0.005). (<b>b</b>) Hepatocytes isolated from <span class="html-italic">Sirt2−/−</span> liver showed significantly reduced OCR compared to <span class="html-italic">Sirt2+/+</span> isolated hepatocytes when given media containing only pyruvate as an energy source (<span class="html-italic">p</span> &lt; 0.00001). (<b>c</b>) <span class="html-italic">Sirt2−/−</span> liver tissue lysates (N = 4) had a significantly reduced FAO response to palmitate (16C chain fatty acid) compared to <span class="html-italic">Sirt2+/+</span> liver lysate (N = 4) during a radio-labeled redox assay (<span class="html-italic">p</span> &lt; 0.05). (<b>d</b>) <span class="html-italic">Sirt2-/-</span> liver lysates (N = 6) had a significantly reduced response to octanoate (8C chain fatty acid) compared to liver lysate prepared from <span class="html-italic">Sirt2+/+</span> mice (N = 6; <span class="html-italic">p</span> &lt; 0.05). (<b>e</b>) The maximal mitochondrial capacity was significantly lower in mitochondria isolated from male <span class="html-italic">Sirt2−/−</span> mouse hepatocytes (N = 3) compared to <span class="html-italic">Sirt2+/+</span> isolated mitochondria (N = 3; <span class="html-italic">p</span> &lt; 0.05). (<b>f</b>) The maximal mitochondrial capacity was relatively the same in mitochondria isolated from female <span class="html-italic">Sirt2−/−</span> mouse hepatocytes (N = 3) compared to <span class="html-italic">Sirt2+/+</span> isolated mitochondria (N = 3; <span class="html-italic">p</span> &gt; 0.05). All comparisons were performed with a Student’s nonparametric <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.0001, **** <span class="html-italic">p</span> &lt; 0.00001.</p>
Full article ">
13 pages, 3580 KiB  
Article
Novel Function of Osteocalcin in Chondrocyte Differentiation and Endochondral Ossification Revealed on a CRISPR/Cas9 bglap–bglap2 Deficiency Mouse Model
by Xiang-Fang Yu, Bin Teng, Jun-Feng Li, Jian V. Zhang, Zhe Su and Pei-Gen Ren
Int. J. Mol. Sci. 2024, 25(18), 9945; https://doi.org/10.3390/ijms25189945 (registering DOI) - 15 Sep 2024
Viewed by 264
Abstract
Endochondral ossification is the process by which cartilage is mineralized into bone, and is essential for the development of long bones. Osteocalcin (OCN), a protein abundant in bone matrix, also exhibits high expression in chondrocytes, especially hypertrophic chondrocytes, while its role in endochondral [...] Read more.
Endochondral ossification is the process by which cartilage is mineralized into bone, and is essential for the development of long bones. Osteocalcin (OCN), a protein abundant in bone matrix, also exhibits high expression in chondrocytes, especially hypertrophic chondrocytes, while its role in endochondral ossification remains unclear. Utilizing a new CRISPR/Cas9-mediated bglap–bglap2 deficiency (OCNem) mouse model generated in our laboratory, we provide the first evidence of OCN’s regulatory function in chondrocyte differentiation and endochondral ossification. The OCNem mice exhibited significant delays in primary and secondary ossification centers compared to wild-type mice, along with increased cartilage length in growth plates and hypertrophic zones during neonatal and adolescent stages. These anomalies indicated that OCN deficiency disturbed endochondral ossification during embryonic and postnatal periods. Mechanism wise, OCN deficiency was found to increase chondrocyte differentiation and postpone vascularization process. Furthermore, bone marrow mesenchymal stromal cells (BMSCs) from OCNem mice demonstrated an increased capacity for chondrogenic differentiation. Transcriptional network analysis implicated that BMP and TGF-β signaling pathways were highly affected in OCNem BMSCs, which is closely associated with cartilage development and maintenance. This elucidation of OCN’s function in chondrocyte differentiation and endochondral ossification contributes to a more comprehensive understanding of its impact on skeletal development and homeostasis. Full article
Show Figures

Figure 1

Figure 1
<p>CRISPR/Cas9-mediated <span class="html-italic">bglap–bglap2</span> deficiency (OCN<sup>em</sup>) mice delayed development of early endochondral ossification and shortened primary ossification center (POC) in embryonic period. (<b>A</b>) Gross morphology of WT and OCN<sup>em</sup> mice. (<b>B</b>) Body length. (<b>C</b>–<b>E</b>) Alizarin red S/alcian blue staining on hindlimbs of WT and OCN<sup>em</sup> mice; L: left, R: right. (<b>C</b>) Metacarpals ossification at E17.5; arrows indicated metacarpal POC, scale bars: 500 μm. (<b>D</b>) Metacarpals/phalanges ossification at P0; ovals circled the middle phalangeal POC, scale bars: 500 μm. (<b>E</b>) Femurs and tibias ossification at E15.5, E17.5, and P0; dot lines indicated femoral and tibial POC, scale bars: 2 mm. (<b>F</b>,<b>G</b>) Quantification of the POC length of femur and tibia. There were 5 mice in each group, no samples were excluded and all data were expressed as mean ± SD. The statistical significance was denoted as * <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 by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>The formation of secondary ossification centers (SOCs) was delayed in OCN<sup>em</sup> mice after birth. (<b>A</b>) Representative safranin O/fast green staining of the proximal tibia of WT and OCN<sup>em</sup> mice at P7, P14 and P21 (Scale bars: 100 μm). AC: articular cartilage; gray dashed line outlined the SOC. Quantification of the SOC size (<b>B</b>) and AC thickness (<b>C</b>). There were 5 mice in each group, no samples were excluded and all data were expressed as mean ± SD. The statistical significance was denoted as ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>Knockout of OCN significantly increased the cartilage length. (<b>A</b>,<b>B</b>) Representative safranin O/fast green staining of tibial growth plate (GP) (upper panel) and higher magnification view of hypertrophic zone (HZ) (lower panel) of WT and OCN<sup>em</sup> mice at P0 (<b>A</b>) and P21 (<b>B</b>). At upper panel, black dashed lines indicated growth plates (Scale bars: 100 μm); at lower panel, gray dashed lines outlined HZ (Scale bars: 50 μm). (<b>C</b>,<b>D</b>) Quantification of the overall length of GP, PZ and HZ at P0 (<b>C</b>) and P21 (<b>D</b>). GP: growth plate, PZ: proliferative zone, HZ: hypertrophic zone. There were 5 mice in each group, no samples were excluded and all data were expressed as mean ± SD. The statistical significance was denoted as * <span class="html-italic">p</span> &lt; 0.05 by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Knockout of OCN increased chondrocyte differentiation and postponed vascularization process. (<b>A</b>) IF staining to detect proliferation, differentiation and vascular invasion markers in mice growth plate. Scale bars: 200 μm; (<b>B</b>–<b>E</b>) Quantification of the expression of PCNA, Col II, MMP 13, and VEGF in the growth plate. There were 5 mice in each group, no samples were excluded and all data were expressed as mean ± SD. The statistical significance was denoted as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p>Knockout of OCN decreased the differentiation potential of bone marrow mesenchymal stromal cells (BMSCs) to bone, but more importantly increased the differentiation potential to cartilage. (<b>A</b>) A volcano plot illustrating differentially regulated gene expression between the WT and OCN<sup>em</sup> BMSCs. Values are presented as the log10 of tag counts; (<b>B</b>) The GO functional clustering of differentially expressed genes associated with chondrogenesis and osteogenesis process. * means this biological process was affected; (<b>C</b>,<b>D</b>) GSEA analyzed the genes associated with osteoblast development and chondrocyte development. NES, normalized enrichment score; FDR, false discovery rate. There were 3 samples in each group and no samples were excluded.</p>
Full article ">Figure 6
<p>Chondrogenic differentiation of BMSCs from OCN<sup>em</sup> mice was significantly enhanced. (<b>A</b>) Photomicrographs of WT BMSCs and OCN<sup>em</sup> BMSCs pellets on days 14 and 28 stained with alcian blue/nuclear fast red staining. Scale bars: 50 μm; (<b>B</b>) Quantified the percentage of alcian blue positive area; (<b>C</b>) sGAG content in WT BMSCs and OCN<sup>em</sup> BMSC pellets on day 14 and 28; (<b>D</b>,<b>E</b>) Detection of chondrogenic differentiation–related genes <span class="html-italic">Col2a1</span> expression by RT–qPCR and RNA–sequencing; (<b>F</b>) GSEA analyzed the genes associated with cartilage development; (<b>G</b>) Transcriptional network of the differentially expressed genes related to cartilage development, BMP signaling and TGF–β signaling. The node size was set based on foldchange in the differentially expressed genes. Red-colored circles indicate genes present in 2 or more purple–colored circles. There were more than 3 samples in each group, no samples were excluded and all data were expressed as mean ± SD. The statistical significance was denoted as ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
12 pages, 4519 KiB  
Article
Determination and Analysis of Endogenous Hormones and Cell Wall Composition between the Straight and Twisted Trunk Types of Pinus yunnanensis Franch
by Hailin Li, Rong Xu, Cai Wang, Xiaolin Zhang, Peiling Li, Zhiyang Wu and Dan Zong
Forests 2024, 15(9), 1626; https://doi.org/10.3390/f15091626 (registering DOI) - 14 Sep 2024
Viewed by 260
Abstract
Pinus yunnanensis Franch., one of the pioneer species of wild mountain afforestation in southwest China, plays an essential role in the economy, society and environment of Yunnan Province. Nonetheless, P. yunnanensis’ trunk twisting and bending phenomenon has become more common, which significantly [...] Read more.
Pinus yunnanensis Franch., one of the pioneer species of wild mountain afforestation in southwest China, plays an essential role in the economy, society and environment of Yunnan Province. Nonetheless, P. yunnanensis’ trunk twisting and bending phenomenon has become more common, which significantly restricts its use and economic benefits. In order to clarify the compositional differences between the straight and twisted trunk types of P. yunnanensis and to investigate the reasons for the formation of twisted stems, the present study was carried out to dissect the macroscopic and microscopic structure of the straight and twisted trunk types of P. yunnanensis, to determine the content of cell wall components (lignin, cellulose, hemicellulose), determine the content of endogenous hormones, and the expression validation of phytohormone-related differential genes (GA2OX, COI1, COI2) and cell wall-related genes (XTH16, TCH4). The results showed that the annual rings of twisted trunk types were unevenly distributed, eccentric growth, insignificant decomposition of early and late wood, rounding and widening of the tracheid cells, thickening of the cell wall, and reduction of the cavity diameter; the lignin and hemicellulose contents of twisted trunk types were higher; in twisted trunk types, the contents of gibberellin (GA) and jasmonic acid (JA) increased, and the content of auxin (IAA) was reduced; the GA2OX were significantly down-regulated in twisted trunk types, and the expressions of the genes associated with the cell wall, COI1, COI2, TCH4 and XTH16, were significantly up-regulated. In conclusion, the present study found that the uneven distribution of endogenous hormones may be an important factor leading to the formation of twisted trunk type of P. yunnanensis, which adds new discoveries to reveal the mechanism of the genesis of different trunk types in plants, and provides a theoretical basis for the genetic improvement of forest trees. Full article
(This article belongs to the Section Genetics and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Macrostructural observation of straight and twisted <span class="html-italic">P. yunnanensis</span>. (<b>a</b>): Straight type (S), (<b>b</b>): twisted type (T); the arrows point to the eccentricity of the distribution of annual rings from the fifth year onwards.</p>
Full article ">Figure 2
<p>Microanatomy of straight and twisted trunk types of <span class="html-italic">P. yunnanensis</span>. Straight trunked type (<b>a</b>–<b>c</b>), twisted trunk type (<b>d</b>–<b>f</b>). Comparison of early and latewood demarcation (<b>a</b>,<b>d</b>); comparison of earlywood (<b>b</b>,<b>e</b>); comparison of latewood (<b>c</b>,<b>f</b>); comparison of average tracheid cell width (<b>g</b>), comparison of cavity diameter size (<b>h</b>), comparison of average double-wall thickness (<b>i</b>). S, straight trunk type; T, twisted trunk type. Double asterisk (**) means extremely significant difference (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>Differences in lignin (<b>a</b>), hemicellulose (<b>b</b>) and cellulose (<b>c</b>) contents of straight and twisted trunk types of <span class="html-italic">P. yunnanensis</span>. S, straight trunk type; T, twisted trunk type. Double asterisk (**) means extremely significant difference (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Differences in endogenous hormone content of straight and twisted trunk types of <span class="html-italic">P. yunnanensis</span>. S, straight trunk type; T, twisted trunk type. Asterisk (*) means significant difference (<span class="html-italic">p</span> &lt; 0.05), and double asterisk (**) means extremely significant difference (<span class="html-italic">p</span> &lt; 0.01). GA<sub>7</sub>: Gibberellin A7; GA<sub>4</sub>: Gibberellin A4; GA<sub>3</sub>: Gibberellin A3;GA<sub>1</sub>: Gibberellin A1; ICA: Indole-3-carboxaldehyde; IBA: 3-Indolebutyric acid; IAA: Indole-3-acetic acid; BR: Brassinolide; TZ: trans-Zeatin; IP: N6-Isopentenyladenine; IPR: isopentenyladenine riboside; SA: Salicylic acid; JA-ILE: Jasmonoyl-L-Isoleucine; JA: Jasmonic acid; H2JA: Dihydrojasmonic acid.</p>
Full article ">Figure 5
<p>Differential gene expression of straight and twisted trunk types of <span class="html-italic">P. yunnanensis</span>. S, straight trunk type; T, twisted trunk type. Double asterisk (**) means extremely significant difference (<span class="html-italic">p</span> &lt; 0.01), and three asterisks (***) means extremely significant difference (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
11 pages, 3082 KiB  
Article
The Synergistic Effect of Baloxavir and Neuraminidase Inhibitors against Influenza Viruses In Vitro
by Xiaojia Guo, Lei Zhao, Wei Li, Ruiyuan Cao and Wu Zhong
Viruses 2024, 16(9), 1467; https://doi.org/10.3390/v16091467 - 14 Sep 2024
Viewed by 310
Abstract
Influenza viruses remain a major threat to human health. Four classes of drugs have been approved for the prevention and treatment of influenza infections. Oseltamivir, a neuraminidase inhibitor, is a first-line anti-influenza drug, and baloxavir is part of the newest generation of anti-influenza [...] Read more.
Influenza viruses remain a major threat to human health. Four classes of drugs have been approved for the prevention and treatment of influenza infections. Oseltamivir, a neuraminidase inhibitor, is a first-line anti-influenza drug, and baloxavir is part of the newest generation of anti-influenza drugs that targets the viral polymerase. The emergence of drug resistance has reduced the efficacy of established antiviral drugs. Combination therapy is one of the options for controlling drug resistance and enhancing therapeutical efficacies. Here, we evaluate the antiviral effects of baloxavir combined with neuraminidase inhibitors (NAIs) against wild-type influenza viruses, as well as influenza viruses with drug-resistance mutations. The combination of baloxavir with NAIs led to significant synergistic effects; however, the combination of baloxavir with laninamivir failed to result in a synergistic effect on influenza B viruses. Considering the rapid emergence of drug resistance to baloxavir, we believe that these results will be beneficial for combined drug use against influenza. Full article
(This article belongs to the Special Issue Antiviral Agents to Influenza Virus 2024)
Show Figures

Figure 1

Figure 1
<p>Inhibitory effect of baloxavir in combination with NAIs in cells infected with H1N1-PR8-I38T mutant strain. The antiviral activity of baloxavir + NAIs was evaluated against H1N1-PR8-I38T in MDCK cells. Graphs showing the antiviral activity of baloxavir in combination with (<b>A</b>) oseltamivir acid; (<b>B</b>) zanamivir; (<b>C</b>) laninamivir; and (<b>D</b>) peramivir.</p>
Full article ">Figure 2
<p>Inhibitory effect of baloxavir in combination with NAIs in cells infected with H1N1-WSN-I38T mutant strain. The antiviral activity of baloxavir + NAIs was evaluated against H1N1-WSN-I38T in MDCK cells. Graphs showing the antiviral activity of baloxavir in combination with (<b>A</b>) oseltamivir acid; (<b>B</b>) zanamivir; (<b>C</b>) laninamivir; and (<b>D</b>) peramivir.</p>
Full article ">Figure 3
<p>Inhibitory effect of baloxavir in combination with NAIs in cells infected with H1N1-PR8-R292K mutant strain. The antiviral activity of baloxavir + NAIs was evaluated against H1N1-PR8-R292K in MDCK cells. Graphs showing the antiviral activity of baloxavir in combination with (<b>A</b>) oseltamivir acid; (<b>B</b>) laninamivir; (<b>C</b>) zanamivir; and (<b>D</b>) peramivir.</p>
Full article ">Figure 4
<p>Inhibitory effect of baloxavir in combination with NAIs in cells infected with influenza B strain. The antiviral activity of baloxavir + NAIs was evaluated against the influenza B strain in MDCK cells. Graphs showing the antiviral activity of baloxavir in combination with (<b>A</b>) oseltamivir acid; (<b>B</b>) laninamivir; (<b>C</b>) zanamivir; and (<b>D</b>) peramivir.</p>
Full article ">Figure 5
<p>Inhibitory effect of baloxavir in combination with NAIs in cells infected with influenza H3N2. The antiviral activity of baloxavir + NAIs was evaluated against influenza H3N2 in MDCK cells. Graphs showing the antiviral activity of baloxavir in combination with (<b>A</b>) oseltamivir acid; (<b>B</b>) laninamivir; (<b>C</b>) zanamivir; and (<b>D</b>) peramivir.</p>
Full article ">
19 pages, 2263 KiB  
Article
Improving Beneficial Traits in Bacillus cabrialesii subsp. cabrialesii TE3T through UV-Induced Genomic Changes
by Pamela Helué Morales Sandoval, María Edith Ortega Urquieta, Valeria Valenzuela Ruíz, Kevin Montañez Acosta, Kevin Alejandro Campos Castro, Fannie I. Parra Cota, Gustavo Santoyo and Sergio de los Santos Villalobos
Plants 2024, 13(18), 2578; https://doi.org/10.3390/plants13182578 - 14 Sep 2024
Viewed by 270
Abstract
It is essential to hunt for new technologies that promote sustainable practices for agroecosystems; thus, the bioprospecting of beneficial microorganisms complementing with mutation induction techniques to improve their genomic, metabolic, and functional traits is a promising strategy for the development of sustainable microbial [...] Read more.
It is essential to hunt for new technologies that promote sustainable practices for agroecosystems; thus, the bioprospecting of beneficial microorganisms complementing with mutation induction techniques to improve their genomic, metabolic, and functional traits is a promising strategy for the development of sustainable microbial inoculants. Bacillus cabrialesii subsp. cabrialesii strain TE3T, a previously recognized plant growth-promoting and biological control agent, was subjected to UV mutation induction to improve these agro-biotechnological traits. Dilutions were made which were spread on Petri dishes and placed under a 20 W UV lamp at 10-min intervals for 60 min. After the UV-induced mutation of this strain, 27 bacterial colonies showed morphological differences compared to the wild-type strain; however, only a strain named TE3T-UV25 showed an improvement in 53.6% of the biocontrol against Bipolaris sorokiniana vs. the wild-type strain, by competition of nutrient and space (only detected in the mutant strain), as well as diffusible metabolites. Furthermore, the ability to promote wheat growth was evaluated by carrying out experiments under specific greenhouse conditions, considering un-inoculated, strain TE3T, and strain TE3T-UV25 treatments. Thus, after 120 days, biometric traits in seedlings were quantified and statistical analyses were performed, which showed that strain TE3T-UV25 maintained its ability to promote wheat growth in comparison with the wild-type strain. On the other hand, using bioinformatics tools such as ANI, GGDC, and TYGS, the Overall Genome Relatedness Index (OGRI) and phylogenomic relationship of mutant strain TE3T-UV25 were performed, confirming that it changed its taxonomic affiliation from B. cabrialesii subsp. cabrialesii to Bacillus subtilis. In addition, genome analysis showed that the mutant, wild-type, and B. subtilis strains shared 3654 orthologous genes; however, a higher number of shared genes (3954) was found between the TE3T-UV25 mutant strain and B. subtilis 168, while the mutant strain shared 3703 genes with the wild-type strain. Genome mining was carried out using the AntiSMASH v7.0 web server and showed that mutant and wild-type strains shared six biosynthetic gene clusters associated with biocontrol but additionally, pulcherriminic acid cluster only was detected in the genome of the mutant strain and Rhizocticin A was exclusively detected in the genome of the wild-type strain. Finally, using the PlaBase tool, differences in the number of genes (17) associated with beneficial functions in agroecosystems were detected in the genome of the mutant vs. wild-type strain, such as biofertilization, bioremediation, colonizing plant system, competitive exclusion, phytohormone, plant immune response stimulation, putative functions, stress control, and biocontrol. Thus, the UV-induced mutation was a successful strategy to improve the bioactivity of B. cabrialesii subsp. cabrialesii TE3T related to the agro-biotecnology applications. The obtained mutant strain, B. subtilis TE3T-UV25, is a promising strain to be further studied as an active ingredient for the bioformulation of bacterial inoculants to migrate sustainable agriculture. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Bipolaris sorokiniana</span> TPQ3 growth (<b>a</b>), and its inhibition by <span class="html-italic">Bacillus cabrialesii</span> subsp. <span class="html-italic">cabrialesii</span> TE3<sup>T</sup> (<b>b</b>) and strain TE3<sup>T</sup>-UV25 (<b>c</b>).</p>
Full article ">Figure 2
<p>Mycelial growth of <span class="html-italic">B. sorokiniana</span> TPQ3, measured in mm<sup>2</sup>, in response to cell-free culture filtrate (CF) from the mutant TE3T-UV25 and the wild-type strain TE3<sup>T</sup> grown in LB broth. * and ** Significant difference among studied strains and the control treatment (<span class="html-italic">p</span> ≥ 0.05). Six replicates were performed, and the <span class="html-italic">p</span>-value was 0.0031 according to the Kruskal–Wallis test.</p>
Full article ">Figure 3
<p>Swarming motility of <span class="html-italic">Bacillus cabrialesii</span> subsp. <span class="html-italic">cabrialesii</span> TE3<sup>T</sup> (<b>a</b>) and mutant strain TE3<sup>T</sup>-UV25 (<b>b</b>) on LB agar.</p>
Full article ">Figure 4
<p>Phylogenomic relationship between mutant strain TE3<sup>T</sup>-UV25 and closely related species. Tree inferred with FastME 2.1.6.1 (Lefort et al., 2015 [<a href="#B23-plants-13-02578" class="html-bibr">23</a>]) from GBDP distances calculated from genome sequences. The branch lengths are scaled in terms of the GBDP distance formula <span class="html-italic">d</span>5. The numbers above branches are GBDP pseudo-bootstrap support values &gt; 60% from 100 replications, with an average branch support of 84.3%. Same color between two sequences refers that they belong to the same species and/or subspecies.</p>
Full article ">Figure 5
<p>Venn diagram of the number of shared and unique genes between mutant strain TE3<sup>T</sup>-UV25, wild-type strain TE3<sup>T</sup>, and <span class="html-italic">Bacillus subtilis</span>, through OrthoVenn3.</p>
Full article ">
11 pages, 2312 KiB  
Article
Disulfide Bond Engineering of Soluble ACE2 for Thermal Stability Enhancement
by Yoon Soo Kim, Myeongbin Kim, Hye Min Park, Hyun Jin Kim and Seong Eon Ryu
Int. J. Mol. Sci. 2024, 25(18), 9919; https://doi.org/10.3390/ijms25189919 (registering DOI) - 14 Sep 2024
Viewed by 327
Abstract
Although the primary pandemic of SARS-CoV-2 is over, there are concerns about the resurgence of the next wave of related viruses, including a wide range of variant viruses. The soluble ACE2 (sACE2) inhibits the SARS-CoV-2 spike protein ACE2 interaction and has potential as [...] Read more.
Although the primary pandemic of SARS-CoV-2 is over, there are concerns about the resurgence of the next wave of related viruses, including a wide range of variant viruses. The soluble ACE2 (sACE2) inhibits the SARS-CoV-2 spike protein ACE2 interaction and has potential as a variant-independent therapeutic against SARS-CoV-2. Here, we introduce novel disulfide bonds in the wild-type sACE2-Fc by structure-guided mutagenesis, aiming to improve its stability. The stability of each mutant was assessed by a thermal shift assay to screen mutants with increased thermal stability. As a result, we identified a mutant sACE2-Fc with a significantly increased melting temperature. X-ray crystal structure determination of the sACE2 mutant confirmed the correct formation of the designed disulfide bond, and there were no significant structural disturbances. We also proved that the thermostable sACE2-Fc preserved the spike protein binding affinity comparable to the wild-type sACE2-Fc in both molecular and cellular environments, suggesting its therapeutic potential. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Design of novel disulfide bonds. (<b>a</b>) Selected residue pairs for novel disulfide bonds are presented as balls in magenta. The figure is drawn from the sACE2:SARS-CoV-2 spike (RBD) complex structure (PDB ID: 6VW1) [<a href="#B31-ijms-25-09919" class="html-bibr">31</a>]. The sACE2 and SARS-CoV-2 spike proteins are in green and blue, respectively. The nine disulfide bond pairs in <a href="#ijms-25-09919-t001" class="html-table">Table 1</a> (M1–M9) are indicated on the pairs of the magenta balls. (<b>b</b>) The close-up view of the loop 331–347 region contains the M1 and M2 disulfide bond pairs. The figure is drawn from PDB ID: 1R42 and is in the same view as (<b>a</b>), and loop 331–347 is in magenta. The residues mutated for the M1 (N53C/Q340C) and M2 (I54C/K341C) disulfide bonds are presented as sticks.</p>
Full article ">Figure 2
<p>Purified sACE2-Fc proteins. The Coomassie blue-stained SDS-PAGE gel image is presented for the purified sACE2-Fc proteins. The sACE2-Fc proteins are attached to Fc for expression and purification (see <a href="#sec3-ijms-25-09919" class="html-sec">Section 3</a> Methods). The minor bands in the 200 kDa region are likely the dimer of sACE2-Fc fusion that remains in the sample due to incomplete reduction of the disulfide bond-linked original sACE2-Fc dimer. All bands in the SDS-PAGE gel contain Fc because the samples were purified by the protein A affinity chromatography, and the sizes of the bands indicate that they are sACE2-Fc fusion proteins. The SDS-PAGE gel is shown here to verify the purity of the samples. The spike protein binding activity of the samples was verified in the molecular and cellular assays (see the <a href="#sec2dot4-ijms-25-09919" class="html-sec">Section 2.4</a> and <a href="#sec2dot5-ijms-25-09919" class="html-sec">Section 2.5</a>). Lane 1 represents the wild-type protein, and lanes 2–10 are for the mutant sACE2-Fc proteins (M1–M9) in <a href="#ijms-25-09919-t001" class="html-table">Table 1</a>. Positions for the molecular weight markers are indicated.</p>
Full article ">Figure 3
<p>Melting curves by the thermal shift assay. The first derivative melting curves of the wild-type and mutant sACE2-Fc proteins are presented. The Tm experiments were carried out with sACE2-Fc fusion proteins. Because we observed one melting transition point, the unfoldings of sACE2 and Fc domains are likely correlated. The dotted line represents the –d(RFU)/dT value at 25 °C of the wild type. (<b>a</b>) The melting curves of the wild type (black), the M2 mutant (red), and the other mutants in <a href="#ijms-25-09919-t001" class="html-table">Table 1</a> (M1: yellow, M3: gold, M4: violet, M5: green, M6: grey, M7: brown, M8: blue, and M9: olive) are presented together. We performed three independent experiments, and the Tm results were similar. (<b>b</b>) The melting curves of only the wild type (black) and the M2 mutant (red) are presented for comparison.</p>
Full article ">Figure 4
<p>Structural analysis of the M2 mutant sACE2. (<b>a</b>) The 2mFo-DFc map of the sACE2 mutant. The map is contoured at 0.7σ. In the protein model, the oxygen, nitrogen, and sulfur atoms are in red, blue, and yellow, respectively. The carbon atoms are in green. (<b>b</b>) Structural superposition of the wild-type (PDB ID: 1R42) (gray) and mutant (cyan) sACE2 proteins. The average Cα RMSD value is 0.58 Å. The Cys54–Cys341 disulfide bond is indicated as a yellow circle.</p>
Full article ">Figure 5
<p>Competitive binding assay in cells. The binding affinity of the sACE2-Fc’s in a cellular condition was estimated by a competitive assay. The wild-type (<b>a</b>) and the M2 mutant (<b>b</b>) sACE2-Fcs were added to the spike protein/ACE2 binding assay (see the text). The Y-axis represents the percentage of the spike protein bound to the surface ACE2 on the cell. The X-axis represents different concentrations of sACE2-Fc used in the assay. Bars 1–8 in (<b>a</b>) and (<b>b</b>) represent 0.1, 0.3, 0.8, 2.5, 7.4, 22.2, 66.7, and 200 nM of the wild-type (<b>a</b>) and the mutant sACE2-Fc (<b>b</b>) proteins, respectively. The amount of the spike protein bound to sACE2-Fc decreased in a dose-dependent manner. The assay was triplicated, and the standard deviation is indicated in each bar. In the assay, we prepared the 200 nM protein solution and serial-diluted it three-fold each time. Thus, the lowest concentration was 0.1 nM (bar 1), not 0.0.</p>
Full article ">
17 pages, 9924 KiB  
Article
Osmanthus fragrans Ethylene Response Factor OfERF1-3 Delays Petal Senescence and Is Involved in the Regulation of ABA Signaling
by Gongwei Chen, Fengyuan Chen, Dandan Zhang, Yixiao Zhou, Heng Gu, Yuanzheng Yue, Lianggui Wang and Xiulian Yang
Forests 2024, 15(9), 1619; https://doi.org/10.3390/f15091619 - 14 Sep 2024
Viewed by 282
Abstract
Osmanthus fragrans is widely used in gardening, but the short flowering period of O. fragrans affects its ornamental and economic value. ERF, as a plant ethylene response factor, is an important link in the regulation of plant senescence. In this study, we conducted [...] Read more.
Osmanthus fragrans is widely used in gardening, but the short flowering period of O. fragrans affects its ornamental and economic value. ERF, as a plant ethylene response factor, is an important link in the regulation of plant senescence. In this study, we conducted a comprehensive analysis of the functional role of OfERF1-3 within the petals of O. fragrans. Specifically, the OfERF1-3 gene was cloned and subjected to rigorous sequence analysis. Subsequently, to evaluate its expression patterns and effects, gene overexpression experiments were carried out on both Nicotiana tabacum and O. fragrans. The results showed that OfERF1-3-overexpressing tobacco plants exhibited longer petal opening times compared with those of wild plants. Measurements of physiological parameters also showed that the flowers of overexpressed tobacco plants contained lower levels of malondialdehyde (MDA) and hydrogen peroxide (H2O2) than those of the wild type. There was a lower expression of senescence marker genes in overexpressed tobacco and O. fragrans. A yeast two-hybrid assay showed that OfERF1-3 interacted with OfSKIP14 in a manner related to the regulation of ABA. In summary, OfERF1-3 can play a delaying role in the petal senescence process in O. fragrans, and it interacts with OfSKIP14 to indirectly affect petal senescence by regulating the ABA pathway. Full article
(This article belongs to the Section Genetics and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>The identification of the <span class="html-italic">OfERF1-3</span> gene. (<b>A</b>) Phylogenetic tree of ERF1-3. (<b>B</b>) Alignment of the deduced amino acids <span class="html-italic">OfERF1-3</span>, <span class="html-italic">AtERF1-3</span>, and <span class="html-italic">OeERF1-3</span>. <span class="html-italic">OfERF1-3</span>: <span class="html-italic">Osmanthus fragrans ERF1-3</span>; <span class="html-italic">AtERF1-3</span>: <span class="html-italic">Arabidopsis thaliana ERF1-3</span>; <span class="html-italic">OeERF1-3</span>: <span class="html-italic">Olea europaea</span> var. sylvestris <span class="html-italic">ERF1-3</span>. The amino acids with light blue backgrounds indicate part homology.</p>
Full article ">Figure 2
<p>The expression level of <span class="html-italic">OfERF1-3</span> in petals of <span class="html-italic">O. fragrans</span> at different flowering stages. (<b>A</b>) <span class="html-italic">OfERF1-3</span> expression in the S1–S5 time periods in <span class="html-italic">O. fragrans</span>. FPKM is a unit of gene expression commonly used to measure the relative level of gene expression in the transcriptome. Groups labeled with the same letter indicate <span class="html-italic">p</span> &gt; 0.05, while different letters indicate <span class="html-italic">p</span> &lt; 0.05. Transcriptome data was obtained from the published article: “Analysis of the Aging-Related AP2/ERF Transcription Factor Gene Family in <span class="html-italic">Osmanthus fragrans</span>”. (<b>B</b>) The five flowering periods of <span class="html-italic">O. fragrans</span>: S1: linggeng stage, S2: xiangyan stage, S3: initial flowering stage, S4: full flowering stage, and S5: late flowering stage.</p>
Full article ">Figure 3
<p>Expression of the <span class="html-italic">OfERF1-3</span> in transgenic <span class="html-italic">Nicotiana tabacum</span> petals. (<b>A</b>) The expression of <span class="html-italic">OfERF1-3</span> in the S1–S5 time periods in <span class="html-italic">N. tabacum</span>. Groups labeled with the same letter indicate <span class="html-italic">p</span> &gt; 0.05, while different letters indicate <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) The five flowering periods of tobacco: S1: tight bud stage, S2: mature bud stage, S3: initial flowering stage, S4: full flowering stage, and S5: late flowering stage.</p>
Full article ">Figure 4
<p>Phenotypes of transgenic plants of tobacco with <span class="html-italic">OfERF1-3</span>. WT: wild-type plants; OE: overexpression plants. (<b>A</b>) Comparison of flowering time between wild-type and overexpression plants as a whole. (<b>B</b>) Single flowers from wild-type and overexpression plants from the early flowering stage period to abscission.</p>
Full article ">Figure 5
<p>Changes in the expression of senescence marker genes and physiological indexes in petals of <span class="html-italic">OfERF1-3</span> overexpressing tobacco. (<b>A</b>) Expression of <span class="html-italic">NtSAG12</span> in WT and OE petals. (<b>B</b>) Expression of <span class="html-italic">NtACO1</span> in WT and OE petals. (<b>C</b>) MDA content in WT and OE petals. Groups labeled with the same letter indicate <span class="html-italic">p</span> &gt; 0.05, while different letters indicate <span class="html-italic">p</span> &lt; 0.05. (<b>D</b>) H<sub>2</sub>O<sub>2</sub> content in WT and OE petals. WT: wild-type plants; OE: overexpression plants. ** indicate <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Phenotype analysis of transgenic petals of <span class="html-italic">O. fragrans</span> transformed with <span class="html-italic">OfERF1-3</span>; EV: pSuper1300 empty vector. (<b>A</b>) Phenotypes of transgenic petals of <span class="html-italic">O. fragrans</span> transformed with <span class="html-italic">OfERF1-3</span> over a 48 h period. (<b>B</b>) Comparative analysis of <span class="html-italic">OfERF1-3</span> expression of empty vector and transgenic petals of <span class="html-italic">O. fragrans</span>. (<b>C</b>) Expression of <span class="html-italic">OfSAG21</span> in pSuper1300 empty vector and pSuper1300-<span class="html-italic">OfERF1-3</span> transgenic petals. (<b>D</b>) Expression of <span class="html-italic">OfACS1</span> in pSuper1300 empty vector and pSuper1300-<span class="html-italic">OfERF1-3</span> transgenic petals. * indicate <span class="html-italic">p</span> &lt; 0.05 and *** indicate <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Yeast self-activation assay and two-hybrid sieve library assay results. (<b>A</b>) The results of the yeast self-activation assay showed that <span class="html-italic">OfERF1-3</span> exhibits no self-activating activity. The pGBKT7-Lam + pGADT7-T control vector served as a negative control. The pGBKT7-53 + pGADT7-T control vector served as a positive control. (<b>B</b>) A total of 21 positive yeast monoclones were obtained via yeast two-hybrid library screening.</p>
Full article ">Figure 8
<p>A yeast two-hybrid assay identified the <span class="html-italic">OfERF1-3</span> that interacted with <span class="html-italic">OfSKIP14</span>. The pGBKT7-Lam + pGADT7-T control vector served as a negative control. The pGBKT7-53 + pGADT7-T control vector served as a positive control.</p>
Full article ">
12 pages, 2290 KiB  
Article
Mild Heat Stress Alters the Physical State and Structure of Membranes in Triacylglycerol-Deficient Fission Yeast, Schizosaccharomyces pombe
by Péter Gudmann, Imre Gombos, Mária Péter, Gábor Balogh, Zsolt Török, László Vígh and Attila Glatz
Cells 2024, 13(18), 1543; https://doi.org/10.3390/cells13181543 - 13 Sep 2024
Viewed by 302
Abstract
We investigated whether the elimination of two major enzymes responsible for triacylglycerol synthesis altered the structure and physical state of organelle membranes under mild heat shock conditions in the fission yeast, Schizosaccharomyces pombe. Our study revealed that key intracellular membrane structures, lipid [...] Read more.
We investigated whether the elimination of two major enzymes responsible for triacylglycerol synthesis altered the structure and physical state of organelle membranes under mild heat shock conditions in the fission yeast, Schizosaccharomyces pombe. Our study revealed that key intracellular membrane structures, lipid droplets, vacuoles, the mitochondrial network, and the cortical endoplasmic reticulum were all affected in mutant fission yeast cells under mild heat shock but not under normal growth conditions. We also obtained direct evidence that triacylglycerol-deficient cells were less capable than wild-type cells of adjusting their membrane physical properties during thermal stress. The production of thermoprotective molecules, such as HSP16 and trehalose, was reduced in the mutant strain. These findings suggest that an intact system of triacylglycerol metabolism significantly contributes to membrane protection during heat stress. Full article
(This article belongs to the Special Issue Advances in Biophysics of Cellular Membranes)
Show Figures

Figure 1

Figure 1
<p>Analysis of lipid droplets in WT and DKO cells subjected to HS. (<b>A</b>) LD540 staining of LDs in WT and DKO cells before and after HS. (<b>B</b>) Quantitative analysis of the LDs in the WT and mutant cells. Data are mean ± SEM of <span class="html-italic">n</span> = 3 independent experiments with ≥500 cells/experiment; * <span class="html-italic">p</span> &lt; 0.05, 30 °C vs. 40 °C, 1 h; <span>$</span> <span class="html-italic">p</span> &lt; 0.05 WT vs. DKO.</p>
Full article ">Figure 2
<p>Analysis of vacuoles of WT and DKO exposed to HS. (<b>A</b>) Representative images of MDY-64-stained WT and DKO cells. Red arrows indicate enlarged vacuoles in heat-shocked DKO cells. Analysis of the vacuolar size (<b>B</b>), quantity (<b>C</b>), and surface/volume (<b>D</b>) after HS in both strains. Data are mean ± SEM of <span class="html-italic">n</span> = 3 independent experiments with &gt;200 vacuoles analyzed; * <span class="html-italic">p</span> &lt; 0.05, 30 °C vs. 40 °C 1 h.</p>
Full article ">Figure 3
<p>Changes in the mitochondrial network and cortical ER of heat-stressed WT and DKO <span class="html-italic">S. pombe</span> cells. (<b>A</b>) MitoTracker CMXRos staining; disordered parts of the mitochondrial network are indicated by arrows. (<b>B</b>) ER-thermo-yellow staining; dot-like ER structures are indicated by red arrows.</p>
Full article ">Figure 4
<p>Assessment of lipid packing of membranes in the wild-type and DKO strains. (<b>A</b>) False colored GP images of the WT (upper panel) and DKO (lower panel) cells during heat shock. (<b>B</b>) Median GP values of plasma membranes (PM) and endomembranes (EM) of control and heat-shocked WT and DKO cells. Data are mean ± SEM of <span class="html-italic">n</span> = 4 independent experiments, &gt;90 cells per data point analyzed; * <span class="html-italic">p</span> &lt; 0.05 30 °C vs. 40 °C, 1 h; <sup><span>$</span></sup> <span class="html-italic">p</span> &lt; 0.05 WT vs. DKO; <sup>+</sup> <span class="html-italic">p</span> &lt; 0.05 PM vs. EM. GP, generalized polarization.</p>
Full article ">Figure 5
<p>Comparison of the expression of the thermoprotectants HSP16 and trehalose in heat-stressed WT and DKO cells. (<b>A</b>) Induction of HSP16-GFP in the WT (BRC40) and DKO (BRC62) backgrounds. (<b>B</b>) Accumulation of trehalose in WT and DKO cells during heat treatment. Data are mean ± SD for n = 3 independent experiments; * <span class="html-italic">p</span> &lt; 0.05 at 30 °C vs. 40 °C for 1 h; <sup><span>$</span></sup> <span class="html-italic">p</span> &lt; 0.05 in WT vs. DKO.</p>
Full article ">
23 pages, 2417 KiB  
Article
Effects of Omega-3 Polyunsaturated Fatty Acids on the Formation of Adipokines, Cytokines, and Oxylipins in Retroperitoneal Adi-Pose Tissue of Mice
by Tatjana Wenderoth, Martin Feldotto, Jessica Hernandez, Julia Schäffer, Stephan Leisengang, Fabian Johannes Pflieger, Janne Bredehöft, Konstantin Mayer, Jing X. Kang, Jens Bier, Friedrich Grimminger, Nadine Paßlack and Christoph Rummel
Int. J. Mol. Sci. 2024, 25(18), 9904; https://doi.org/10.3390/ijms25189904 - 13 Sep 2024
Viewed by 392
Abstract
Oxylipins and specialized pro-resolving lipid mediators (SPMs) derived from polyunsaturated fatty acids (PUFAs) are mediators that coordinate an active process of inflammation resolution. While these mediators have potential as circulating biomarkers for several disease states with inflammatory components, the source of plasma oxylipins/SPMs [...] Read more.
Oxylipins and specialized pro-resolving lipid mediators (SPMs) derived from polyunsaturated fatty acids (PUFAs) are mediators that coordinate an active process of inflammation resolution. While these mediators have potential as circulating biomarkers for several disease states with inflammatory components, the source of plasma oxylipins/SPMs remains a matter of debate but may involve white adipose tissue (WAT). Here, we aimed to investigate to what extent high or low omega (n)-3 PUFA enrichment affects the production of cytokines and adipokines (RT-PCR), as well as oxylipins/SPMs (liquid chromatography–tandem mass spectrometry) in the WAT of mice during lipopolysaccharide (LPS)-induced systemic inflammation (intraperitoneal injection, 2.5 mg/kg, 24 h). For this purpose, n-3 PUFA genetically enriched mice (FAT-1), which endogenously synthesize n-3 PUFAs, were compared to wild-type mice (WT) and combined with n-3 PUFA-sufficient or deficient diets. LPS-induced systemic inflammation resulted in the decreased expression of most adipokines and interleukin-6 in WAT, whereas the n-3-sufficient diet increased them compared to the deficient diet. The n-6 PUFA arachidonic acid was decreased in WAT of FAT-1 mice, while n-3 derived PUFAs (eicosapentaenoic acid, docosahexaenoic acid) and their metabolites (oxylipins/SPMs) were increased in WAT by genetic and nutritional n-3 enrichment. Several oxylipins/SPMs were increased by LPS treatment in WAT compared to PBS-treated controls in genetically n-3 enriched FAT-1 mice. Overall, we show that WAT may significantly contribute to circulating oxylipin production. Moreover, n-3-sufficient or n-3-deficient diets alter adipokine production. The precise interplay between cytokines, adipokines, and oxylipins remains to be further investigated. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in mRNA expression of inflammatory marker genes. Expression of inflammatory marker genes in white adipose tissue (WAT) 24 h p.i. with lipopolysaccharide (LPS, i.p., 2.5 mg/kg) or phosphate buffered saline (PBS). TLR4: toll-like receptor 4 (<b>A</b>); IκBα: inhibitor of nuclear factor κB (<b>B</b>); IL-6: interleukin-6 (<b>C</b>). WT: wild-type; FAT-1: transgenic n-3 PUFA-enriched mice; def: n-3 PUFA-deficient diet; suf: n-3 PUFA-sufficient diet. Analyzed by two-way ANOVA (treatment, genotype), separately for def and suf; represent mean ± SEM with symbols indicating individual mice (<span class="html-italic">n</span> = 5–6). #: Main effect of treatment (LPS vs. PBS) separately analyzed for def or suf. §: Effect of diet analyzed by <span class="html-italic">t</span>-test comparing suf with def diet regardless of their genotype or treatment. §§§: <span class="html-italic">p</span> &lt; 0.001, ####: <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Changes in adipokine mRNA expression in WAT. Expression of white adipose tissue (WAT) adipokine levels 24 h p.i. with lipopolysaccharide (LPS, i.p., 2.5 mg/kg) or phosphate buffered saline (PBS). Leptin (<b>A</b>); adiponectin (<b>B</b>); CTRP3: C1q/TNF-Related Protein 3 (<b>C</b>). WT: wild-type; FAT-1: transgenic n-3 PUFA-enriched mice; def: n-3 PUFA-deficient diet; suf: n-3 PUFA-sufficient diet. Analyzed by two-way ANOVA (treatment, genotype), separately for def and suf; bars represent mean ± SEM with symbols indicating individual mice (<span class="html-italic">n</span> = 5–6). #: Main effect of treatment (LPS vs. PBS) separately analyzed for def or suf. §: Effect of diet analyzed by <span class="html-italic">t</span>-test comparing suf with def diet regardless of their genotype or treatment. ##: <span class="html-italic">p</span> &lt; 0.001, §§§: <span class="html-italic">p</span> &lt; 0.001, ####/§§§§: <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>The n-6:n-3 PUFA ratio and n-6 lipid mediators in WAT. Detection and quantification of arachidonic acid (AA) and its metabolites by liquid chromatography–tandem mass spectrometry in white adipose tissue (WAT) 24 h p.i. with lipopolysaccharide (LPS, i.p., 2.5 mg/kg) or phosphate-buffered saline (PBS). n-6 (AA):n-3 (EPA+DHA) ratio (<b>A</b>); AA (<b>B</b>); leukotriene B4 (LTB4) (<b>C</b>). n-6: Omega-6; n-3: Omega-3; WT: wild-type; FAT-1: transgenic n-3 PUFA-enriched mice; def: n-3 PUFA-deficient diet; suf: n-3 PUFA-sufficient diet. Analyzed by two-way ANOVA (treatment, genotype), separately for def and suf; Šídák post hoc tests were conducted if a significant interaction was detected; interaction n-6:n-3 ratio deficient group <span class="html-italic">p</span> &lt; 0.0001. Bars represent mean ± SEM with symbols indicating individual mice (<span class="html-italic">n</span> = 5–6). n.d.: not detected. *: Post hoc effects of treatment and genotype. +: Main effect of genotype (FAT-1 vs. WT) separately analyzed for def or suf. #: Main effect of treatment (LPS vs. PBS) separately analyzed for def or suf. §: Effect of diet analyzed by <span class="html-italic">t</span>-test comparing suf with def diet regardless of their genotype or treatment. §: <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.</p>
Full article ">Figure 4
<p>The n-3 lipid mediator EPA and its metabolites in WAT. Detection and quantification of n-3 lipid mediators and metabolites from eicosapentaenoic acid (EPA) by liquid chromatography–tandem mass spectrometry in white adipose tissue (WAT) of mice 24 h after in vivo stimulation with lipopolysaccharide (LPS, i.p., 2.5 mg/kg) or phosphate buffered saline (PBS). EPA (<b>A</b>) and its metabolites 18-hydroxyeicosapentaenoic acid (18-HEPE) (<b>B</b>), resolvin E1 (RvE1) (<b>C</b>), and 18-oxo RvE1 (<b>D</b>). WT: wild-type; FAT-1: transgenic n-3 PUFA-enriched mice; def: n-3 PUFA-deficient diet; suf: n-3 PUFA-sufficient diet. Analyzed by two-way ANOVA (treatment, genotype), separately for def and suf; Šídák post hoc tests were conducted if a significant interaction was detected; interaction EPA def group <span class="html-italic">p</span> &lt; 0.001. Bars represent mean ± SEM with symbols indicating individual mice (<span class="html-italic">n</span> = 5–6). *: Post hoc effects of treatment and genotype. +: Main effect of genotype (FAT-1 vs. WT) separately analyzed for def or suf. #: Main effect of treatment (LPS vs. PBS) separately analyzed for def or suf. §: Effect of diet; analyzed by <span class="html-italic">t</span>-test comparing suf with def diet, regardless of their genotype or treatment. #/§: <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.</p>
Full article ">Figure 5
<p>The n-3 lipid mediator DHA and its metabolites in WAT. Detection and quantification of n-3 lipid mediators and metabolites of docosahexaenoic acid (DHA) by liquid chromatography–tandem mass spectrometry. Concentrations in white adipose tissue (WAT) 24 h p.i. with lipopolysaccharide (LPS, i.p., 2.5 mg/kg.) or phosphate-buffered saline (PBS). DHA (<b>A</b>) and its metabolites 14(S)-hydroxy-DHA (14(S)-HDHA) (<b>B</b>), 17(S)-hydroxy-DHA (17(S)-HDHA) (<b>C</b>), resolvin (Rv) D1 (<b>D</b>), and RvD2 (<b>E</b>). WT: wild-type; FAT-1: transgenic n-3 PUFA-enriched mice; def: n-3 PUFA-deficient diet; suf: n-3 PUFA-sufficient diet. Analyzed by two-way ANOVA (treatment, genotype), separately for def and suf; Šídák post hoc tests were conducted if a significant interaction was detected; interaction: DHA def group: <span class="html-italic">p</span> &lt; 0.001, 14(S)-HDHA def group: <span class="html-italic">p</span> &lt; 0.01, 17(S)-HDHA def group: <span class="html-italic">p</span> &lt; 0.05, RvD2 suf group: <span class="html-italic">p</span> &lt; 0.05. Bars represent mean ± SEM with symbols indicating individual mice (<span class="html-italic">n</span> = 5–6). *: Post hoc effects of treatment and genotype. +: Main effect of genotype (FAT-1 vs. WT) separately analyzed for def or suf. #: Main effect of treatment (LPS vs. PBS) separately analyzed for def or suf. §: Effect of diet analyzed by <span class="html-italic">t</span>-test comparing suf with def diet regardless of their genotype or treatment. */#/+: <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.</p>
Full article ">
19 pages, 7244 KiB  
Article
Reduction of Mitochondrial Calcium Overload via MKT077-Induced Inhibition of Glucose-Regulated Protein 75 Alleviates Skeletal Muscle Pathology in Dystrophin-Deficient mdx Mice
by Mikhail V. Dubinin, Anastasia E. Stepanova, Irina B. Mikheeva, Anastasia D. Igoshkina, Alena A. Cherepanova, Eugeny Yu. Talanov, Ekaterina I. Khoroshavina and Konstantin N. Belosludtsev
Int. J. Mol. Sci. 2024, 25(18), 9892; https://doi.org/10.3390/ijms25189892 - 13 Sep 2024
Viewed by 220
Abstract
Duchenne muscular dystrophy is secondarily accompanied by Ca2+ excess in muscle fibers. Part of the Ca2+ accumulates in the mitochondria, contributing to the development of mitochondrial dysfunction and degeneration of muscles. In this work, we assessed the effect of intraperitoneal administration [...] Read more.
Duchenne muscular dystrophy is secondarily accompanied by Ca2+ excess in muscle fibers. Part of the Ca2+ accumulates in the mitochondria, contributing to the development of mitochondrial dysfunction and degeneration of muscles. In this work, we assessed the effect of intraperitoneal administration of rhodacyanine MKT077 (5 mg/kg/day), which is able to suppress glucose-regulated protein 75 (GRP75)-mediated Ca2+ transfer from the sarcoplasmic reticulum (SR) to mitochondria, on the Ca2+ overload of skeletal muscle mitochondria in dystrophin-deficient mdx mice and the concomitant mitochondrial dysfunction contributing to muscle pathology. MKT077 prevented Ca2+ overload of quadriceps mitochondria in mdx mice, reduced the intensity of oxidative stress, and improved mitochondrial ultrastructure, but had no effect on impaired oxidative phosphorylation. MKT077 eliminated quadriceps calcification and reduced the intensity of muscle fiber degeneration, fibrosis level, and normalized grip strength in mdx mice. However, we noted a negative effect of MKT077 on wild-type mice, expressed as a decrease in the efficiency of mitochondrial oxidative phosphorylation, SR stress development, ultrastructural disturbances in the quadriceps, and a reduction in animal endurance in the wire-hanging test. This paper discusses the impact of MKT077 modulation of mitochondrial dysfunction on the development of skeletal muscle pathology in mdx mice. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of MKT077 on mitochondrial calcium homeostasis. (<b>A</b>) Mitochondrial Ca<sup>2+</sup> load assay; 0.1 mg/mL alamethicin (ALM) was added to induce maximal calcium release from the mitochondrial matrix. The results are presented as means ± SEM (<span class="html-italic">n</span> = 4). (<b>B</b>) Changes in the external [Ca<sup>2+</sup>] upon the successive addition of 20 μM Ca<sup>2+</sup> pulses to the suspension of the skeletal muscle mitochondria of the experimental animals. (<b>C</b>) Calcium retention capacity of the skeletal muscle mitochondria. The results are presented as means ± SEM.</p>
Full article ">Figure 2
<p>Western blotting of GRP75 and GAPDH (<b>A</b>), quantification of GRP75/GAPDH ratio (<b>B</b>) and mRNA expression of <span class="html-italic">Hspa9</span> relative to <span class="html-italic">Rplp2</span> (<b>C</b>) in the skeletal muscles of experimental groups of mice. The results are presented as means ± SEM.</p>
Full article ">Figure 3
<p>Effect of MKT077 on lipid peroxidation in mitochondria. Lipid peroxidation was assessed by the level of TBARS in the skeletal muscle mitochondria of the experimental animals. The results are presented as means ± SEM.</p>
Full article ">Figure 4
<p>The effect of MKT077 on ATP levels in the quadriceps of mice. The data are presented as means ± SEM.</p>
Full article ">Figure 5
<p>Representative transmission electron micrographs of mouse quadriceps sections. Mitochondria of the subsarcolemmal population are highlighted with red arrows. The asterisk marks the same mitochondria at low (<b>left side</b>) and high (<b>right side</b>) magnifications. Scale bar is 1 μm.</p>
Full article ">Figure 6
<p>Electron micrograph (<a href="#ijms-25-09892-f005" class="html-fig">Figure 5</a>) profiles: percentage of MAM surface area per mitochondrion perimeter in each microscopic field (<b>A</b>); mitochondrial perimeter (<b>B</b>) and number of mitochondria per plate (<b>C</b>). The data are presented as means ± SEM (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>The relative mRNA expression of <span class="html-italic">Pink1</span> (<b>A</b>) and <span class="html-italic">Parkin</span> (<b>B</b>) relative to <span class="html-italic">Rplp2</span> in the skeletal muscles of experimental groups of mice. The data are presented as means ± SEM (<span class="html-italic">n</span> = 5).</p>
Full article ">Figure 8
<p>Western blotting of GRP78 and GAPDH (<b>A</b>), quantification of GRP78/GAPDH ratio (<b>B</b>) and mRNA expression of <span class="html-italic">Hspa5</span> relative to <span class="html-italic">Rplp2</span> (<b>C</b>) in the skeletal muscles of experimental groups of mice. The results are presented as means ± SEM.</p>
Full article ">Figure 9
<p>Representative histological images of the quadricep muscles of experimental mice: H&amp;E staining, Alizarin red (AR) staining (calcified area indicated by arrows), and Sirius red (SR) staining (fibrotic area indicated by arrows). Scale bar is 75 μm (H&amp;E and SR) or 150 μm (AR).</p>
Full article ">Figure 10
<p>The percentage of CNF (<b>A</b>), mean minimal Feret’s diameter (<b>B</b>), fiber size distribution (<b>C</b>), percentage of the total fiber number), Alizarin red staining area (<b>D</b>), and the percentage of fibrosis (<b>E</b>) in the quadriceps of experimental animals. The data are presented as means ± SEM (<span class="html-italic">n</span> = 5).</p>
Full article ">Figure 11
<p>The effect of MKT077 on the activity of creatine kinase in the blood serum of mice. The data are presented as the mean ± SEM.</p>
Full article ">Figure 12
<p>The effect of MKT077 on muscle strength and endurance in mice. (<b>A</b>) Grip strength test. (<b>B</b>) Wire-hanging test. The data are presented as means ± SEM.</p>
Full article ">
17 pages, 1828 KiB  
Article
Whole-Genome and Poly(A)+Transcriptome Analysis of the Drosophila Mutant agnts3 with Cognitive Dysfunctions
by Aleksandr V. Zhuravlev, Dmitrii E. Polev, Anna V. Medvedeva and Elena V. Savvateeva-Popova
Int. J. Mol. Sci. 2024, 25(18), 9891; https://doi.org/10.3390/ijms25189891 - 13 Sep 2024
Viewed by 232
Abstract
The temperature-sensitive Drosophila mutant agnts3 exhibits the restoration of learning defects both after heat shock (HS) and under hypomagnetic conditions (HMC). Previously, agnts3 was shown to have an increased level of LIM kinase 1 (LIMK1). However, its limk1 sequence did not [...] Read more.
The temperature-sensitive Drosophila mutant agnts3 exhibits the restoration of learning defects both after heat shock (HS) and under hypomagnetic conditions (HMC). Previously, agnts3 was shown to have an increased level of LIM kinase 1 (LIMK1). However, its limk1 sequence did not significantly differ from that of the wild-type strain Canton-S (CS). Here, we performed whole-genome and poly(A)-enriched transcriptome sequencing of CS and agnts3 males normally, after HMC, and after HS. Several high-effect agnts3-specific mutations were identified, including MED23 (regulation of HS-dependent transcription) and Spn42De, the human orthologs of which are associated with intellectual disorders. Pronounced interstrain differences between the transcription profiles were revealed. Mainly, they included the genes of defense and stress response, long non-coding RNAs, and transposons. After HS, the differences between the transcriptomes became less pronounced. In agnts3, prosalpha1 was the only gene whose expression changed after both HS and HMC. The normal downregulation of prosalpha1 and Spn42De in agnts3 was confirmed by RT-PCR. Analysis of limk1 expression did not reveal any interstrain differences or changes after stress. Thus, behavioral differences between CS and agnts3 both under normal and stressed conditions are not due to differences in limk1 transcription. Instead, MED23, Spn42De, and prosalpha1 are more likely to contribute to the agnts3 phenotype. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Differentially expressed genes in <span class="html-italic">agn<sup>ts3</sup></span> relative to <span class="html-italic">Canton-S</span> (<span class="html-italic">CS</span>) under different conditions (Venn diagrams). The names of five genes showing the maximum change in expression under specific conditions are shown. <span>$</span> refers to genes that showed the interstrain difference of expression across all conditions.</p>
Full article ">Figure 2
<p>GO Biological Processes enrichment. Abbreviations: Down, downregulated; FDR, false discovery rate; HMC, hypomagnetic conditions; HS, heat shock; N, normal conditions; NR, negative regulation; Up, upregulated. Because there is significant overlap between the genes from the different processes, only five processes are shown for each case.</p>
Full article ">Figure 3
<p>Normalized gene expression levels in <span class="html-italic">CS</span> and <span class="html-italic">agn<sup>ts3</sup></span> (RT-PCR data). Expression levels are indicated in conditional units (C.U.). The average expression level of each gene was set to 1. Statistical differences: <span>$</span> from <span class="html-italic">CS</span> (two-sided t-тест, n = 5, p is given above the box plot). <span class="html-italic">Cry</span> #1 and <span class="html-italic">Cry</span> #2 correspond to different pairs of primers. The median is shown as a dark line. Outliers are indicated by a diamond.</p>
Full article ">
10 pages, 1746 KiB  
Article
Association of Wild-Type TP53 with Downregulation of Lovastatin Sensitivity in Human Non-Small Cell Lung Cancer Cells
by Yu-Yao Chang, Tsung-Ying Yang and Gwo-Tarng Sheu
Curr. Issues Mol. Biol. 2024, 46(9), 10130-10139; https://doi.org/10.3390/cimb46090604 - 13 Sep 2024
Viewed by 222
Abstract
Statins inhibit 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR), the rate-limiting enzyme of the mevalonate pathway, and reduce cholesterol synthesis. They also have been demonstrated to improve prognosis in patients with various cancers, suggesting a potential anti-cancer effect of statins. However, there is no consensus on the [...] Read more.
Statins inhibit 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR), the rate-limiting enzyme of the mevalonate pathway, and reduce cholesterol synthesis. They also have been demonstrated to improve prognosis in patients with various cancers, suggesting a potential anti-cancer effect of statins. However, there is no consensus on the molecular targets of statins for their anti-cancer effects. Docetaxel (DOC) is a microtubule-stabilizing agent currently used as a chemotherapeutic drug in several cancers, including lung cancer. Interestingly, the anti-cancer effects of either drug that are related to abnormal or wild-type TP53 gene have been implied. Therefore, the drug sensitivity of DOC and lovastatin in human lung cancer cells was evaluated. We found that H1355 (mutant TP53-E285K), CL1 (mutant TP53-R248W), and H1299 (TP53-null) human non-small cell lung cancer cells were more sensitive to lovastatin than A549 and H460 cells expressing wild-type TP53. Conversely, A549 and H460 cells showed higher sensitivity to DOC than H1299 and CL1 cells, as demonstrated by the MTT assay. When endogenous TP53 activity was inhibited by pifithrin-α in A549 and H460 cells, lovastatin sensitivities significantly increased, and cancer cell viabilities markedly reduced. These results indicate that TP53 status is associated with the anti-cancer effect of statins in human lung cancer cells. Mutated or null TP53 status is correlated with higher statin sensitivity. Furthermore, DOC-resistant H1299 (H1299/D8) cells showed significant sensitivity to lovastatin treatment compared to DOC-resistant A549 (A549/D16) cells, indicating a potential application of statins/chemotherapy combination therapy to control wild-type and abnormal TP53-containing human lung tumors. Full article
Show Figures

Figure 1

Figure 1
<p>Determination of the DOC and lovastatin sensitivities in lung cancer cells by MTT assay. To evaluate the sensitivity of DOC and lovastatin in lung cancer cells with different TP53 statuses, varied concentrations of DOC (<b>a</b>) (2.5 to 40 nM) and (<b>d</b>) lovastatin (2.5 to 40 μM) were applied on A549, H460, H1299, H1355, CL1-0, and CL1-5 cells. (<b>b</b>) The half-maximal inhibitory concentration (IC50) of each cell line to DOC and (<b>e</b>) lovastatin were determined accordingly. IC50 for each cell line to (<b>c</b>) Either DOC or (<b>f</b>) lovastatin sensitivity were determined and statistically analyzed with the TP53 status, respectively. * A value of <span class="html-italic">p</span> &lt; 0.05 was considered to be statistically significant.</p>
Full article ">Figure 2
<p>Inhibition of wt-TP53 protein by PFTα followed by DOC or lovastatin to measure drug sensitivity of lung cancer cells. (<b>a</b>) H460 cells were pretreated with DMSO or PFTα (10 μM) for 48 h, then treated either with DOC (20 nM) or lovastatin (10 μM) for another 48 h. Similar treatments were applied to (<b>c</b>) A549 cells, and the images of the cells were taken with 100× magnification using a light microscope. The viability of (<b>b</b>) H460 cells and (<b>d</b>) A549 cells were calculated, followed by an MTT assay. * Statistically significant.</p>
Full article ">Figure 3
<p>Lovastatin sensitivities of DOC-resistant A549 and H1299 sublines. (<b>a</b>) A549 cells were pretreated either with varied concentrations of lovastatin alone for 30 min or in combination with DOC (16 nM) for an additional 48 h. Similar treatments were applied to A549/D16 cells followed by MTT assay (<b>b</b>) H1299 cells and H299/D8 cells were treated either with varied concentrations of lovastatin alone or combined with DOC (8 nM) for 48 h followed by MTT assay.</p>
Full article ">Figure 4
<p>A simplified conclusion has been obtained and summarized.</p>
Full article ">
Back to TopTop