[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

Article Types

Countries / Regions

Search Results (425)

Search Parameters:
Keywords = dual-reporter genes

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 5477 KiB  
Article
Transcriptome Analysis Identified PyNAC42 as a Positive Regulator of Anthocyanin Biosynthesis Induced by Nitrogen Deficiency in Pear (Pyrus spp.)
by Jianhui Zhang, Bobo Song, Guosong Chen, Guangyan Yang, Meiling Ming, Shiqiang Zhang, Zhaolong Xue, Chenhui Han, Jiaming Li and Jun Wu
Horticulturae 2024, 10(9), 980; https://doi.org/10.3390/horticulturae10090980 (registering DOI) - 16 Sep 2024
Viewed by 182
Abstract
Anthocyanins are important secondary metabolites in plants, which contribute to fruit color and nutritional value. Anthocyanins can be regulated by environmental factors such as light, low temperature, water conditions, and nutrition limitations. Nitrogen (N) is an essential macroelement for plant development, its deficiency [...] Read more.
Anthocyanins are important secondary metabolites in plants, which contribute to fruit color and nutritional value. Anthocyanins can be regulated by environmental factors such as light, low temperature, water conditions, and nutrition limitations. Nitrogen (N) is an essential macroelement for plant development, its deficiency as a kind of nutrition limitation often induces anthocyanin accumulation in many plants. However, there is a lack of reports regarding the effect of nitrogen deficiency on anthocyanin biosynthesis in pears. In this study, we found that N deficiency resulted in anthocyanin accumulation in pear callus and upregulated the expression of anthocyanin biosynthesis pathway structural genes (PyPAL, PyCHS, PyCHI, PyF3H, PyDFR, PyANS, and PyUFGT) and key regulatory factors (PyMYB10, PyMYB114, and PybHLH3). Through analysis of transcriptome data of treated pear callus and RT-qPCR assay, a differentially expressed gene PyNAC42 was identified as significantly induced by the N deficiency condition. Overexpression of PyNAC42 promoted anthocyanin accumulation in “Zaosu” pear peels. Additionally, dual luciferase assay and yeast one-hybrid assay demonstrated that PyNAC42 could not directly activate the expression of PyDFR, PyANS, and PyUFGT. Furthermore, yeast two-hybrid and pull-down assays confirmed that PyNAC42 interacted with PyMYB10 both in vivo and in vitro. Co-expression of PyNAC42 and PyMYB10 significantly enhanced anthocyanin accumulation in “Zaosu” pear peels. Dual luciferase assay showed that PyNAC42 significantly enhanced the activation of PyDFR, PyANS, and PyUFGT promoters by interacting with PyMYB10, which suggests that PyNAC42 can form the PyNAC42-PyMYB10 complex to regulate anthocyanin biosynthesis in pear. Thus, the molecular mechanism underlying anthocyanin biosynthesis induced by N deficiency is preliminarily elucidated. Our finding has expanded the regulatory network of anthocyanin biosynthesis and enhanced our understanding of the mechanisms underlying nutrient deficiency modulates anthocyanin biosynthesis in pear. Full article
(This article belongs to the Section Genetics, Genomics, Breeding, and Biotechnology (G2B2))
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Effects of the MS control (MS) and N-deficiency treatment (-N) on anthocyanin accumulation in pear callus. (<b>B</b>) Total anthocyanin contents of pear callus subjected to two treatments. The <span class="html-italic">X</span>-axis represents (MS) and (-N) treatments. (<b>C</b>) Expression levels of anthocyanin biosynthesis-related genes in pear callus. The data were normalized to the <span class="html-italic">PyGAPDH</span> expression level. The error bars in (<b>B</b>,<b>C</b>) represent the SD (<span class="html-italic">n</span> = 3 independent biological replicates). Student’s <span class="html-italic">t</span>-test was used for statistical analysis (*** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Analysis of RNA-seq data. (<b>A</b>) Summary of RNA-sequencing data. (<b>B</b>) The volcano plot showed the DEGs in N-deficiency treatment (-N) vs. MS control (MS) comparison. The up- and downregulated genes were represented by red and blue dots, respectively. The dashed lines represent thresholds for differentially expressed genes, with vertical dashed lines indicating |log2 Fold Change| = 1 and the horizontal dashed line indicating -log10 (FDR) = 2.</p>
Full article ">Figure 3
<p>GO enrichment and KEGG pathway analyses of DEGs from N-deficiency treatment vs. MS control comparison. Bubble diagram showing enriched GO (<b>A</b>) biological process, (<b>B</b>) cellular component, and (<b>C</b>) molecular function terms. (<b>D</b>) Bubble diagram showing enriched KEGG pathways terms. The color represents the significance. The red color corresponds to a higher significance, while the blue color corresponds to a lower significance. The important enriched terms were marked in red.</p>
Full article ">Figure 4
<p>RT-qPCR validation of nine candidate DEGs from RNA-Seq data analysis. (<b>A</b>–<b>I</b>) The relative expression of nine candidate DEGs in MS control (MS) and N-deficiency treatment (-N). The columns represent the relative expression levels from RT-qPCR (left y axis) The lines represent FPKM values from the RNA-seq data (right y axis). The Pearson correlation coefficient (r) values between FPKM and the relative expression levels are shown. The error bars mean the SD (<span class="html-italic">n</span> = 3). Student’s <span class="html-italic">t</span>-test was used for statistical analysis (*** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>(<b>A</b>) Phylogenetic relationship between PyNAC42 (marked with a red frame) and other NAC TF from different species (for a complete list of species, see <a href="#app1-horticulturae-10-00980" class="html-app">Table S2</a>). (<b>B</b>) Amino acid sequence alignment of NAC TF PyNAC42 (marked with a red frame), AtNAC42, MdNAC2, AtNAC78, PbrNSC, and PpBL. (*) represents the omitted number of amino acids. The shading color indicates the level of conservation in the amino acid sequence, which is categorized into three levels denoted by black, pink, and blue respectively based on their degree of conservatism. (<b>C</b>) Detection of fluorescence signals in tobacco leaf transformed with PyNAC42-GFP. In the detection, mCherry was used to show the nuclear localization. Bars represent 20 μm.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>D</b>) Functional analysis of the role of PyNAC42 in anthocyanin biosynthesis in “Zaosu” pear peels. (<b>A</b>) Phenotypes of “Zaosu” pear peels transiently transformed with the empty vector (EV) and a vector overexpressing PyNAC42 (<span class="html-italic">PyNAC42</span>-OE). The injection areas were outlined in dash circles. (<b>B</b>) Anthocyanin contents of “Zaosu” pear peels. The <span class="html-italic">X</span>-axis represents two treatments. Expression levels of (<b>C</b>) <span class="html-italic">PyNAC42</span> and (<b>D</b>) anthocyanin biosynthesis structural genes in pear peels. The data were normalized to the <span class="html-italic">PyGAPDH</span> expression level. (<b>E</b>,<b>F</b>) Effect of PyNAC42 on the promoter activity of anthocyanin biosynthesis structural genes. (<b>E</b>) Vector construction for effector and reporters used in dual luciferase assay. (<b>F</b>) Activation effect of PyNAC42 on promoters of <span class="html-italic">PyDFR</span>, <span class="html-italic">PyANS</span>, and <span class="html-italic">PyUFGT</span>. The error bars are SD (<span class="html-italic">n</span> = 3). Student’s <span class="html-italic">t</span>-test was used for statistical analysis (*** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>(<b>A</b>,<b>B</b>) PyNAC42 interacts with PyMYB10 both in vivo and in vitro. (<b>A</b>) Y2H assay of interactions between PyNAC42 and PyMYB10, PybHLH3, and PyMYB114. Yeast cells were plated on SD/-Leu/-Trp, SD/-Leu/-Trp/-His/-Ade, and SD/-Leu/-Trp/-His/-Ade/+X-α-Gal media. Growth trend aligned with that of the positive control indicates an interaction between the two proteins. (<b>B</b>) Pull-down assay. GST and the PyMYB10-GST fusion protein were bound to GST-Sefinose Resin and incubated with the PyNAC42-HIS protein, respectively. The PyMYB10-GST fusion protein could pull down the PyNAC42-HIS fusion protein, which indicated that PyNAC42 interacts with PyMYB10 in vitro. (<b>C</b>–<b>F</b>) Functional analysis of the effects of the interaction between PyNAC42 and PyMYB10 on anthocyanin biosynthesis in “Zaosu” pear peels. (<b>C</b>) Phenotypes, (<b>D</b>) total anthocyanin contents, and (<b>E</b>,<b>F</b>) expression levels of <span class="html-italic">PyNAC42</span>, <span class="html-italic">PyMYB10</span>, <span class="html-italic">PyDFR</span>, <span class="html-italic">PyANS</span>, and <span class="html-italic">PyUFGT</span> in “Zaosu” pear peels. The <span class="html-italic">X</span>-axis represents four treatments. <span class="html-italic">PyGAPDH</span> was used as reference gene for RT-qPCR, and one-way ANOVA with Tukey’s post hoc test was used for statistical analysis. Different letters indicate a significant difference between means.</p>
Full article ">Figure 8
<p>Effects of PyNAC42-PyMYB10 complex on the promoter activation of LBGs. (<b>A</b>) Vector construction for effectors and reporters used in dual luciferase assays. (<b>B</b>) Activation of the promoters of <span class="html-italic">PyDFR</span>, <span class="html-italic">PyANS</span>, and <span class="html-italic">PyUFGT</span> by four treatments. <span class="html-italic">X</span>-axis represents the group of four treatments. Specifically, EV was added to balance the concentration when PyNAC42/PyMYB10 was co-injected with each promoter. The error bars are SD (<span class="html-italic">n</span> = 3). One-way ANOVA with Tukey’s post hoc test was used for statistical analysis, and different letters indicate <span class="html-italic">p</span> &lt; 0.05. The LUC/REN ratio represents the promoter activity.</p>
Full article ">Figure 9
<p>Proposed model for underlying N deficiency induced anthocyanin biosynthesis mediated by PyNAC42.</p>
Full article ">
14 pages, 3173 KiB  
Article
MiR-206 Suppresses Triacylglycerol Accumulation via Fatty Acid Elongase 6 in Dairy Cow Mammary Epithelial Cells
by Xin Zhao, Yu Liu, Yupeng Li, Yuxin Zhang, Chunlei Yang and Dawei Yao
Animals 2024, 14(17), 2590; https://doi.org/10.3390/ani14172590 - 6 Sep 2024
Viewed by 248
Abstract
Cow milk possesses high nutritional value due to its rich array of beneficial fatty acids. It is important to understand the mechanisms involved in lipid metabolism in dairy cows. These mechanisms are driven by a complex molecular regulatory network. In addition, there are [...] Read more.
Cow milk possesses high nutritional value due to its rich array of beneficial fatty acids. It is important to understand the mechanisms involved in lipid metabolism in dairy cows. These mechanisms are driven by a complex molecular regulatory network. In addition, there are many regulatory factors involved in the process of fatty acid metabolism, including transcription factors and non-coding RNAs, amongst others. MicroRNAs (miRNAs) can regulate the expression of target genes and modulate various biological processes, including lipid metabolism. Specifically, miR-206 has been reported to impair lipid accumulation in nonruminant hepatocytes. However, the effects and regulatory mechanisms of miR-206 on lipid metabolism in bovine mammary cells remain unclear. In the present study, we investigated the effects of miR-206 on lipid-related genes and TAG accumulation. The direct downstream gene of miR-206 was subsequently determined via a dual-luciferase assay. Finally, the fatty acid content of bovine mammary epithelial cells (BMECs) upon ELOVL6 inhibition was examined. The results revealed that miR-206 overexpression significantly decreased triacylglycerol (TAG) concentration and abundances of the following: acetyl-coenzyme A carboxylase alpha (ACACA); fatty acid synthase (FASN); sterol regulatory element binding transcription factor 1 (SREBF1); diacylglycerol acyltransferase 1 (DGAT1); 1-acylglycerol-3-phosphate O-acyltransferase 6 (AGPAT6); lipin 1 (LPIN1); and fatty acid elongase 6 (ELOVL6). Overexpression of miR-206 was also associated with an increase in patatin-like phospholipase domain-containing 2 (PNPLA2), while inhibition of miR-206 promoted milk fat metabolism in vitro. In addition, we found that ELOVL6 is a direct target gene of miR-206 through mutation of the binding site. Furthermore, ELOVL6 intervention significantly decreased the TAG levels and elongation indexes of C16:0 and C16:1n-7 in BMECs. Finally, ELOVL6 siRNA partially alleviated the increased TAG accumulation caused by miR-206 inhibition. In summary, we found that miR-206 inhibits milk fatty acid synthesis and lipid accumulation by targeting ELOVL6 in BMECs. The results presented in this paper may contribute to the development of strategies for enhancing the quality of cow milk and its beneficial fatty acids, from the perspective of miRNA–mRNA networks. Full article
Show Figures

Figure 1

Figure 1
<p>Expression analysis of miR-206 in BMECs. (<b>A</b>) Detection of transfection efficiency of miR-206 mimics. (<b>B</b>) Detection of transfection efficiency of miR-206 inhibitor. (<b>C</b>,<b>D</b>) Effect of overexpression or interference of miR-206 on intracellular triglyceride content. **: <span class="html-italic">p</span> &lt; 0.01 v. control; * <span class="html-italic">p</span> &lt; 0.05 v. control.</p>
Full article ">Figure 2
<p>Effect of overexpression or interference of miR-206 on expression of genes related to lipid accumulation and lipid metabolism. (<b>A</b>) Variation in genes related to lipid metabolism after overexpression of miR-206. (<b>B</b>) Variation in genes related to lipid metabolism after miR-206 inhibition. Quantitative PCR data were calculated using the 2<sup>−ΔΔCt</sup> method and are presented as mean ± standard error of the means for three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 v. control; * <span class="html-italic">p</span> &lt; 0.05 v. control.</p>
Full article ">Figure 3
<p>Prediction of miR-206 target genes. (<b>A</b>) Analysis of binding sites of miR-206 on <span class="html-italic">ELOVL6</span> 3′TR in different species. (<b>B</b>) Results of site-directed mutation of miR-206 target sites. (<b>C</b>,<b>D</b>) Dual-luciferase assay results of miR-206 mimics or inhibitors co-transfected with recombinant vector. ** <span class="html-italic">p</span> &lt; 0.01 v. control.</p>
Full article ">Figure 4
<p>Detection of <span class="html-italic">ELOVL6</span> interference efficiency and effect of <span class="html-italic">ELOVL6</span> inhibition on genes related to lipid metabolism. (<b>A</b>) All three siELOVL6 could significantly interfere with the expression of <span class="html-italic">ELOVL6</span>. (<b>B</b>) Expression changes in genes related to de novo fatty acid synthesis and desaturation after interference with <span class="html-italic">ELOVL6</span>. (<b>C</b>) Expression changes in genes related to fatty acid uptake and transport after <span class="html-italic">ELOVL6</span> inhibition. (<b>D</b>) Expression of TAG synthesis-related genes after <span class="html-italic">ELOVL6</span> inhibition. (<b>E</b>) Expression of TAG degradation related genes after <span class="html-italic">ELOVL6</span> interference. * <span class="html-italic">p</span> &lt; 0.05 v. control, ** <span class="html-italic">p</span> &lt; 0.01 v. control.</p>
Full article ">Figure 5
<p>Changes in FA desaturation index and elongation index after overexpression or inhibition of <span class="html-italic">ELOVL6</span> in BMEC. (<b>A</b>,<b>B</b>) Desaturation index of 16:1 and 18:1 after the treatment of siRNA-Scr or siRNA-ELOVL6. (<b>C</b>,<b>D</b>) Elongation index of 16:0 and 16:1n-7 after treatment with siRNA-Scr or siRNA-ELOVL6. Data are presented as mean ± standard error of the means for 3 individual experiments. * <span class="html-italic">p</span> &lt; 0.05 v. control.</p>
Full article ">Figure 6
<p>Effects of <span class="html-italic">ELOVL6</span> interference on fatty acid composition and triglyceride content in cells. (<b>A</b>) Changes in intracellular FA fractions after <span class="html-italic">ELOVL6</span> inhibition in BMEC. (<b>B</b>) TAG levels in cells transfected with siRNA-Scr or siRNA-ELOVL6. (<b>C</b>) TAG levels in cells transfected with negative control inhibitor + control siRNA, inhibitor miR-206+ control siRNA, and inhibitor miR-206+ sirNA-ELOVL6. * <span class="html-italic">p</span> &lt; 0.05 v. control.</p>
Full article ">
19 pages, 7254 KiB  
Article
Prolactin Modulates the Proliferation and Secretion of Goat Mammary Epithelial Cells via Regulating Sodium-Coupled Neutral Amino Acid Transporter 1 and 2
by Xiaoyue Ma, Hanling Liu, Wentao Li, Jianguo Chen, Zhenliang Cui, Zixia Wang, Changmin Hu, Yi Ding and Hongmei Zhu
Cells 2024, 13(17), 1461; https://doi.org/10.3390/cells13171461 - 30 Aug 2024
Viewed by 390
Abstract
The prolactin (PRL) hormone is a major regulator of mammary gland development and lactation. However, it remains unclear whether and how PRL contributes to mammary epithelial cell proliferation and secretion. The Boer and Macheng black crossbred goats are superior in reproduction, meat, and [...] Read more.
The prolactin (PRL) hormone is a major regulator of mammary gland development and lactation. However, it remains unclear whether and how PRL contributes to mammary epithelial cell proliferation and secretion. The Boer and Macheng black crossbred goats are superior in reproduction, meat, and milk, and are popular in Hubei province. To elucidate the mechanisms of PRL on mammary growth and lactation, to improve the local goat economic trade, we have performed studies on these crossbred goats during pregnancy and early lactation, and in goat mammary epithelial cells (GMECs). Here, we first found that the amino acid transporters of SNAT1 and SNAT2 expression in vivo and in vitro were closely associated with PRL levels, the proliferation and secretion of GMECs; knockdown and over-expression of SNAT1/2 demonstrated that PRL modulated the proliferation and lactation of GMECs through regulating SNAT1/2 expression. Transcriptome sequencing and qPCR assays demonstrated the effect of PRL on the transcriptional regulation of SNAT1 and SNAT2 in GMECs. Dual-luciferase reporter gene assays further verified that the binding of the potential PRL response element in the SNAT1/2 promoter regions activated SNAT1/2 transcription after PRL stimulation. Additionally, silencing of either PRLR or STAT5 nearly abolished PRL-stimulated SNAT1/2 promoter activity, suggesting PRLR–STAT5 signaling is involved in the regulation of PRL on the transcriptional activation of SNAT1/2. These results illustrated that PRL modulates the proliferation and secretion of GMECs via PRLR–STAT5-mediated regulation of the SNAT1/2 pathway. This study provides new insights into how PRL affects ruminant mammary development and lactation through regulation of amino acid transporters. Full article
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the whole experiment. Youth: The goats are in the stage of puberty; Pd91: The goats are in the stage of pregnancy, day 91; Pd137: The goats are in the stage of pregnancy, day 137; Ld4: The goats are in the stage of lactation, day 4; Ld31: The goats are in the stage of lactation, day 31; PRL: prolactin; SNAT1/2: sodium-coupled neutral amino acid transporter 1/2; GMECs: goat mammary epithelial cells (diagram by Figdraw).</p>
Full article ">Figure 2
<p>PRL and SNAT1/2 levels in different stages of goats. PRL: prolactin; SNAT1/2: sodium-coupled neutral amino acid transporter 1/2. (<b>A</b>) Serum PRL concentrations in goats. (<b>B</b>) SNAT1/2 protein expressions in different stages of goat mammary tissue. Youth: The goats are in the stage of puberty; Pd91: The goats are in the stage of pregnancy, day 91; Pd137: The goats are in the stage of pregnancy, day 137; Ld4: The goats are in the stage of lactation, day 4; Ld31: The goats are in the stage of lactation, day 31. Data are represented as means ± SD. Values with different lowercase letters indicate significant difference (<span class="html-italic">p</span> &lt; 0.05). Original images can be found in <a href="#app1-cells-13-01461" class="html-app">Supplementary Files</a>.</p>
Full article ">Figure 3
<p>(<b>A</b>) GMECs are treated with different concentrations of PRL (0, 20, 50, 100, 200, 400, 800, and 1600 ng/mL) for 12, 24, and 48 h. Cell viability is determined by CCK-8 assay. (<b>B</b>) GMECs are treated with different concentrations of PRL (0, 200, and 1600 ng/mL) for 48 h. Cell viability is determined by EdU assay (400 μm). (<b>C</b>) Protein levels of SNAT1/2 in GMECs treated with PRL are detected by western blot analysis. Data are represented as means ± SD, * <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. Values with different lowercase letters indicate significant difference (<span class="html-italic">p</span> &lt; 0.05). Original images can be found in <a href="#app1-cells-13-01461" class="html-app">Supplementary Files</a>.</p>
Full article ">Figure 4
<p>The effect of PRL on GMECs’ lactation and SNAT1/2 expression. (<b>A</b>) Western blot analysis of milk protein and SNAT1/2 in GMECs treated with PRL. (<b>B</b>) QPCR analysis of <span class="html-italic">CSN</span> and <span class="html-italic">BLG</span> mRNA levels in GMECs treated with PRL. <span class="html-italic">CSN</span>: Casein; <span class="html-italic">BLG</span>: Beta-lactoglobulin. (<b>C</b>) GPO-PAP assay of TG content in the supernatant of GMECs after induction of lactation. EGF: epidermal growth factor; ITS: insulin-transferrin-selenium; HC: hydrocortisone; Control: GMECs are treated with only culture medium; IL: GMECs are treated with an induced lactation system (EGF + ITS + HC); IL + PRL2500/5000/7500/10,000 ng/mL: GMECs are treated with IL and 2500, 5000, 7500, or 10,000 ng/mL PRL. Data are represented as means ± SD. Values with different lowercase letters indicate significant difference (<span class="html-italic">p</span> &lt; 0.05). Original images can be found in <a href="#app1-cells-13-01461" class="html-app">Supplementary Files</a>.</p>
Full article ">Figure 5
<p>The effect of knockdown and overexpression of SNAT1/2 on GMECs’ proliferation and lactation. (<b>A</b>) The effect of knockdown and overexpression of SNAT1 or SNAT2 in GMECs on mRNA levels of <span class="html-italic">SNAT1/2</span>, <span class="html-italic">CSN</span>, and <span class="html-italic">BLG</span>. (<b>B</b>) Protein changes of SNAT1 and SNAT2 after they are knocked down or are overexpressed in GMECs. (<b>C</b>) Cell proliferative activity is detected by CCK-8 assay after knockdown or overexpression of SNAT1/2. (<b>D</b>) Cell proliferation activity is detected by EdU assay after knockdown or overexpression of SNAT1/2 (400 μm). (<b>E</b>) CSN and BLG protein changes in GMECs after being knocked down or overexpressed SNAT1/2. Control: GMECs did not receive any treatment; NC: GMECs transfected with corresponding scrambled siRNA for SNAT1 or SNAT2 as negative control; SiSNAT1/2: GMECs are transfected with siRNAs for SNAT1/2; C1: pEGFP-C1, GMECs are transfected with pEGFP-C1 empty vector (without a SNAT1 or SNAT2 sequence insert); C1-SNAT1/2: pEGFP-SNAT1/2-C1, GMECs are transfected with an overexpression vector for SNAT1/2. Data are represented as means ± SD, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Original images can be found in <a href="#app1-cells-13-01461" class="html-app">Supplementary Files</a>.</p>
Full article ">Figure 6
<p>SNAT1/2 nucleic acid level detection. (<b>A</b>) RNA levels of <span class="html-italic">SNAT1/2</span> are detected by transcriptome sequencing after treatment with 200 ng/mL of PRL. (<b>B</b>) qPCR assays of <span class="html-italic">SNAT1/2</span> mRNA levels in GMECs after being treated with PRL (0, 100, 200, 400, 800 ng/mL). (<b>C</b>) qPCR assays for mRNA levels of <span class="html-italic">SNAT1/2</span> in GMECs after being treated with induced lactation system (EGF + ITS + HC) supplemented with PRL (0, 2500, 5000, 7500, 10,000 ng/mL). EGF: epidermal growth factor; ITS: insulin-transferrin-selenium; HC: hydrocortisone; IL: GMECs are treated with an induced lactation system with EGF, ITS, and HC; IL + PRL2500/5000/7500/10,000 ng/mL: GMECs are treated with IL and 2500, 5000, 7500, or 10,000 ng/mL PRL. Data are represented as means ± SD, NS: no statistical significance, ** <span class="html-italic">p</span> &lt; 0.01. Values with different lowercase letters indicate significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>The effect of PRL on SNAT1/2 transcription initiation. (<b>A</b>) Dual-luciferase reporter gene assays on SNAT1/2 gene transcriptional activity in GMECs treated with PRL. (<b>B</b>) Dual-luciferase reporter gene assays of truncated SNAT1/2 promoter activity in GMECs treated with PRL. pGL3-SNAT1/2: pGL3 vectors containing the SNAT1/2 promoter sequence (−1724/+166)/(−1695/+95); PRL (200 ng/mL): GMECs are treated with 200 ng/mL prolactin; Control: GMECs did not receive any treatment; SNAT1-1/2/3: The SNAT1 promoter sequence is truncated into three segments, which are respectively inserted into pGL3 vectors; SNAT2-1/2/3: The SNAT2 promoter sequence is truncated into three segments, which are respectively inserted into pGL3 vectors; pGL3-b: GMECs are transfected into pGL3-basic vector. Data are represented as Means ± SD, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Values with different lowercase letters indicate significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>The effect of PRLR-STAT5 on transcriptional activity of SNAT1/2 regulated by PRL. PRLR: prolactin receptor; STAT5: signal transducers and activators of transcription 5. (<b>A</b>) SNAT1/2 promoter fluorescence activity is measured after PRLR is knocked down. (<b>B</b>) Promoter fluorescence activity of SNAT1/2 in GMECs after knocked down STAT5. (<b>C</b>) Protein and mRNA levels of STAT5 and SNAT1/2 are measured after knockdown of PRLR. (<b>D</b>) Protein and mRNA levels of SNAT1/2 in GMECs after knocked down STAT5. pSNAT1/2-NC: negative control for PRLR/STAT5 siRNAs during SNAT1/2 promoter assay; pSNAT1/2 + siPRLR/siSTAT5: The SNAT1/2 promoter–reporter assay was accompanied by PRLR or STAT5 RNA interference; pSNAT1/2: The SNAT1/2 promoter–reporter assay accompanied with pGL3-basic vector; NC: the negative control for PRLR or STAT5 RNA interference; siPRLR/siSTAT5: siRNAs for PRLR or STAT5. Data are represented as means ± SD, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Original images can be found in <a href="#app1-cells-13-01461" class="html-app">Supplementary Files</a>.</p>
Full article ">
45 pages, 3449 KiB  
Review
Non-Muscle Myosin II A: Friend or Foe in Cancer?
by Wasim Feroz, Briley SoYoung Park, Meghna Siripurapu, Nicole Ntim, Mary Kate Kilroy, Arwah Mohammad Ali Sheikh, Rosalin Mishra and Joan T. Garrett
Int. J. Mol. Sci. 2024, 25(17), 9435; https://doi.org/10.3390/ijms25179435 - 30 Aug 2024
Viewed by 394
Abstract
Non-muscle myosin IIA (NM IIA) is a motor protein that belongs to the myosin II family. The myosin heavy chain 9 (MYH9) gene encodes the heavy chain of NM IIA. NM IIA is a hexamer and contains three pairs of peptides, [...] Read more.
Non-muscle myosin IIA (NM IIA) is a motor protein that belongs to the myosin II family. The myosin heavy chain 9 (MYH9) gene encodes the heavy chain of NM IIA. NM IIA is a hexamer and contains three pairs of peptides, which include the dimer of heavy chains, essential light chains, and regulatory light chains. NM IIA is a part of the actomyosin complex that generates mechanical force and tension to carry out essential cellular functions, including adhesion, cytokinesis, migration, and the maintenance of cell shape and polarity. These functions are regulated via light and heavy chain phosphorylation at different amino acid residues. Apart from physiological functions, NM IIA is also linked to the development of cancer and genetic and neurological disorders. MYH9 gene mutations result in the development of several autosomal dominant disorders, such as May-Hegglin anomaly (MHA) and Epstein syndrome (EPS). Multiple studies have reported NM IIA as a tumor suppressor in melanoma and head and neck squamous cell carcinoma; however, studies also indicate that NM IIA is a critical player in promoting tumorigenesis, chemoradiotherapy resistance, and stemness. The ROCK-NM IIA pathway regulates cellular movement and shape via the control of cytoskeletal dynamics. In addition, the ROCK-NM IIA pathway is dysregulated in various solid tumors and leukemia. Currently, there are very few compounds targeting NM IIA, and most of these compounds are still being studied in preclinical models. This review provides comprehensive evidence highlighting the dual role of NM IIA in multiple cancer types and summarizes the signaling networks involved in tumorigenesis. Furthermore, we also discuss the role of NM IIA as a potential therapeutic target with a focus on the ROCK-NM IIA pathway. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>The expression and alteration profile of <span class="html-italic">MYH9</span> at the pancancer level. (<b>A</b>) <span class="html-italic">MYH9</span> mRNA expression from 32 TCGA datasets. (<b>B</b>) The alteration profile of <span class="html-italic">MYH9</span> from the same 32 TCGA datasets. The expression and alteration frequency data were obtained from cBioPortal (<a href="http://www.cbioportal.org" target="_blank">www.cbioportal.org</a>).</p>
Full article ">Figure 2
<p>This figure illustrates the two assembly states of non-muscle myosin II (NMII): the 10S assembly-incompetent state and the 6S assembly-competent state. In the 10S assembly-incompetent state (<b>left</b>), the myosin molecule is folded, with the globular head and heavy chain regions interacting through various tail-binding sites, leading to a compact structure. Many intramolecular interactions keep the 10S state in an inactive stable form (<a href="#ijms-25-09435-t001" class="html-table">Table 1</a>). The interactions involve the Blocked Head (BH), which is the myosin head prevented from binding to actin, and the Free Head (FH), another myosin head that is inhibited but not directly involved in actin binding in the 10S state. The transition to the 6S assembly-competent state (<b>right</b>) occurs upon the phosphorylation of the regulatory light chain (RLC), resulting in an extended, active conformation where the heavy chain regions are aligned, allowing for actin binding and ATPase activity, which are essential for NMII’s role in cell contractility and motility. ELC: Essential Light Chain; RLC: Regulatory Light Chain; FH: Free Head; BH: Blocked Head; TF: Tail–Free Head Interaction; TB: Tail–Blocked Head Interaction; TT: Tail–Tail Interaction.</p>
Full article ">Figure 3
<p>The figure shows the specific kinases involved in the phosphorylation of serine and threonine residues of both RLCs and heavy chains. PKC, protein kinase C; MLCK, myosin light chain kinase; ROCK, Rho-associated protein kinase; TRPM7, transient receptor potential melastatin 7; PKCβ, protein kinase Cβ; CK II, casein kinase II. The figure was adapted from Pecci et al. [<a href="#B28-ijms-25-09435" class="html-bibr">28</a>].</p>
Full article ">Figure 4
<p>This figure illustrates the regulation of myosin II filament formation and activity. Specific serine residues on the myosin heavy chain (S1916, S1927, and S1943) are phosphorylated by various kinases, including MHCKA, MHCKB, MHCKC, TRPM6, TRPM7, PKC, and CK2. The phosphorylation of these sites is depicted as favoring the filamentous state of myosin, which is essential for mechanotransduction and ATP hydrolysis-driven interaction with actin filaments. Phosphatases are the enzymes responsible for dephosphorylation, which may reverse the phosphorylation effect, potentially leading to a shift back to the monomeric state of myosin. Mts1, also known as S100A4, is a calcium-binding protein that regulates myosin II function by modulating filament assembly. It binds to the myosin heavy chain, influencing the balance between monomeric and filamentous forms of myosin II. Mts1 typically inhibits filament formation, thereby controlling myosin’s contractile activity and its ability to interact with actin.</p>
Full article ">Figure 5
<p>The figure shows the structure and orientation of cell migration in the 2D environment. At the front, actin filaments within lamellipodia and filopodia are oriented with their rapidly polymerizing ends in the forward direction. In the main body, actin and myosin filaments form bipolar structures to aid in cell retraction. NM IIA and NM IIB show distinct localizations inside the cell, with NM IIA predominantly being found at the leading edge where actin dynamics are most active. NM IIB is predominant toward the rear end. The region between the leading and trailing edges contains varying concentrations of NM IIA and NM IIB. Additional molecules, such as RhoA, Rac1, Cdc42, Ca<sup>2+</sup> ions, and αPKC, also play significant roles in this cellular organization and migration process.</p>
Full article ">Figure 6
<p>The formation of plasma membrane blebs consists of three phases: initiation, expansion, and retraction. Blebbing stimuli, such as Ca<sup>2+</sup> influx and apoptosis, induce the initiation of membrane protrusion. Actomyosin contractility drives the expansion of blebs, which are devoid of the F-actin cortex. Rho-ROCK signaling then drives bleb retraction via actomyosin contractility. NM IIA contractile forces promote bleb retraction.</p>
Full article ">Figure 7
<p>A schematic representation of the <span class="html-italic">MYH9</span> exons with common mutations found in patients with <span class="html-italic">MYH9</span>-RD. The color coding of exon organization is as follows: black, motor domain; green, neck; orange, coiled coil domain; and brown, non-helical tail.</p>
Full article ">
20 pages, 3712 KiB  
Article
Overexpression of SlALC Increases Drought and Salt Tolerance and Affects Fruit Dehiscence in Tomatoes
by Zihan Gao, Yuqing Tu, Changguang Liao, Pengyu Guo, Yanling Tian, Ying Zhou, Qiaoli Xie, Guoping Chen and Zongli Hu
Int. J. Mol. Sci. 2024, 25(17), 9433; https://doi.org/10.3390/ijms25179433 - 30 Aug 2024
Viewed by 299
Abstract
The bHLH transcription factors are important plant regulators against abiotic stress and involved in plant growth and development. In this study, SlALC, a gene coding for a prototypical DNA-binding protein in the bHLH family, was isolated, and SlALC-overexpression tomato (SlALC [...] Read more.
The bHLH transcription factors are important plant regulators against abiotic stress and involved in plant growth and development. In this study, SlALC, a gene coding for a prototypical DNA-binding protein in the bHLH family, was isolated, and SlALC-overexpression tomato (SlALC-OE) plants were generated by Agrobacterium-mediated genetic transformation. SlALC transgenic lines manifested higher osmotic stress tolerance than the wild-type plants, estimated by higher relative water content and lower water loss rate, higher chlorophyll, reducing sugar, starch, proline, soluble protein contents, antioxidant enzyme activities, and lower MDA and reactive oxygen species contents in the leaves. In SlALC-OE lines, there were more significant alterations in the expression of genes associated with stress. Furthermore, SlALC-OE fruits were more vulnerable to dehiscence, with higher water content, reduced lignin content, SOD/POD/PAL enzyme activity, and lower phenolic compound concentrations, all of which corresponded to decreased expression of lignin biosynthetic genes. Moreover, the dual luciferase reporter test revealed that SlTAGL1 inhibits SlALC expression. This study revealed that SlALC may play a role in controlling plant tolerance to drought and salt stress, as well as fruit lignification, which influences fruit dehiscence. The findings of this study have established a foundation for tomato tolerance breeding and fruit quality improvement. Full article
(This article belongs to the Special Issue Advances in Tomato Breeding and Molecular Research)
Show Figures

Figure 1

Figure 1
<p>Bioinformatics analysis, subcellular localization, and expression pattern of <span class="html-italic">SlALC</span>. (<b>A</b>) Homology analysis of SlALC; the black box indicates the PIFs domain. (<b>B</b>) SlALC phylogenetic tree. The accession numbers are SlALC (NP_001361317, highlighted in red box), StSPATULA-like (XP_049391100), CaALC (NP_001361321), NtSPATULA-like (XP_009627217.1), InSPATULA-like (XP_019182870), AtALC (AT5G67110), AtPIF3 (AT1G09530), and SlSPT (NP_001361318.1). (<b>C</b>) Subcellular localization assay of SlALC protein. GFP: green fluorescent protein; RFP: red fluorescent protein. Red fluorescent protein is used to locate the nucleus. Scale bar = 50 µm. (<b>D</b>) Quantitative RT–PCR analysis of the expression of the <span class="html-italic">SlALC</span> gene in roots (RT), stems (ST), young leaves (YL), mature leaves (ML), senescent leaves (SL), sepals (SE), flowers (FL), and fruits (pericarp) at immature green (IMG), mature green (MG), breaker (B), B+ 4 and B + 7 stages. (<b>E</b>,<b>F</b>) Expression patterns of <span class="html-italic">SlALC</span> in leaves under the dehydration and salt treatments. (<b>G</b>) Relative expression of <span class="html-italic">SlALC</span> in leaves of <span class="html-italic">SlALC</span>-OE T0 plants. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Mannitol and salt tolerance analysis of WT and transgenic seeds and seedlings. (<b>A</b>) Germination phenotype of WT and <span class="html-italic">SlALC</span>-OE seeds under 0, 50, 100, and 150 mM Mannitol treatments for 3 weeks. Scale bar = 1 cm. (<b>B</b>–<b>E</b>) Seed germination rates of WT and <span class="html-italic">SlALC</span>-OE lines under 0, 50, 100, and 150 mM Mannitol treatment, respectively. (<b>F</b>) Phenotypic map of <span class="html-italic">SlALC</span>-OE seedlings under 0, 75, 150, and 300 mM mannitol treatment. Scale bar = 5 cm. (<b>G</b>,<b>H</b>) Root and seedling length of WT and <span class="html-italic">SlALC</span>-OE plants under normal and 0, 75, 150, and 300 mM mannitol treatment, respectively. (<b>I</b>) Germination phenotype of WT and <span class="html-italic">SlALC</span>-OE seeds under 0, 40, and 80 mM NaCl treatments for 3 weeks. Scale bar = 1 cm. (<b>J</b>–<b>L</b>) Seed germination rates of WT and <span class="html-italic">SlALC</span>-OE lines under 0, 40, and 80 mM NaCl treatments, respectively. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>The phenotype of WT and <span class="html-italic">SlALC</span>-OE transgenic tomato plants under drought and salt stress. (<b>A</b>) Growth Status of WT and <span class="html-italic">SlALC</span>-OE under drought and salt stress. Scale bar = 10 cm. (<b>B</b>–<b>I</b>) Comparisons of water loss rate (<b>B</b>), relative water content (<b>C</b>), soluble protein content (<b>D</b>), proline content (<b>E</b>), total chlorophyll content (<b>F</b>), reducing sugar content (<b>G</b>), starch content (<b>I</b>). (<b>H</b>) KI/I2 staining. Scale bar = 1 cm. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Comparison of cell damage indicators, ROS content between WT and <span class="html-italic">SlALC</span>-OE lines under drought and salt treatments. (<b>A</b>) Trypan blue staining, (<b>B</b>) MDA content, (<b>C</b>) relative conductivity, (<b>D</b>) NBT staining, (<b>E</b>) H<sub>2</sub>O<sub>2</sub> content, (<b>F</b>) CAT activity, (<b>G</b>) DAB staining, (<b>H</b>) SOD activity, (<b>I</b>) POD activity. Scale bar = 1 cm. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Expression levels of stress-related genes before and after drought and salt treatment in WT and <span class="html-italic">SlALC</span>-OE mature leaves. (<b>A</b>–<b>O</b>) qRT-PCR analysis of the expression levels of <span class="html-italic">PR1</span>, <span class="html-italic">PR5</span>, <span class="html-italic">Lea</span>, <span class="html-italic">Prg</span>, <span class="html-italic">Dhn</span>, <span class="html-italic">P5CS</span>, <span class="html-italic">Cat1</span>, <span class="html-italic">Cat2</span>, <span class="html-italic">FRK2</span>, <span class="html-italic">BoGH3B</span>, <span class="html-italic">Cab7</span>, <span class="html-italic">Golden2-like1</span>, <span class="html-italic">Golden2-like2</span>, <span class="html-italic">Sgr1</span>, <span class="html-italic">DCL</span>. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p><span class="html-italic">SlALC</span>-OE lines with cracks in fruit. (<b>A</b>) The phenotype of cracking fruit after rain. Black arrow indicates tomato fruit cracks. (<b>B</b>) Fruit cracks in <span class="html-italic">SlALC</span>-OE lines compared with WT. (<b>C</b>) Fruit cracking rate. (<b>D</b>) Fruit water content. Scale bar = 1 cm. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p><span class="html-italic">SlALC</span>-OE fruits are less lignified compared with wild-type. (<b>A</b>–<b>C</b>) Comparison of wounds in transgenic and WT fruits. Fruit at 0 days of treatment (<b>A</b>). Fruits at 7 days of cultivation (<b>B</b>). Fruit shoulders at 7 days of cultivation (<b>C</b>). (<b>D</b>) Resorcinol staining. (<b>E</b>–<b>I</b>) Lignin content (<b>E</b>). Total phenol content (<b>F</b>). POD (<b>G</b>), SOD (<b>H</b>), and PAL (<b>I</b>) activity. Scale bar = 1 cm. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p><span class="html-italic">SlALC</span> is involved in the regulation of fruit lignification. (<b>A</b>–<b>H</b>) qRT-PCR analysis of the expression levels of <span class="html-italic">PAL</span>, <span class="html-italic">C4H</span>, <span class="html-italic">LeCCR1</span>, <span class="html-italic">LeCCR2</span>, <span class="html-italic">4CL</span>, <span class="html-italic">CAD</span>, <span class="html-italic">SOD</span>, and <span class="html-italic">POD</span> in fruit wounds of WT and <span class="html-italic">SlALC</span>-OE. (<b>I</b>) Effector and reporter constructs used for dual-luciferase assay. (<b>J</b>) SlTAGL1 activates <span class="html-italic">SlALC</span> promoter by dual-luciferase assay. Data are means ± SD of three biological replicates. Statistically significant differences were determined using Student’s <span class="html-italic">t</span> test (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
13 pages, 3492 KiB  
Article
Rad6 Regulates Conidiation by Affecting the Biotin Metabolism in Beauveria bassiana
by Yuhan Guo, Haomin He, Yi Guan and Longbin Zhang
J. Fungi 2024, 10(9), 613; https://doi.org/10.3390/jof10090613 - 28 Aug 2024
Viewed by 331
Abstract
Rad6 is a canonical ubiquitin-conjugating enzyme known for its role in regulating chromosome-related cellular processes in yeast and has been proven to have multiple functions in Beauveria bassiana, including insect-pathogenic lifestyle, UV damage repair, and conidiation. However, previous studies have only reported [...] Read more.
Rad6 is a canonical ubiquitin-conjugating enzyme known for its role in regulating chromosome-related cellular processes in yeast and has been proven to have multiple functions in Beauveria bassiana, including insect-pathogenic lifestyle, UV damage repair, and conidiation. However, previous studies have only reported the key role of Rad6 in regulating conidial production in a nutrient-rich medium, without any deep mechanism analyses. In this study, we found that the disruption of Rad6 leads to a profound reduction in conidial production, irrespective of whether the fungus is cultivated in nutrient-rich or nutrient-poor environments. The absence of rad6 exerts a suppressive effect on the transcription of essential genes in the central developmental pathway, namely, brlA, abaA, and wetA, resulting in a direct downregulation of conidiation capacity. Additionally, mutant strains exhibited a more pronounced decline in both conidial generation and hyphal development when cultured in nutrient-rich conditions. This observation correlates with the downregulation of the central developmental pathway (CDP) downstream gene vosA and the upregulation of flaA in nutrient-rich cultures. Moreover, single-transcriptomics analyses indicated that irregularities in biotin metabolism, DNA repair, and tryptophan metabolism are the underlying factors contributing to the reduced conidial production. Comprehensive dual transcriptomics analyses pinpointed abnormal biotin metabolism as the primary cause of conidial production decline. Subsequently, we successfully restored conidial production in the Rad6 mutant strain through the supplementation of biotin, further confirming the transcriptomic evidence. Altogether, our findings underscore the pivotal role of Rad6 in influencing biotin metabolism, subsequently impacting the expression of CDP genes and ultimately shaping the asexual life cycle of B. bassiana. Full article
Show Figures

Figure 1

Figure 1
<p>Structural analysis of <span class="html-italic">B. bassiana</span> Rad6 and its phylogenetic relationship with other fungal homologues. (<b>a</b>) Domain analysis of <span class="html-italic">B. bassiana</span> Rad6. (<b>b</b>) The phylogenetic relationship between <span class="html-italic">B. basiana</span> Rad6 and other fungal homologues. Bootstrap values obtained through 1000 replications are presented at the nodes for reference. The branch length in the scale is proportional to the genetic distance, as determined using the neighbor-joining method within the MEGA7 program, accessible at <a href="http://www.megasoftware.net/" target="_blank">http://www.megasoftware.net/</a> (accessed on 8 October 2019) (note: the accession number of each protein in NCBI and its similarity to the corresponding protein of <span class="html-italic">Beauveria bassiana</span> Rad6 are shown in parentheses).</p>
Full article ">Figure 2
<p>Impacts of Rad6 disruption on conidiation capability and hyphal growth of <span class="html-italic">B. bassiana</span> in different culture media. (<b>a</b>) Conidia yield of each strain cultured in nutrient-rich SDAY plates. (<b>b</b>) Dry hyphal biomass of strains collected from nutrient-rich SDAY plates. (<b>c</b>) Conidia yield of each strain cultured in nutrient-poor CDA plates. (<b>d</b>) Dry hyphal biomass of strains collected from nutrient-poor CDA plates. All experiments were initiated by spreading 100 μL of a 10<sup>7</sup> conidia mL<sup>−1</sup> suspension. Asterisked bars in each group significantly differ from unmarked bars (Tukey’s HSD, <span class="html-italic">p</span> &lt; 0.05). Error bar: SD from three replicates.</p>
Full article ">Figure 3
<p>Transcriptome analysis of Δ<span class="html-italic">rad6</span> vs. WT collected from different culture media. (<b>a</b>,<b>b</b>) Distribution of log<sub>2</sub> FC and <span class="html-italic">p</span> values for genes identified in the transcriptomes from SDAY and CDA cultures. (<b>c</b>,<b>d</b>) KEGG pathway enrichment analysis of DEGs from SDAY and CDA cultures. Differentially expressed genes (DEGs) are defined as log<sub>2</sub> FC ≤ –1 or log<sub>2</sub> FC ≥ 1 at the level of <span class="html-italic">p</span> &lt; 0.05. The remaining genes are insignificantly affected (–1 ≤ log<sub>2</sub> FC ≤ 1). (Note: DnR refers to downregulated, UpR refers to upregulated, NDR refers to non-significant differentially regulated. The pathways are ordered based on the significant results in the KEGG enrichment analysis. <span class="html-italic">p</span>-value &lt; 0.001 labeled ***, <span class="html-italic">p</span>-value &lt; 0.01 labeled **, and <span class="html-italic">p</span>-value &lt; 0.05 labeled *).</p>
Full article ">Figure 4
<p>Analysis of genes co-occurring in the transcriptome of different media in <span class="html-italic">B. bassiana</span>. (<b>a</b>–<b>c</b>) Overlapped in the total, upregulated, and downregulated DEGs from SDAY and CDA cultures in <span class="html-italic">B. bassiana</span>. (<b>d</b>) KEGG pathway enrichment analysis of common DEGs from SDAY and CDA cultures. The pathways are ordered based on the significant results in the KEGG enrichment analysis. <span class="html-italic">p</span>-value &lt; 0.001 labeled ***, <span class="html-italic">p</span>-value &lt; 0.01 labeled **, and <span class="html-italic">p</span>-value &lt; 0.05 labeled *.</p>
Full article ">Figure 5
<p>Phenotype verification of biotin addition. (<b>a</b>,<b>c</b>) Conidiation of each strain in SDAY and CDA plates amended with different concentrations of biotin. (<b>b</b>,<b>d</b>) Relative conidiation inhibition rate of biotin addition at different concentrations. Asterisked bars in each group significantly differ from unmarked bars (Tukey’s HSD, <span class="html-italic">p</span> &lt; 0.05). Error bar: SD from three replicates.</p>
Full article ">Figure 6
<p>qRT-PCR verification with or without exogenous biotin. (<b>a</b>,<b>c</b>) Transcription levels of CDP-related genes in biotin-free SDAY and CDA plates. (<b>b</b>,<b>d</b>) Transcription levels of CDP-related genes in SDAY and CDA plates with biotin added. Asterisked bars in each group significantly differ from unmarked bars (Tukey’s HSD, <span class="html-italic">p</span> &lt; 0.05). Error bar: SD from three replicates.</p>
Full article ">
21 pages, 4352 KiB  
Article
PDCD10 Is a Key Player in TMZ-Resistance and Tumor Cell Regrowth: Insights into Its Underlying Mechanism in Glioblastoma Cells
by Yuan Zhu, Su Na Kim, Zhong-Rong Chen, Rainer Will, Rong-De Zhong, Philipp Dammann and Ulrich Sure
Cells 2024, 13(17), 1442; https://doi.org/10.3390/cells13171442 - 28 Aug 2024
Viewed by 393
Abstract
Overcoming temozolomide (TMZ)-resistance is a major challenge in glioblastoma therapy. Therefore, identifying the key molecular player in chemo-resistance becomes urgent. We previously reported the downregulation of PDCD10 in primary glioblastoma patients and its tumor suppressor-like function in glioblastoma cells. Here, we demonstrate that [...] Read more.
Overcoming temozolomide (TMZ)-resistance is a major challenge in glioblastoma therapy. Therefore, identifying the key molecular player in chemo-resistance becomes urgent. We previously reported the downregulation of PDCD10 in primary glioblastoma patients and its tumor suppressor-like function in glioblastoma cells. Here, we demonstrate that the loss of PDCD10 causes a significant TMZ-resistance during treatment and promotes a rapid regrowth of tumor cells after treatment. PDCD10 knockdown upregulated MGMT, a key enzyme mediating chemo-resistance in glioblastoma, accompanied by increased expression of DNA mismatch repair genes, and enabled tumor cells to evade TMZ-induced cell-cycle arrest. These findings were confirmed in independent models of PDCD10 overexpressing cells. Furthermore, PDCD10 downregulation led to the dedifferentiation of glioblastoma cells, as evidenced by increased clonogenic growth, the upregulation of glioblastoma stem cell (GSC) markers, and enhanced neurosphere formation capacity. GSCs derived from PDCD10 knockdown cells displayed stronger TMZ-resistance and regrowth potency, compared to their parental counterparts, indicating that PDCD10-induced stemness may independently contribute to tumor malignancy. These data provide evidence for a dual role of PDCD10 in tumor suppression by controlling both chemo-resistance and dedifferentiation, and highlight PDCD10 as a potential prognostic marker and target for combination therapy with TMZ in glioblastoma. Full article
(This article belongs to the Special Issue Molecular and Cellular Mechanisms of Cancers: Glioblastoma III)
Show Figures

Figure 1

Figure 1
<p>Knockdown of PDCD10 confers TMZ-resistance on GBM cells. (<b>A</b>) Confirmation of PDCD10 knockdown in lentiviral transduced U87 and T98g cells by RT<sup>2</sup>-PCR (<b>a</b>), western blot (<b>b</b>), and semi-quantitation of the blots (<b>c</b>). ev and sh: empty vector- and PDCD10 shRNA-transduced cells, respectively. IOD: integrated optical density. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, compared with ev. (<b>B</b>) Knockdown of PDCD10 in GBM cells leads to a resistance to TMZ-induced cell death. U87 and T98g cells received the treatment with 150 µM (<b>a</b>) and 300 µM (<b>b</b>) of TMZ, respectively, for 72 h. Thereafter, TMZ was washed-out. Remaining viable cells were cultured in the TMZ-free medium for 3 d, which is defined as the post-treatment phase. Control cells (C) were treated with vehicle DMSO (0.1% and 0.2% for U87 and T98g, respectively). MTT assay was performed to determine the viability of cells at 72 h after TMZ treatment (treatment phase) and 3 d after washing-out of TMZ (post-treatment phase). *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, compared with ev; ##, <span class="html-italic">p</span> &lt; 0.01; ###, <span class="html-italic">p</span> &lt; 0.0001, compared with evC in the same phase; +++, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding group in treatment phase.</p>
Full article ">Figure 2
<p>Knockdown of PDCD10 enhances cell viability after rechallenge with TMZ in regrown cells (RG) generated from the established acquired TMZ-resistant model. (<b>A</b>) Confirmation of PDCD10 knockdown in shU87-RG and shT98g-RG cells by RT<sup>2</sup>-PCR (<b>a</b>) and by FACS of respective transduced cells that expressed red-fluorescence protein (RFP) and green-fluorescence protein (GFP) (<b>b</b>). *, <span class="html-italic">p</span> &lt; 0.05, compared with ev. (<b>B</b>) MTT assay in RG cells in treatment phase and post-treatment phase. MTT assay was performed with ev/shU87 and ev/shT98g cells that received the treatment with 150 µM (<b>a</b>) and 300 µM (<b>b</b>) of TMZ, respectively, for 72 h (treatment phase) and 2 d and 4 d after washing-out TMZ (post-treatment phase, without reseeding). Control cells (C) received vehicle DMSO (0.1% and 0.2% for U87 and T98g, respectively). *, <span class="html-italic">p</span> &lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001, compared with evRG C (72 h); ###, <span class="html-italic">p</span> &lt; 0.001, compared with evRG-TMZ (72 h). (<b>C</b>) MTT assay in RG cells in a second post-treatment model with reseeding. ev/shU87-RG and ev/shT98g-RG cells received the treatment with 150 µM (<b>a</b>) and 300 µM (<b>b</b>) of TMZ, respectively, for 72 h. Thereafter, TMZ–containing media and dead cells were washed-out, and the viable cells were harvested and reseeded at the same density, followed by 2, 4, and 6 d of culture in drug-free medium. A significantly more rapid regrowth was observed in both TMZ-treated shU87-RG and shT98g-RG cells, compared with the corresponding evRG cells after reseeding and culturing in drug-free media. *, <span class="html-italic">p</span> &lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding evRG.</p>
Full article ">Figure 3
<p>Overexpression of PDCD10 sensitizes GBM cells to TMZ treatment. (<b>A</b>) Confirmation of overexpression of PDCD10 in lentiviral transduced U87 and T98g cells by RT<sup>2</sup>-PCR (<b>a</b>) and western blot (<b>b</b>) and semi-quantitation of the blots (<b>c</b>). Western blotting with anti-V5 antibody distinguishes between the expression of transgenic C-terminal V5-tagged PDCD10 protein and endogenous protein. ev and ox: empty vector-transduced and PDCD10-overexpressing cells, respectively. IOD: integrated optical density. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, compared with ev. (<b>B</b>) Overexpression of PDCD10 significantly reduces cell viability in a concentration-dependent manner after 72 h of TMZ treatment in both oxU87 (<b>a</b>) and oxT98g (<b>b</b>) cells. Control cells (C) were treated with vehicle DMSO (0.1% and 0.2% for U87 and T98g, respectively). *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding ev groups. (<b>C</b>) Overexpression of PDCD10 sensitizes GBM cells to TMZ treatment 72 h after TMZ treatment (treatment phase) and 3 d after washing-out TMZ (post-treatment phase). ev/oxU87 and ev/oxT98g cells received the treatment with 150 µM (<b>a</b>) and 300 µM (<b>b</b>) of TMZ for 72 h, respectively. Thereafter, TMZ-containing medium and dead cells were washed-out and the viable cells were further cultured in drug-free medium for 3 d followed by MTT assay. Control cells (C) were treated with vehicle DMSO (0.1% and 0.2% for U87 and T98g, respectively). *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding ev groups; +++, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding evC in the treatment phase; #, <span class="html-italic">p</span> &lt; 0.05, ###, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding evC in the same phase.</p>
Full article ">Figure 4
<p>DNA replication in response to TMZ treatment is dependent on PDCD10 expression. DNA replication was detected by EdU incorporation followed by FACS at 72 h of TMZ treatment (treatment phase) and at 3 d after TMZ-washing-out and -culturing in drug-washout media (post-treatment phase). ev/shT98g-RG and ev/oxT98g cells received 500 and 300 µM TMZ, respectively. (<b>A</b>,<b>C</b>) Histograms of EdU-positive (EdU+) and -negative (EdU−) cell populations in ev/shT98g-RG and ev/oxT98g cells, respectively. (<b>B</b>,<b>D</b>) Bar graphs of EdU+/− populations based on the corresponding histograms in (<b>A</b>,<b>C</b>). Knockdown of PDCD10 leads to an increase in DNA replication in both the treatment and post-treatment phases of T98g-RG cells, whereas overexpression of PDCD10 suppresses DNA replication in response to TMZ treatment. The data are representative of at least three independent experiments.</p>
Full article ">Figure 5
<p>Alteration in cell cycle checkpoints in response to TMZ treatment is dependent on PDCD10 expression. Cell cycle assay was performed by FACS after 72 h of TMZ treatment (treatment phase) and at 3 d after TMZ-washing-out and -culturing in drug-washout media (post-treatment phase). ev/shT98g-RG and ev/oxT98g cells received 500 and 300 µM TMZ, respectively. (<b>A</b>,<b>C</b>) are representative of cell cycle histograms in ev/shT98g-RG and ev/oxT98g cells, respectively. DNA content-based cell cycle distributions were defined using FlowJo with the Dean–Jett–Fox algorithm and presented in histograms. Each cell cycle phase is shown in different colors: G0/G1- (blue), S- (yellow), and G2/M-phase (green). (<b>B</b>,<b>D</b>) Stacked bar graphs of the distribution of the cell population in the cell cycle based on the corresponding histograms in (<b>A</b>,<b>C</b>). Knockdown of PDCD10 leads to the escape of cells from G2/M arrest and increases the population in the S phase (DNA replication phase) in both the treatment and post-treatment phases, whereas overexpression does the opposite. The data are representative of at least three independent experiments.</p>
Full article ">Figure 6
<p>Knockdown of PDCD10 in T98g-RG cells leads to deregulation of DNA damage response (DDR) genes. ev/shT98g-RG and ev/oxT98g cells received 300 µM of TMZ or vehicle (0.2% DMSO) treatment (no TMZ treatment). Cells were harvested for PCR detection of DDR genes after 72 h of TMZ treatment (treatment phase) and at 3 d after TMZ-washing-out and culturing in drug-free media (post-treatment phase). (<b>A</b>) Expression of MGMT in T98g-RG (<b>a</b>) cells and ev/oxT98g (<b>b</b>) cells in the no treatment, treatment (300 µM, 72 h), and post-treatment (3 d after washing out) phases. (<b>B</b>) Western blot (<b>a</b>) and semi-quantitation of the blots (<b>b</b>) of the MGMT protein expression in ev/shT98g-RG and ev/oxT98g cells. (<b>C</b>) Expression of DDR genes (<span class="html-italic">MSH2</span>, <span class="html-italic">MSH6,</span> and <span class="html-italic">PMS2</span>) in T98g-RG cells in the no treatment (<b>a</b>), treatment (<b>b</b>), and post-treatment phases (<b>c</b>), respectively. (<b>D</b>) Expression of DDR genes in ev/oxT98g cells in the no treatment (<b>a</b>), treatment (<b>b</b>), and post-treatment phases (<b>c</b>), respectively. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01 and ***, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding ev.</p>
Full article ">Figure 7
<p>PDCD10 expression determines the colony formation capacity of GBM cells. ev/shT98g-RG (<b>A</b>) and ev/oxT98g (<b>B</b>) cells received 300 µM of TMZ or vehicle (0.2% DMSO) treatment. Cells were harvested for colony formation assay in a 12-well plate in triplicate after 72 h of TMZ treatment (treatment phase) and at 3 d after TMZ-washing-out and culturing in drug-free media (post-treatment phase). The number of colonies was quantified after staining with 0.5% crystal violet using the ImageJ software (version 1.54j). Representative images of colony formation in ev/shT98g-RG and ev/oxT98g are shown in (<b>Aa</b>) and (<b>Ba</b>), respectively. Quantitative analysis of the colony numbers is presented in (<b>Ab</b>) and (<b>Bb</b>) for ev/shT98g-RG and ev/oxT98g cells, respectively. ***, <span class="html-italic">p</span> &lt; 0.001, compared with ev; ##, <span class="html-italic">p</span> &lt; 0.01 and ###, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding C; +++, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding group in the treatment phase.</p>
Full article ">Figure 8
<p>Knockdown of PDCD10 enhances the self-renewal capacity of U87-RG cells and GSCs generated from the parental cell line U87-RG (U87-RG-GSCs), and increases the expression of stem cell markers in RG-GSCs. (<b>A</b>) Representative images of neurospheres derived from U87-RG cells and U87-RG-GSCs. Scale bar: 100 µm. (<b>B</b>) Quantitative analysis of neurospheres formation efficiency (SFE). ***, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding ev; ###, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding parental cells. (<b>C</b>) Knockdown of PDCD10 increases the mRNA expression of stemness genes in RG-GSCs. The expression of stem cell markers <span class="html-italic">Nestin</span> and <span class="html-italic">KLF4</span> was detected by RT<sup>2</sup>-PCR in untreated U87-RG-GSCs, and in shU87-RG-GSCs treated with TMZ (150 µM) for 72 h. *, <span class="html-italic">p</span> &lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001, compared with corresponding ev; #, <span class="html-italic">p</span> &lt; 0.05; ###, <span class="html-italic">p</span> &lt; 0.001, compared with evC; +, <span class="html-italic">p</span> &lt; 0.05, compared with corresponding C. (<b>D</b>) Knockdown of PDCD10 enhances the viability of parental U87-RG cells and their GSC variants. U87-RG cells (left) and U87-RG-GSCs (right) received TMZ (150 M) or vehicle DMSO (0.1%) treatment for 72 h. Cell viability was detected after 72 h of 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, compared with corresponding ev; #, <span class="html-italic">p</span> &lt; 0.05; ##, <span class="html-italic">p</span> &lt; 0.01, compared with evC; +, <span class="html-italic">p</span> &lt; 0.05, ++, <span class="html-italic">p</span> &lt; 0.01, compared with corresponding parental cells. (<b>E</b>) Knockdown of PDCD10 reduces the mRNA expression of MMR genes (<span class="html-italic">MSH2</span>, <span class="html-italic">MSH6</span>, and <span class="html-italic">PMS2</span>) in U87-RG-GSCs. U87-RG-GSCs received treatment of 150 µM TMZ or vehicle (0.1% DMSO; no treatment). Cells were harvested for PCR detection of MMR genes after 72 h of treatment. Expression of MMR genes in non-treated (<b>Ea</b>) and TMZ-treated U87-RG-GSCs (<b>Eb</b>). *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001, compared with ev.</p>
Full article ">Figure 9
<p>Schematic summary of the role and mechanism of PDCD10 in acquired TMZ-resistance. Knockdown of PDCD10 (shPDCD10) in GBM cells significantly increased cell survival in response to TMZ treatment, and strongly promoted tumor cell regrowth in the post-treatment phase, which collectively accounted for acquired TMZ-resistance. Mechanism studies revealed that the loss of PDCD10 modulated the expression of DNA damage response genes (i.e., upregulating MGMT and downregulating MMR genes <span class="html-italic">MSH2</span>, <span class="html-italic">MSH6</span>, and <span class="html-italic">PMS2</span>), and altered the cell cycle process, as evidenced by the evasion of tumor cells from arrest at the G2/M phase, and the increase in tumor cells in the proliferating S phase. In addition, shPDCD10-GBM cells exhibited higher cell plasticity, as demonstrated by an increased capacity for colony formation and transformation of shPDCD10-GBM cells into GSC-like cells that expressed higher levels of the stem cell markers Nestin and KLF4. In support of these findings, overexpression of PDCD10 (oxPDCD10) induced contrary changes in the molecular and cell behaviors observed in shPDCD10-GBM cells, increasing the sensitivity of oxPDCD10-GBM cells to TMZ treatment and suppressing tumor cell regrowth after TMZ treatment. Our results indicate that PDCD10 plays a pivotal role in acquired TMZ-resistance and thus represents a promising target for perturbing TMZ-resistance and tumor recurrence.</p>
Full article ">
16 pages, 3649 KiB  
Article
Pan-Genome Analysis of TRM Gene Family and Their Expression Pattern under Abiotic and Biotic Stresses in Cucumber
by Lili Zhao, Ke Wang, Zimo Wang, Shunpeng Chu, Chunhua Chen, Lina Wang and Zhonghai Ren
Horticulturae 2024, 10(9), 908; https://doi.org/10.3390/horticulturae10090908 - 27 Aug 2024
Viewed by 377
Abstract
Cucumber (Cucumis sativus L.) is a vital economic vegetable crop, and the TONNEAU1 Recruiting Motif (TRM) gene plays a key role in cucumber organ growth. However, the pan-genomic characteristics of the TRM gene family and their expression patterns under different stresses have [...] Read more.
Cucumber (Cucumis sativus L.) is a vital economic vegetable crop, and the TONNEAU1 Recruiting Motif (TRM) gene plays a key role in cucumber organ growth. However, the pan-genomic characteristics of the TRM gene family and their expression patterns under different stresses have not been reported in cucumber. In this study, we identified 29 CsTRMs from the pan-genomes of 13 cucumber accessions, with CsTRM29 existing only in PI183967. Most CsTRM proteins exhibited differences in sequence length, except five CsTRMs having consistent protein sequence lengths among the 13 accessions. All CsTRM proteins showed amino acid variations. An analysis of CsTRM gene expression patterns revealed that six CsTRM genes strongly changed in short-fruited lines compared with long-fruited lines. And four CsTRM genes strongly responded to salt and heat stress, while CsTRM14 showed responses to salt stress, powdery mildew, gray mold, and downy mildew. Some CsTRM genes were induced or suppressed at different treatment timepoints, suggesting that cucumber TRM genes may play different roles in responses to different stresses, with expression patterns varying with stress changes. Remarkably, the expression of CsTRM21 showed considerable change between long and short fruits and in responses to abiotic stresses (salt stress and heat stress), as well as biotic stresses (powdery mildew and gray mold), suggesting a dual role of CsTRM21 in both fruit shape determination and stress resistance. Collectively, this study provided a base for the further functional identification of CsTRM genes in cucumber plant growth and stress resistance. Full article
(This article belongs to the Special Issue Vegetable Genomics and Breeding Research)
Show Figures

Figure 1

Figure 1
<p>Comparison of the conserved motifs and gene structures of <span class="html-italic">CsTRM07</span> (<b>A</b>), <span class="html-italic">CsTRM17</span> (<b>B</b>), and <span class="html-italic">CsTRM24</span> (<b>C</b>) in the 13 cucumber accessions.</p>
Full article ">Figure 2
<p>Synteny analysis of <span class="html-italic">TRMs</span> among cucumber and other plant species: Gray lines indicate the collinear blocks, while red lines highlight the collinear gene pairs involving TRM genes. <span class="html-italic">‘C. sativus’</span>, <span class="html-italic">‘Z. mays’</span>, <span class="html-italic">‘O. sativa’</span>, <span class="html-italic">‘A. thaliana’</span>, and <span class="html-italic">‘S. lycopersicum’</span> indicate <span class="html-italic">Cucumis sativus</span>, <span class="html-italic">Zea mays</span>, <span class="html-italic">Oryza sativa</span>, <span class="html-italic">Arabidopsis thaliana</span>, and <span class="html-italic">Solanum lycopersicum</span>, respectively.</p>
Full article ">Figure 3
<p>Expression analysis of <span class="html-italic">CsTRMs</span> in the fruit: The transcriptional levels of <span class="html-italic">CsTRM</span> genes in GFC (carpel number = 5) and 32X (carpel number = 3) (<b>A</b>), 408 (long fruit) and 409 (short fruit) (<b>B</b>), and WT and <span class="html-italic">CsFUL1<sup>A</sup></span>-OX (<b>C</b>) are shown on the heatmaps. A color scale range of −2.0 to 2.0 and −1.5 to 1.5 was applied, based on the normalized values. The color gradient, from blue to red, represents increasing expression levels. GFC, mutant Gui Fei Cui (GFC) from South China-type cucumber 32X. The carpel number changed from 3 in 32X to 5 in GFC, despite the number of other floral organs, such as sepal, petal, and stamen, remaining unchanged. WT, empty vector/control transgenic plants. FC, fold-change. (<b>D</b>) qRT-PCR analysis of <span class="html-italic">CsTRM</span> expression of the cucumber ovary at 4 days before anthesis (4 DBA) and 0 days after anthesis (0 DAA) at the long fruit CSSL2-7 and the round fruit RNS7. The gene of cucumber Actin served as reference gene. The standard error of the mean is represented by the error bars (<span class="html-italic">n</span> = 3). Significance analysis was performed with the two-tailed Student’s <span class="html-italic">t</span>-test (ns <span class="html-italic">p</span> &gt; 0.05, * <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).</p>
Full article ">Figure 4
<p>Expression patterns of <span class="html-italic">CsTRM</span> genes in response to abiotic stress: The heatmap displays the gene expression levels of <span class="html-italic">CsTRM</span> genes in response to salt (<b>A</b>) and heat (<b>B</b>) tolerance. A color scale range of –3.0 to 3.0 was applied, based on the normalized values. The color gradient, from blue to red, represents increasing expression levels. Abbreviations include CT for control treatment; HT for heat treatment; HT0h for HT at 0 h; HT3h for HT at 3 h; and HT6h for HT at 6 h. FC, fold-change.</p>
Full article ">Figure 5
<p>Expression analysis of <span class="html-italic">CsTRMs</span> under biotic stresses: The heatmaps displays the transcriptional levels of <span class="html-italic">CsTRM</span> genes in response to powdery mildew (PM) for 48 h (<b>A</b>), gray mold (GM) for 96 h (<b>B</b>), and downy mildew (DM) for 1–8 days post inoculation (<b>C</b>). A color scale range of –3.0 to 3.0 was applied, based on the normalized values. The color gradient, from blue to red, represents increasing expression levels. Abbreviations include ID for PM-inoculated susceptible cucumber line D8 leaves; NID for non-inoculated D8 leaves; IS for PM-inoculated resistant cucumber line SSL508-28 leaves; NIS for non-inoculated SSL508-28 leaves; CT for without inoculation; DPI for days post inoculation; and FC for fold-change. (<b>D</b>) qRT-PCR analysis of <span class="html-italic">CsTRM</span> expression of the cotyledons of cucumber seedlings inoculated with gray mold (GM) at 0 h, 6 h, 24 h, and 72 h, and maintaining environmental humidity after inoculation was necessary. The gene of cucumber Actin served as reference gene. The standard error of the mean is represented by the error bars (<span class="html-italic">n</span> = 3). Significance analysis was performed with the two-tailed Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
9 pages, 2006 KiB  
Article
Let-7f-5p Modulates Lipid Metabolism by Targeting Sterol Regulatory Element-Binding Protein 2 in Response to PRRSV Infection
by Dongfeng Jiang, Liyu Yang, Xiangge Meng, Qiuliang Xu, Xiang Zhou and Bang Liu
Vet. Sci. 2024, 11(9), 392; https://doi.org/10.3390/vetsci11090392 - 26 Aug 2024
Viewed by 572
Abstract
Porcine reproductive and respiratory syndrome (PRRS) has caused substantial damage to the pig industry. MicroRNAs (miRNAs) were found to play crucial roles in modulating the pathogenesis of PRRS virus (PRRSV). In the present study, we revealed that PRRSV induced let-7f-5p to influence lipid [...] Read more.
Porcine reproductive and respiratory syndrome (PRRS) has caused substantial damage to the pig industry. MicroRNAs (miRNAs) were found to play crucial roles in modulating the pathogenesis of PRRS virus (PRRSV). In the present study, we revealed that PRRSV induced let-7f-5p to influence lipid metabolism to regulate PRRSV pathogenesis. A transcriptome analysis of PRRSV-infected PK15CD163 cells transfected with let-7f-5p mimics or negative control (NC) generated 1718 differentially expressed genes, which were primarily associated with lipid metabolism processes. Furthermore, the master regulator of lipogenesis SREBP2 was found to be directly targeted by let-7f-5p using a dual-luciferase reporter system and Western blotting. The findings demonstrate that let-7f-5p modulates lipogenesis by targeting SREBP2, providing novel insights into miRNA-mediated PRRSV pathogenesis and offering a potential antiviral therapeutic target. Full article
(This article belongs to the Special Issue Genetic Diversity, Conservation, and Innovative Breeding in Pigs)
Show Figures

Figure 1

Figure 1
<p>Let-7f-5p restricts PRRSV replication and regulates triglyceride production. (<b>A</b>) The relative expression of let-7f-5p during PRRSV infection. (<b>B</b>) The effect of let-7f-5p mimics on PRRSV replication. (<b>C</b>) The effect of the let-7f-5p inhibitor on PRRSV replication. (<b>D</b>) The effect of the let-7f-5p mimics or inhibitor on triglyceride production. (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>Let7f-5p altered the transcriptomic profile in PRRSV-infected PK-15<sup>CD163</sup> cells. (<b>A</b>) The principal component analysis (PCA). (<b>B</b>) A volcano map of differentially expressed genes (DEGs). (<b>C</b>) A heatmap of DEGs related to lipid metabolism. (<b>D</b>) GO analysis. (<b>E</b>) KEGG pathway analysis.</p>
Full article ">Figure 3
<p>qRT-PCR analysis of expression of lipid metabolism genes in PRRSV-infected PK-15<sup>CD163</sup> cells transfected with let-7f-5p or NC (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Let-7f-5p targeted mRNA of <span class="html-italic">SREBP2</span>. (<b>A</b>) Bioinformatical predicted target sequences of let-7f-5p to 3′UTR of SREBP2 from multiple species. (<b>B</b>) Luciferase activity of SREBP2-3’UTR from PK-15 cells transfected with let-7f-5p mimic, NC or let-7f-5p inhibitor. (<b>C</b>) mRNA expression of SREBP2 from PK-15 cells transfected with let-7f-5p mimic or NC. (<b>D</b>) Protein expression of SREBP2 from PK-15 cells transfected with let-7f-5p mimic or NC. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
8 pages, 1481 KiB  
Brief Report
Honeysuckle-Derived miR2911 Inhibits Replication of Porcine Reproductive and Respiratory Syndrome Virus by Targeting Viral Gene Regions
by Xinyan Cao, Jiaxi Xue, Adnan Ali, Manyi Zhang, Jinliang Sheng, Yanming Sun and Yanbing Zhang
Viruses 2024, 16(9), 1350; https://doi.org/10.3390/v16091350 - 23 Aug 2024
Viewed by 387
Abstract
The highly abundant and stable antiviral small RNA derived from honeysuckle, known as miR2911, has been shown to play a key role in inhibiting influenza virus infection and SARS-CoV-2 infection. However, whether miR2911 inhibits the replication of porcine reproductive and respiratory syndrome virus [...] Read more.
The highly abundant and stable antiviral small RNA derived from honeysuckle, known as miR2911, has been shown to play a key role in inhibiting influenza virus infection and SARS-CoV-2 infection. However, whether miR2911 inhibits the replication of porcine reproductive and respiratory syndrome virus (PRRSV) remains unknown. Hence, this study investigated the mechanisms underlying the action of miR2911 during PRRSV infection. Six targets of miR2911 within the PRRSV orf1 (Nsp2: 2459 to 2477, 1871 to 1892, 954 to 977, and 1271 to 1292; Nsp1: 274 to 296 and 822 to 841) were successfully identified by using the miRanda v1.0b software. The miR2911 target sequence was analyzed by target sequence comparison, and only individual base mutations existed in different prevalent strains, and the miR2911 target region was highly conserved among different strains. Subsequently, through the dual luciferase reporter gene assay and miR2911 overexpression assay, it was demonstrated that miR2911 significantly inhibits the replication of PRRSV by targeting regions of PRRSV Nsp1 and Nsp2. These findings offer new insights for the development of novel anti-PRRSV drugs. Full article
(This article belongs to the Special Issue Porcine Viruses 2024)
Show Figures

Figure 1

Figure 1
<p>Prediction of miR2911 targeting PRRSV gene analysis (Note: “:” indicates base mismatch). Prediction of miR2911 targeting four regions of the PRRSV Nsp2 gene versus two regions of the Nsp1 gene using miRanda v1.0b software.</p>
Full article ">Figure 2
<p>Evolutionary tree analysis of miR2911 targeting viral regions, targeting PRRSV genes with multi-region sequence evolutionary trees. (<b>A</b>) miR2911 sequences downloaded from the NCBI database targeting the PRRSV gene region, with light colors indicating genes with differences. (<b>B</b>) Construction of an evolutionary tree of miR2911 targeting viral gene regions using the neighborhood method in MEGA 7.0 software.</p>
Full article ">Figure 3
<p>Overexpression of miR2911 inhibited replication of PRRSV via target viral gene regions. (<b>A</b>) HEK293T cells were co-transfected with luciferase reporter pGL3-promoter-PRRSV, pRL-TK, and miR2911 mimic or NC mimic; after transfection for 24 h, double luciferase reporter was detected. (<b>B</b>,<b>C</b>) miR2911 or NC was transfected into Marc-145 cells. After 24 h, the cells were infected with PRRSV at a MOI of 0.1 for 24 h. The mRNA and protein level of PRRSV orf7 was measured by qRT-PCR and Western blot, respectively. (<b>D</b>) The supernatants were collected at 24 h post-infection (hpi) for TCID50 assay. Data are presented as mean ± SEM pooled from one separated experiment; <span class="html-italic">n</span> ≥ 3 for each of the analyzed parameters. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
8 pages, 1940 KiB  
Article
Hsa-miR-874-3p Reduces Endogenous Expression of RGS4-1 Isoform In Vitro
by Feng-Ling Xu and Bao-Jie Wang
Genes 2024, 15(8), 1057; https://doi.org/10.3390/genes15081057 - 11 Aug 2024
Viewed by 584
Abstract
Background: The level of the regulator of G-protein signaling 4-1 (RGS4-1) isoform, the longest RGS4 isoform, is significantly reduced in the dorsolateral prefrontal cortex (DLPFC) of people with schizophrenia. However, the mechanism behind this has not been clarified. The 3′untranslated regions (3′UTRs) are [...] Read more.
Background: The level of the regulator of G-protein signaling 4-1 (RGS4-1) isoform, the longest RGS4 isoform, is significantly reduced in the dorsolateral prefrontal cortex (DLPFC) of people with schizophrenia. However, the mechanism behind this has not been clarified. The 3′untranslated regions (3′UTRs) are known to regulate the levels of their mRNA splice variants. Methods: We constructed recombinant pmir-GLO vectors with a truncated 3′ regulatory region of the RGS4 gene (3R1, 3R2, 3R3, 3R4, 3R5, and 3R6). The dual-luciferase reporter assay was conducted to find functional regions in HEK-293, SK-N-SH, and U87cells and then predicted miRNA binding to these regions. We performed a dual-luciferase reporter assay and a Western blot analysis after transiently transfecting the predicted miRNAs. Results: The dual-luciferase reporter assay found that regions +401–+789, +789–+1152, and +1562–+1990 (with the last base of the termination codon being +1) might be functional regions. Hsa-miR-874-3p, associated with many psychiatric disorders, might target the +789–+1152 region in the 3′UTR of the RGS4 gene. In the dual-luciferase reporter assay, the hsa-miR-874-3p mimic, co-transfected with 3R1, down-regulated the relative fluorescence intensities. However, this was reversed when the hsa-miR-874-3p mimic was co-transfected with m3R1 (deletion of +853–+859). The hsa-miR-874-3p mimic significantly decreased the endogenous expression of the RGS4-1 isoform in HEK-293 cells. Conclusions: Hsa-miR-874-3p inhibits the expression of the RGS4-1 isoform by targeting +853–+859. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The truncated 3′ regulatory region of the <span class="html-italic">RGS4</span> gene recombined into pmir-GLO vectors.</p>
Full article ">Figure 2
<p>The predicted miR-874-3p binding site in the RGS4 mRNA 3′-UTR and the deletion sequences in m3R1. The red bases are the deleted ones in m3R1.</p>
Full article ">Figure 3
<p>The relative fluorescence intensities of recombined pmir-GLO vectors (3R1–3R6) in HEK-293, U87, and SK-N-SH cells. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 4
<p>The relative fluorescence intensities of 3R1 and m3R1 co-transfected with hsa-miR-874-3p mimic or NC. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 5
<p>Western blot assay measuring the endogenous protein levels of the RGS4-1 inform after transfection of NC (1, 2, 3), the hsa-miR-874-3p mimic (4, 5, 6), an NC inhibitor (7, 8, 9), and an hsa-miR-874-3p inhibitor (10, 11, 12), in HEK-293 (<b>A</b>,<b>D</b>), SK (<b>B</b>,<b>E</b>), and U87 (<b>C</b>,<b>F</b>). ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001, (<b>A</b>–<b>C</b>) each panel summarizes data from two separate western blots.</p>
Full article ">
16 pages, 6271 KiB  
Article
MicroRNA-148a Targets DNMT1 and PPARGC1A to Regulate the Viability, Proliferation, and Milk Fat Synthesis of Ovine Mammary Epithelial Cells
by Jiqing Wang, Na Ke, Xinmiao Wu, Huimin Zhen, Jiang Hu, Xiu Liu, Shaobin Li, Fangfang Zhao, Mingna Li, Bingang Shi, Zhidong Zhao, Chunyan Ren and Zhiyun Hao
Int. J. Mol. Sci. 2024, 25(16), 8558; https://doi.org/10.3390/ijms25168558 - 6 Aug 2024
Viewed by 507
Abstract
In this study, the expression profiles of miR-148a were constructed in eight different ovine tissues, including mammary gland tissue, during six different developmental periods. The effect of miR-148a on the viability, proliferation, and milk fat synthesis of ovine mammary epithelial cells (OMECs) was [...] Read more.
In this study, the expression profiles of miR-148a were constructed in eight different ovine tissues, including mammary gland tissue, during six different developmental periods. The effect of miR-148a on the viability, proliferation, and milk fat synthesis of ovine mammary epithelial cells (OMECs) was investigated, and the target relationship of miR-148a with two predicted target genes was verified. The expression of miR-148a exhibited obvious tissue-specific and temporal-specific patterns. miR-148a was expressed in all eight ovine tissues investigated, with the highest expression level in mammary gland tissue (p < 0.05). Additionally, miR-148a was expressed in ovine mammary gland tissue during each of the six developmental periods studied, with its highest level at peak lactation (p < 0.05). The overexpression of miR-148a increased the viability of OMECs, the number and percentage of Edu-labeled positive OMECs, and the expression levels of two cell-proliferation marker genes. miR-148a also increased the percentage of OMECs in the S phase. In contrast, transfection with an miR-148a inhibitor produced the opposite effect compared to the miR-148a mimic. These results indicate that miR-148a promotes the viability and proliferation of OMECs in Small-tailed Han sheep. The miR-148a mimic increased the triglyceride content by 37.78% (p < 0.01) and the expression levels of three milk fat synthesis marker genes in OMECs. However, the miR-148a inhibitor reduced the triglyceride level by 87.11% (p < 0.01). These results suggest that miR-148a promotes milk fat synthesis in OMECs. The dual-luciferase reporter assay showed that miR-148a reduced the luciferase activities of DNA methyltransferase 1 (DNMT1) and peroxisome proliferator-activated receptor gamma coactivator 1-A (PPARGC1A) in wild-type vectors, suggesting that they are target genes of miR-148a. The expression of miR-148a was highly negatively correlated with PPARGC1A (r = −0.789, p < 0.001) in ovine mammary gland tissue, while it had a moderate negative correlation with DNMT1 (r = −0.515, p = 0.029). This is the first study to reveal the molecular mechanisms of miR-148a underlying the viability, proliferation, and milk fat synthesis of OMECs in sheep. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Expression levels of miR-148a in ovine eight different tissues (<b>A</b>), mammary gland tissue during different developmental periods (<b>B</b>) and mammary gland tissue at peak lactation of Small-tailed Han sheep and Gansu Alpine Merino sheep (<b>C</b>). Values with different lowercase letters above the bars are different (<span class="html-italic">p</span> &lt; 0.05). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Transfection efficiency of miR-148a detected using RT-qPCR (<b>A</b>) and its effect on the viability of ovine mammary epithelial cells (<b>B</b>). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Effect of miR-148a on the proliferation of ovine mammary epithelial cells (OMECs) when the miR-148a mimic and miR-148a inhibitor were transfected into OMECs. (<b>A</b>) Proliferation of OMECs detected using an Edu assay. (<b>B</b>) Percentage of Edu-labeled positive OMECs. (<b>C</b>) Relative expression levels of <span class="html-italic">CDK4</span> and <span class="html-italic">CDK2</span>. ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Effects of the miR-148a mimic (<b>A</b>) and miR-148a inhibitor (<b>C</b>) on the cycle of ovine mammary epithelial cells (OMECs) when compared to the miR-148a mimic NC (<b>B</b>) and miR-148a inhibitor NC (<b>D</b>).</p>
Full article ">Figure 5
<p>Effect of miR-148a on the triglyceride level (<b>A</b>) and expression levels of <span class="html-italic">mTOR</span> (<b>B</b>), <span class="html-italic">DGAT1</span> (<b>C</b>), and <span class="html-italic">ABCG2</span> (<b>D</b>) in ovine mammary epithelial cells (OMECs). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5 Cont.
<p>Effect of miR-148a on the triglyceride level (<b>A</b>) and expression levels of <span class="html-italic">mTOR</span> (<b>B</b>), <span class="html-italic">DGAT1</span> (<b>C</b>), and <span class="html-italic">ABCG2</span> (<b>D</b>) in ovine mammary epithelial cells (OMECs). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Construction and sequencing results of dual luciferase reporter vectors for two target genes. (<b>A</b>) The structural diagram of dual luciferase reporter vectors. (<b>B</b>) Sequence validation of the target gene <span class="html-italic">DNMT1</span> in wild-type (WT) and mutant-type (MUT) pmiR-RB-Report™ vectors by Sanger sequencing. (<b>C</b>) Sequence validation of the target gene <span class="html-italic">PPARGC1A</span> in WT and MUT pmiR-RB-Report™ vectors by Sanger sequencing.</p>
Full article ">Figure 6 Cont.
<p>Construction and sequencing results of dual luciferase reporter vectors for two target genes. (<b>A</b>) The structural diagram of dual luciferase reporter vectors. (<b>B</b>) Sequence validation of the target gene <span class="html-italic">DNMT1</span> in wild-type (WT) and mutant-type (MUT) pmiR-RB-Report™ vectors by Sanger sequencing. (<b>C</b>) Sequence validation of the target gene <span class="html-italic">PPARGC1A</span> in WT and MUT pmiR-RB-Report™ vectors by Sanger sequencing.</p>
Full article ">Figure 7
<p>Validation of miR-148a with the predicted target genes <span class="html-italic">DNMT1</span> and <span class="html-italic">PPARGC1A</span>. (<b>A</b>,<b>B</b>) The luciferase activities of the target genes <span class="html-italic">DNMT1</span> and <span class="html-italic">PPARGC1A</span> for miR-148a detected using a dual luciferase reporter assay. (<b>C</b>,<b>D</b>) Effect of miR-148a on the expression levels of the target genes <span class="html-italic">DNMT1</span> and <span class="html-italic">PPARGC1A</span>. ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Expression levels of the target genes <span class="html-italic">DNMT1</span> (<b>A</b>) and <span class="html-italic">PPARGC1A</span> (<b>B</b>) for miR-148a in ovine mammary gland tissue during six different developmental periods. Values with different lowercase letters above the bars are different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
18 pages, 11995 KiB  
Article
The Novel-m0230-3p miRNA Modulates the CSF1/CSF1R/Ras Pathway to Regulate the Cell Tight Junctions and Blood–Testis Barrier in Yak
by Qiu Yan, Qi Wang, Yong Zhang, Ligang Yuan, Junjie Hu and Xingxu Zhao
Cells 2024, 13(15), 1304; https://doi.org/10.3390/cells13151304 - 5 Aug 2024
Viewed by 672
Abstract
The yak (Bos grunniens) is a valuable livestock animal endemic to the Qinghai–Tibet Plateau in China with low reproductive rates. Cryptorchidism is one of the primary causes of infertility in male yaks. Compared with normal testes, the tight junctions (TJs) of Sertoli cells [...] Read more.
The yak (Bos grunniens) is a valuable livestock animal endemic to the Qinghai–Tibet Plateau in China with low reproductive rates. Cryptorchidism is one of the primary causes of infertility in male yaks. Compared with normal testes, the tight junctions (TJs) of Sertoli cells (SCs) and the integrity of the blood–testis barrier (BTB) in cryptorchidism are both disrupted. MicroRNAs are hairpin-derived RNAs of about 19–25 nucleotides in length and are involved in a variety of biological processes. Numerous studies have shown the involvement of microRNAs in the reproductive physiology of yak. In this study, we executed RNA sequencing (RNA-seq) to describe the expression profiles of mRNAs and microRNAs in yaks with normal testes and cryptorchidism to identify differentially expressed genes. GO and KEGG analyses were used to identify the biological processes and signaling pathways which the target genes of the differentially expressed microRNAs primarily engaged. It was found that novel-m0230-3p is an important miRNA that significantly differentiates between cryptorchidism and normal testes, and it is down-regulated in cryptorchidism with p < 0.05. Novel-m0230-3p and its target gene CSF1 both significantly contribute to the regulation of cell adhesion and tight junctions. The binding sites of novel-m0230-3p with CSF1 were validated by a dual luciferase reporter system. Then, mimics and inhibitors of novel-m0230-3p were transfected in vitro into SCs, respectively. A further analysis using qRT-PCR, immunofluorescence (IF), and Western blotting confirmed that the expression of cell adhesion and tight-junction-related proteins Occludin and ZO-1 both showed changes. Specifically, both the mRNA and protein expression levels of Occludin and ZO-1 in SCs decreased after transfection with the novel-m0230-3p mimics, while they increased after transfection with the inhibitors, with p < 0.05. These were achieved via the CSF1/CSF1R/Ras signaling pathway. In summary, our findings indicate a negative miRNA-mRNA regulatory network involving the CSF1/CSF1R/Ras signaling pathway in yak SCs. These results provide new insights into the molecular mechanisms of CSF1 and suggest that novel-m0230-3p and its target protein CSF1 could be used as potential therapeutic targets for yak cryptorchidism. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of appearance and pathology of normal testis and cryptorchidism in yak. (<b>A</b>,<b>B</b>) Testes weights, ** <span class="html-italic">p</span> &lt; 0.01. (<b>C</b>) H&amp;E staining, magnification of 20×. (<b>D</b>) mRNA expression of β-catenin in normal testis and cryptorchidism of yak. Values represent mean ± SD; <span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> &lt; 0.01. (<b>E</b>) mRNA expression of ZO-1 in normal testis and cryptorchidism of yak. Values represent mean ± SD; <span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> &lt; 0.01. (<b>F</b>) Protein expression of β-catenin and ZO-1 in normal and cryptorchid testes of yaks. (<b>G</b>) Localization of β-catenin and ZO-1 proteins in yak testes, analyzed by immunofluorescence staining. β-catenin and ZO-1 in tissue are shown separately in red and nuclei are colored blue; magnification, 20×. Tes: normal testis; Cry: cryptorchidism; S: spermatogonium; SZ: spermatozoa; ST: seminiferous tubule; PS: primary spermatocyte; SC: Sertoli cell; LC: Leydig cell, PMC: peritubular myoid cell; CV: capillary vessel.</p>
Full article ">Figure 2
<p>Functional annotation and enrichment analyses in normal testis and cryptorchidism. (<b>A</b>) Cluster analysis of six samples. (<b>B</b>) Clustering heatmap of differential mRNA expression. (<b>C</b>) KEGG pathway analysis of source genes of differential mRNAs. (<b>D</b>) Number of DE mRNAs, with red indicating upregulation and green indicating downregulation. (<b>E</b>) GO enrichment analysis of DE mRNAs. Tes: normal testis, Cry: cryptorchidism.</p>
Full article ">Figure 3
<p>miRNA–mRNA interaction network. (<b>A</b>) Number of DE miRNAs, with purple indicating known and blue indicating novel. (<b>B</b>) Expression levels of DE miRNAs. (<b>C</b>) Clustering heatmap of DE miRNA expression. Red corresponds to upregulation, and green corresponds to downregulation. (<b>D</b>) Network plot of novel-m0230-3p and its target gene <span class="html-italic">CSF1</span>. Tes: normal testis, Cry: cryptorchidism.</p>
Full article ">Figure 4
<p>qRT-PCR verification of DE genes in normal testis and cryptorchidism in yak. (<b>A</b>) mRNA expression of DE genes shown by qRT-PCR analysis. (<b>B</b>) Comparison of |log2 FC| expression levels of mRNA-seq and qRT-PCR. (<b>C</b>) mRNA expression of DE miRNAs shown by qRT-PCR analysis. (<b>D</b>) Comparison of |log2 FC| expression levels of miRNA-seq and qRT-PCR. Values represent mean ± SD; <span class="html-italic">n</span> = 3. ** <span class="html-italic">p</span> &lt; 0.01. FC, fold-change. Tes: normal testis, Cry: cryptorchidism.</p>
Full article ">Figure 5
<p>Verification of CSF1 and CSF1R. (<b>A</b>) Expression patterns of CSF1 and CSF1R proteins by Western blot analysis; <span class="html-italic">n</span> = 3. (<b>B</b>–<b>E</b>) Immunofluorescence assay for expression and location of CSF1 and CSF1R in testis and cryptorchidism; magnification, 20×. Tes: normal testis; Cry: cryptorchidism; S: spermatogonium; SZ: spermatozoa; ST: seminiferous tubule; PS: primary spermatocyte; SC: Sertoli cell; LC: Leydig cell, CV: capillary vessel.</p>
Full article ">Figure 6
<p>Targeting relationship between target gene CSF1 and novel-m0230-3p. (<b>A</b>) IF staining identified isolated yak SCs using antibodies against SOX9 (green) and β-tubulin (red); magnification, 20×. (<b>B</b>) Immunofluorescence staining identified isolated yak SCs using antibodies against WT1 (green) and β-tubulin (red); magnification, 20×. (<b>C</b>) Binding site of CSF1 and novel-m0230-3p. (<b>D</b>) Luciferase activity in SCs after co-transfection with mimics of novel-m0230-3p (100 nM) or mimic NC (100 nM) and pmirGLO-CSF1 3′-UTR-WT (400 ng) or pmirGLO-CSF1 3′-UTR-MUT (400 ng). Values represent mean ± SD; <span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Regulation of CSF1 by novel-m0320-3p. (<b>A</b>) Optimal transfection efficiency of mimics and inhibitors of novel-m0230-3p was explored at different concentrations and transfection times. Values represent mean ± SD; <span class="html-italic">n</span> = 3. <span class="html-italic">* p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (<b>B</b>) Localization of CSF1 protein and novel-m0230-3p mimic in SCs was analyzed by immunofluorescence staining. CSF1 was colored red, mimic is shown in green, and nuclei were counterstained with DAPI (blue); magnification, 20×. (<b>C</b>) Localization of CSF1 protein and novel-m0230-3p inhibitor in SCs was analyzed by immunofluorescence staining. CSF1 was colored red, inhibitor is shown in green, and nuclei were counterstained with DAPI (blue); magnification, 20×. (<b>D</b>) mRNA expression of novel-m0230-3p after transfection of 100 nM mimic into SCs for 48 h. Values represent mean ± SD; <span class="html-italic">n</span> = 3. ** <span class="html-italic">p &lt;</span> 0.01. (<b>E</b>,<b>F</b>) mRNA and protein expression of CSF1 after transfection of 100 nM mimic into SCs for 48 h. ** <span class="html-italic">p</span> &lt; 0.01. (<b>G</b>) mRNA expression of novel-m0230-3p after transfection of 100 nM inhibitor into Sertoli cells for 48 h. Values represent mean ± SD; <span class="html-italic">n</span> = 3. ** <span class="html-italic">p</span> &lt; 0.01. (<b>H</b>,<b>I</b>) mRNA and protein expression of CSF1 after transfection of 100 nM inhibitor into SCs for 48 h. Values represent mean ± SD; <span class="html-italic">n</span> = 3. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Novel-m0230-3p regulates CSF1 expression via CSF1/CSF1R/Ras signaling pathway to affect cellular TJs and adhesion. (<b>A</b>) mRNA expression of CSF1R, Ras, Occludin, and ZO-1 was measured by qRT-PCR after transfection of 100 nM mimic into SCs for 48 h. Values represent mean ± SD; <span class="html-italic">n</span> = 3. ** <span class="html-italic">p</span> &lt; 0.01. (<b>B</b>) Protein expression of CSF1R, Ras, Occludin, and ZO-1 was assessed by Western blotting after transfection of 100 nM mimic into SCs for 48 h. (<b>C</b>) mRNA expression of CSF1R, Ras, Occludin, and ZO-1 was measured by qRT-PCR after transfection of 100 nM inhibitor into SCs for 48 h. Values represent mean ± SD; <span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> &lt; 0.01. (<b>D</b>) Protein expression of CSF1R, Ras, Occludin, and ZO-1 was assessed by Western blotting after transfection of 100 nM inhibitor into SCs for 48 h.</p>
Full article ">Figure 9
<p>Regulatory effects of novel-m0230-3p on cell TJs and adhesion via CSF1/CSF1R/Ras signaling pathway in yak SCs. Abbreviations: CSF1: colony-stimulating factor 1; CSF1R: colony-stimulating factor 1 receptor; ZO-1: TJP1 (ZO1) tight junction protein 1.</p>
Full article ">
15 pages, 8438 KiB  
Article
Role of Csdc2 in Regulating Secondary Hair Follicle Growth in Cashmere Goats
by Heqing Zhu, Yingying Li, He Xu, Yuehui Ma, Göran Andersson, Erik Bongcam-Rudloff, Tiantian Li, Jie Zhang, Yan Li, Jilong Han and Min Yang
Int. J. Mol. Sci. 2024, 25(15), 8349; https://doi.org/10.3390/ijms25158349 - 30 Jul 2024
Viewed by 759
Abstract
Cashmere goats possess two types of hair follicles, with the secondary hair follicles producing valuable cashmere fiber used for textiles. The growth of cashmere exhibits a seasonal pattern arising from photoperiod change. Transcription factors play crucial roles during this process. The transcription factor, [...] Read more.
Cashmere goats possess two types of hair follicles, with the secondary hair follicles producing valuable cashmere fiber used for textiles. The growth of cashmere exhibits a seasonal pattern arising from photoperiod change. Transcription factors play crucial roles during this process. The transcription factor, cold-shock domain, containing C2 (Csdc2) plays a crucial role in modulating cell proliferation and differentiation. Our preceding research indicated that the expression of Csdc2 changes periodically during anagen to telogen. However, the mechanisms of Csdc2 in regulating SHF growth remain unclear. Here, we found that the knockdown of Csdc2 inhibits the proliferation of dermal papilla cells. ChIP-Seq analysis showed that Csdc2 had a unique DNA binding motif in SHFs. Through conjoint analysis of ChIP-Seq and RNA-Seq, we revealed a total of 25 candidate target genes of Csdc2. Notably, we discovered a putative Csdc2 binding site within roundabout guidance receptor 2 (Robo2) on chromosome 1 of the goat genome. Furthermore, qRT-PCR and dual-luciferase reporter assay confirmed Csdc2’s positive regulatory influence on Robo2. These findings expand the research field of hair follicle transcriptional regulatory networks, offering insights into molecular breeding strategies to enhance cashmere production in goats. Full article
(This article belongs to the Special Issue Molecular Progression of Genetics in Breeding of Farm Animals)
Show Figures

Figure 1

Figure 1
<p>Immunohistochemistry and two types of HFs under a stereomicroscope. (<b>A</b>) Immunohistochemical results show Csdc2 is highly expressed at DP in anagen (Sep). Scale bars indicate 100 µm. (<b>B</b>) PHFs and SHFs.</p>
Full article ">Figure 2
<p>Culture and characterization of DPCs. (<b>A</b>) Different states of DPCs at different days of culture. Scale bars indicate 100 µm. (<b>B</b>) Using immunofluorescence showing DPCs were positive for vimentin and α-SMA.</p>
Full article ">Figure 3
<p>Alterations of hair follicle development-related genes and cell proliferation after Csdc2 knockdown. (<b>A</b>) Knockdown of Csdc2 significantly inhibited the expression of HF-development-related genes. (<b>B</b>) Cell proliferation after Csdc2 knockdown. *** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Analysis results of ChIP. (<b>A</b>) Average length of annotated peaks. (<b>B</b>) Regional statistics of peak annotation. (<b>C</b>) GO cluster of genes in coding and non-coding regions. (<b>D</b>) KEGG analysis of genes in coding and non-coding regions.</p>
Full article ">Figure 5
<p>Specific binding sites analysis. (<b>A</b>) Sequence map of motif CST6. (<b>B</b>) KEGG enrichment and GO clustering of genes in this motif. (<b>C</b>) Venn map shows the relationship between ChIP data, RNA-Seq data, and motif corresponding genes.</p>
Full article ">Figure 6
<p>Analyses of sequence conservation and visualisation of binding sites. (<b>A</b>) Conservation of motifs. (<b>B</b>) Enrichment of Csdc2 in the intron 6 of <span class="html-italic">Robo2</span>.</p>
Full article ">Figure 7
<p>Knockdown, overexpression, and dual luciferase reporter assay results. (<b>A</b>) Overexpression Csdc2 promoted the expression of Robo2. (<b>B</b>) Knockdown of Csdc2 significantly inhibited the expression of Robo2. (<b>C</b>) The results of the luciferase reporter assays conducted in 293T cells indicate a targeting relationship between Robo2 and Csdc2 (mean ± SD). *, <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.</p>
Full article ">
16 pages, 3076 KiB  
Article
MicroRNA-181a Targets GNAI2 and Affects the Proliferation and Induction Ability of Dermal Papilla Cells: The Potential Involvement of the Wnt/β-Catenin Signaling Pathway
by Mingliang He, Xiaoyang Lv, Joram M. Mwacharo, Yutao Li, Shanhe Wang and Wei Sun
Int. J. Mol. Sci. 2024, 25(14), 7950; https://doi.org/10.3390/ijms25147950 - 20 Jul 2024
Viewed by 800
Abstract
Wool is generated by hair follicles (HFs), which are crucial in defining the length, diameter, and morphology of wool fibers. However, the regulatory mechanism of HF growth and development remains largely unknown. Dermal papilla cells (DPCs) are a specialized cell type within HFs [...] Read more.
Wool is generated by hair follicles (HFs), which are crucial in defining the length, diameter, and morphology of wool fibers. However, the regulatory mechanism of HF growth and development remains largely unknown. Dermal papilla cells (DPCs) are a specialized cell type within HFs that play a crucial role in governing the growth and development of HFs. This study aims to investigate the proliferation and induction ability of ovine DPCs to enhance our understanding of the potential regulatory mechanisms underlying ovine HF growth and development. Previous research has demonstrated that microRNA-181a (miR-181a) was differentially expressed in skin tissues with different wool phenotypes, which indicated that miR-181a might play a crucial role in wool morphogenesis. In this study, we revealed that miR-181a inhibited the proliferation and induction ability of ovine DPCs by quantitative Real-time PCR (qRT-PCR), cell counting Kit-8 (CCK-8), 5-ethynyl-2′-deoxyuridine (EdU), flow cytometry, and alkaline phosphatase staining. Then, we also confirmed G protein subunit alpha i2 (GNAI2) is a target gene of miR-181a by dual luciferase reporter assay, qRT-PCR, and Western blot, and that it could promote the proliferation and induction ability of ovine DPCs. In addition, GNAI2 could also activate the Wnt/β-Catenin signaling pathway in ovine DPCs. This study showed that miR-181a can inhibit the proliferation and induction ability of ovine DPCs by targeting GNAI2 through the Wnt/β-Catenin signaling pathway. Full article
Show Figures

Figure 1

Figure 1
<p>MiR-181a inhibits the proliferation of ovine DPCs. (<b>a</b>,<b>b</b>) Expression of miR-181a after transfection of ovine DPCs with miR-181a mimic and miR-181a inhibitor. (<b>c</b>,<b>d</b>) Expression of <span class="html-italic">PCNA</span> and <span class="html-italic">CDK2</span> after transfection of sheep DPCs with miR-181a mimic and miR-181a inhibitor, respectively. (<b>e</b>,<b>f</b>) CCK-8 assay after transfection of ovine DPCs with miR-181a mimic and miR-181a inhibitor, respectively. (<b>g</b>,<b>h</b>) Cell cycle assay after transfection of sheep DPCs with miR-181a mimic or miR-181a inhibitor. (<b>i</b>,<b>j</b>) EdU assay after transfection of sheep DPCs with miR-181a mimic or miR-181a inhibitor; the scale is 100 µm. The unpaired Student’s <span class="html-italic">t</span>-test was used for statistical significance (<sup>ns</sup> <span class="html-italic">p</span> &gt; 0.05; * <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).</p>
Full article ">Figure 2
<p>MiR-181a inhibits the induction ability of ovine DPCs. (<b>a</b>,<b>b</b>) Alkaline phosphatase staining after transfection of ovine DPCs with miR-181a mimic or miR-181a inhibitor; the scale is 250 µm. (<b>c</b>,<b>d</b>) Expression of <span class="html-italic">FGF7</span>, <span class="html-italic">IGF1</span>, and <span class="html-italic">Versican</span> after transfection of ovine DPCs with miR-181a mimic and miR-181a inhibitor. The unpaired Student’s <span class="html-italic">t</span>-test was used for statistical significance (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p><span class="html-italic">GNAI2</span> is a target gene of miR-181a. (<b>a</b>) The potential binding site of the miR-181a seed sequence in the 3′UTR region of the <span class="html-italic">GNAI2</span> gene. The seed sequences of miR-181a are highlighted in red, and the wild-type and mutant sequences of the 3′UTR region of the <span class="html-italic">GNAI2</span> gene are highlighted in green. (<b>b</b>) The interaction model between miR-181a and the 3′UTR of <span class="html-italic">GNAI2</span> was analyzed by RNAhybrid. (<b>c</b>) <span class="html-italic">GNAI2</span> is expressed in Hu sheep DPCs; the scale is 50 µm. (<b>d</b>) The luciferase assays after transfection of the vectors PMIR-GNAI2-3′UTR-WT or PMIR-GNAI2-3′UTR-MT and miR-181a mimic or mimic-NC into HEK293T cells. (<b>e</b>,<b>f</b>) The mRNA expression level of <span class="html-italic">GNAI2</span> after overexpression or knockdown of miR-181a in ovine DPCs. (<b>g</b>,<b>h</b>) The protein expression of <span class="html-italic">GNAI2</span> after overexpression or knockdown of miR-181a in ovine DPCs. The unpaired Student’s <span class="html-italic">t</span>-test was used for statistical significance (<sup>ns</sup> <span class="html-italic">p</span> &gt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p><span class="html-italic">GNAI2</span> promotes the proliferation of ovine DPCs. (<b>a</b>,<b>b</b>) Expression of <span class="html-italic">GNAI2</span> after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. (<b>c</b>,<b>d</b>) Expression of <span class="html-italic">PCNA</span> and <span class="html-italic">CDK2</span> after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. (<b>e</b>,<b>f</b>) CCK-8 assay after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. (<b>g</b>,<b>h</b>) Cell cycle assay after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. (<b>i</b>,<b>j</b>) EdU assay after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs; the scale is 100 µm. The unpaired Student’s <span class="html-italic">t</span>-test was used for statistical significance (<sup>ns</sup> <span class="html-italic">p</span> &gt; 0.05; * <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).</p>
Full article ">Figure 5
<p><span class="html-italic">GNAI2</span> promotes the induction ability of ovine DPCs. (<b>a</b>,<b>b</b>) Alkaline phosphatase staining after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs; the scale is 250 µm. (<b>c</b>,<b>d</b>) Expression of <span class="html-italic">FGF7</span>, <span class="html-italic">IGF1</span>, and <span class="html-italic">Versican</span> after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. The unpaired Student’s <span class="html-italic">t</span>-test was used for statistical significance (<sup>ns</sup> <span class="html-italic">p</span> &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p><span class="html-italic">GNAI2</span> is involved in the Wnt/β-catenin signaling pathway of ovine DPCs. (<b>a</b>,<b>b</b>) TOP/FOP flash assays after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. (<b>c</b>,<b>d</b>) Expression of <span class="html-italic">CTNNB1</span>, <span class="html-italic">TCF4</span>, <span class="html-italic">LEF1</span>, <span class="html-italic">c-MYC</span>, and <span class="html-italic">cyclinD1</span> after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. (<b>e</b>,<b>f</b>) Expression of β-catenin after overexpression or knockdown of <span class="html-italic">GNAI2</span> in ovine DPCs. The unpaired Student’s <span class="html-italic">t</span>-test was used for statistical significance (<sup>ns</sup> <span class="html-italic">p</span> &gt; 0.05; * <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).</p>
Full article ">Figure 7
<p>Schematic of miR-181a regulating the proliferation and induction ability of DPCs by GNAI2-Wnt/β-catenin signaling-pathway axis.</p>
Full article ">
Back to TopTop