[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (512)

Search Parameters:
Keywords = siRNA transfection

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 3254 KiB  
Article
Sestrin2 Suppression Promotes Endothelial–Mesenchymal Transition and Exacerbates Methylglyoxal-Induced Endothelial Dysfunction
by Shahenda Salah Abdelsalam, Muhammad Ammar Zahid, Sarah Khalaf Ghanem, Abbas Khan, Aijaz Parray and Abdelali Agouni
Int. J. Mol. Sci. 2024, 25(24), 13463; https://doi.org/10.3390/ijms252413463 - 16 Dec 2024
Viewed by 243
Abstract
Sestrin2 (SESN2) is a stress-inducible protein known for its cytoprotective functions, but its role in diabetic vascular complications remains unclear. This study investigated the impact of SESN2 on methylglyoxal (MGO)-induced endothelial–mesenchymal transition (EndMT). Human endothelial cells were transfected with SESN2 siRNA duplexes to [...] Read more.
Sestrin2 (SESN2) is a stress-inducible protein known for its cytoprotective functions, but its role in diabetic vascular complications remains unclear. This study investigated the impact of SESN2 on methylglyoxal (MGO)-induced endothelial–mesenchymal transition (EndMT). Human endothelial cells were transfected with SESN2 siRNA duplexes to silence SESN2 expression, followed by MGO treatment. SESN2 knockdown significantly exacerbated MGO-induced oxidative stress, as evidenced by the reduced expression of antioxidant markers. Furthermore, SESN2 silencing enhanced the inflammatory response to MGO, demonstrated by the increased levels of pro-inflammatory cytokines. Notably, SESN2 deficiency promoted EndMT, a key process in diabetes-induced cardiovascular complications, as shown by the increased expression of mesenchymal markers and the decreased expression of endothelial markers. These findings suggest that SESN2 plays a critical protective role in endothelial cells against MGO-induced damage. The study provides novel insights into the molecular mechanisms underlying diabetic cardiovascular complications and identifies SESN2 as a potential therapeutic target for preventing endothelial dysfunction in diabetes. Our results indicate that SESN2 downregulation may contribute to the pathogenesis of diabetic vascular complications by promoting EndMT, increased oxidative stress, and inflammation. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Obesity and Metabolic Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of MGO treatment on the expression of SESN2 in endothelial cells. Western bot analysis and densitometry data of SESN2 protein expression normalized against loading control GAPDH and expressed as a percentage (%) of the untreated group (CTRL) (n = 4 in each group). Cells were left untreated or incubated with SESN2 siRNA duplexes for 48 h and then either exposed or not exposed to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. CTRL or indicated groups.</p>
Full article ">Figure 2
<p>Effect of <span class="html-italic">SESN2</span> silencing on the mRNA expression levels of endothelial and mesenchymal markers in endothelial cells subjected to MGO. Relative mRNA expression levels of endothelial markers: <span class="html-italic">VE-Cadherin</span> (<b>A</b>), <span class="html-italic">PECAM</span> (<b>B</b>), <span class="html-italic">TIE1</span> (<b>C</b>), <span class="html-italic">TIE2</span> (<b>D</b>), <span class="html-italic">vWF</span> (<b>E</b>), and mesenchymal markers <span class="html-italic">α-SMA</span> (<b>F</b>), <span class="html-italic">Vimentin</span> (<b>G</b>), <span class="html-italic">SM22</span> (<b>H</b>) and <span class="html-italic">FSP-1</span> (<b>I</b>) normalized against housekeeping gene <span class="html-italic">β-actin</span> (n = 6 in each group). Cells were left untreated or incubated with <span class="html-italic">SESN2</span> siRNA duplexes for 48 h, and then exposed or not to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, vs. CTRL or indicated groups.</p>
Full article ">Figure 3
<p>Effect of <span class="html-italic">SESN2</span> silencing on the protein expression of endothelial and mesenchymal markers in endothelial cells subjected to MGO. (<b>A</b>) Western bot analysis of endothelial markers: VE-Cadherin, PECAM, and Tie2. (<b>B</b>–<b>D</b>) Densitometry data of protein expression of VE-Cadherin (<b>B</b>), PECAM (<b>C</b>), and Tie2 (<b>D</b>) normalized against loading control GAPDH and expressed as a percentage (%) of the untreated group (CTRL). (<b>E</b>) Western bot analysis of mesenchymal markers: α-SMA, Vimentin, and TGF-β. (<b>F</b>–<b>H</b>) Densitometry data of protein expression of α-SMA (<b>F</b>), Vimentin (<b>G</b>), and TGF-β (<b>H</b>) normalized against loading control GAPDH and expressed as a percentage (%) of the untreated group (CTRL). Cells were left untreated or incubated with <span class="html-italic">SESN2</span> siRNA duplexes for 48 h, and then exposed to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M (n = 4 in each group). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, vs. CTRL or indicated groups.</p>
Full article ">Figure 4
<p>Effect of <span class="html-italic">SESN2</span> silencing on the expression of oxidative stress markers in endothelial cells subjected to MGO. Relative mRNA expression levels of <span class="html-italic">KEAP-1</span> (<b>A</b>), <span class="html-italic">NRF-2</span> (<b>B</b>), <span class="html-italic">NOX-2</span> (<b>C</b>), <span class="html-italic">PGD</span> (<b>D</b>), and <span class="html-italic">NQO-1</span> (<b>E</b>) normalized against housekeeping gene <span class="html-italic">β-actin</span> (n = 6 in each group). Cells were left untreated or incubated with <span class="html-italic">SESN2</span> siRNA duplexes for 48 h, and then exposed to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, vs. CTRL or indicated groups.</p>
Full article ">Figure 5
<p>Effect of SESN2 silencing on the expression of inflammatory markers in endothelial cells subjected to MGO. Relative mRNA expression levels of <span class="html-italic">IL-6</span> (<b>A</b>), <span class="html-italic">IL-8</span> (<b>B</b>), <span class="html-italic">IL-10</span> (<b>C</b>), <span class="html-italic">IL-1β</span> (<b>D</b>), <span class="html-italic">TNF-α</span> (<b>E</b>), <span class="html-italic">NF-κB</span> (<b>F</b>), <span class="html-italic">NLRP3</span> (<b>G</b>), and <span class="html-italic">COX-2</span> (<b>H</b>) normalized against housekeeping gene <span class="html-italic">β-actin</span> (n = 6 in each group). Cells were left untreated or incubated with <span class="html-italic">SESN2</span> siRNA duplexes for 48 h, and then exposed or not to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, vs. CTRL or indicated groups.</p>
Full article ">Figure 6
<p>Impact of SESN2 silencing on the expression levels of adhesion molecules in endothelial cells subjected to MGO. Relative mRNA expression levels of <span class="html-italic">ICAM-1</span> (<b>A</b>), <span class="html-italic">MCP-1</span> (<b>B</b>), and <span class="html-italic">E-selectin</span> (<b>C</b>) normalized against housekeeping gene <span class="html-italic">β-actin</span> (n = 6 in each group). Cells were left untreated or incubated with <span class="html-italic">SESN2</span> siRNA duplexes for 48 h, and then exposed to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, vs. CTRL or indicated groups.</p>
Full article ">Figure 7
<p>Impact of SESN2 silencing on the expression levels of TGF-β signaling and transcription factors in endothelial cells subjected to MGO. Relative mRNA expression levels of <span class="html-italic">TGF-β</span> (<b>A</b>), <span class="html-italic">SMAD2</span> (<b>B</b>), <span class="html-italic">SMAD3</span> (<b>C</b>), <span class="html-italic">SMAD4</span> (<b>D</b>), <span class="html-italic">β-Catenin</span> (<b>E</b>), <span class="html-italic">COL1A1</span> (<b>F</b>), and <span class="html-italic">SNAI1</span> (<b>G</b>) normalized against housekeeping gene <span class="html-italic">β-actin</span> (n = 6 in each group). Cells were left untreated or incubated with <span class="html-italic">SESN2</span> siRNA duplexes for 48 h, and then exposed or not exposed to MGO (600 μM for 18 h). Data are presented as mean ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, vs. CTRL or indicated groups.</p>
Full article ">
22 pages, 16663 KiB  
Article
Gene-Silencing Therapeutic Approaches Targeting PI3K/Akt/mTOR Signaling in Degenerative Intervertebral Disk Cells: An In Vitro Comparative Study Between RNA Interference and CRISPR–Cas9
by Masao Ryu, Takashi Yurube, Yoshiki Takeoka, Yutaro Kanda, Takeru Tsujimoto, Kunihiko Miyazaki, Hiroki Ohnishi, Tomoya Matsuo, Naotoshi Kumagai, Kohei Kuroshima, Yoshiaki Hiranaka, Ryosuke Kuroda and Kenichiro Kakutani
Cells 2024, 13(23), 2030; https://doi.org/10.3390/cells13232030 - 9 Dec 2024
Viewed by 578
Abstract
The mammalian target of rapamycin (mTOR), a serine/threonine kinase, promotes cell growth and inhibits autophagy. The following two complexes contain mTOR: mTORC1 with the regulatory associated protein of mTOR (RAPTOR) and mTORC2 with the rapamycin-insensitive companion of mTOR (RICTOR). The phosphatidylinositol 3-kinase (PI3K)/Akt/mTOR [...] Read more.
The mammalian target of rapamycin (mTOR), a serine/threonine kinase, promotes cell growth and inhibits autophagy. The following two complexes contain mTOR: mTORC1 with the regulatory associated protein of mTOR (RAPTOR) and mTORC2 with the rapamycin-insensitive companion of mTOR (RICTOR). The phosphatidylinositol 3-kinase (PI3K)/Akt/mTOR signaling pathway is important in the intervertebral disk, which is the largest avascular, hypoxic, low-nutrient organ in the body. To examine gene-silencing therapeutic approaches targeting PI3K/Akt/mTOR signaling in degenerative disk cells, an in vitro comparative study was designed between small interfering RNA (siRNA)-mediated RNA interference (RNAi) and clustered regularly interspaced short palindromic repeat (CRISPR)–CRISPR-associated protein 9 (Cas9) gene editing. Surgically obtained human disk nucleus pulposus cells were transfected with a siRNA or CRISPR–Cas9 plasmid targeting mTOR, RAPTOR, or RICTOR. Both of the approaches specifically suppressed target protein expression; however, the 24-h transfection efficiency differed by 53.8–60.3% for RNAi and 88.1–89.3% for CRISPR–Cas9 (p < 0.0001). Targeting mTOR, RAPTOR, and RICTOR all induced autophagy and inhibited apoptosis, senescence, pyroptosis, and matrix catabolism, with the most prominent effects observed with RAPTOR CRISPR–Cas9. In the time-course analysis, the 168-h suppression ratio of RAPTOR protein expression was 83.2% by CRISPR–Cas9 but only 8.8% by RNAi. While RNAi facilitates transient gene knockdown, CRISPR–Cas9 provides extensive gene knockout. Our findings suggest that RAPTOR/mTORC1 is a potential therapeutic target for degenerative disk disease. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of human disk intracellular PI3K/Akt/mTOR signaling pathway. The mTOR is a serine/threonine kinase that integrates nutrient signals to promote drive cell growth and division. It operates within the following two primary complexes: mTORC1 and mTORC2, which include RAPTOR and RICTOR, respectively. The downstream effectors of mTORC1, such as p70/S6K, are involved in controlling cell proliferation, mRNA translation, and protein synthesis, also associated with senescence and matrix catabolism. Autophagy is tightly suppressed by mTORC1 as well. The regulation of mTORC1 is mediated by the upstream class-I PI3K, with Akt serving as a crucial pro-survival mediator that prevents apoptosis. Furthermore, the negative feedback loop between p70/S6K and the class-I PI3K exists. To analyze the cascade-dependent functions of PI3K/Akt/mTOR signaling, gene suppression was performed using both siRNA-mediated RNAi-based and CRISPR–Cas9-based methods to target <span class="html-italic">mTOR</span> for both mTORC1 and mTORC2, <span class="html-italic">RAPTOR</span> for mTORC1, and <span class="html-italic">RICTOR</span> for mTORC2.</p>
Full article ">Figure 2
<p>Schematic illustration of the in vitro study design. Human degenerative intervertebral disk NP cells were surgically collected from patients who underwent lumbar discectomy or interbody fusion surgery. To retain the phenotype and replicate the physiologically hypoxic intervertebral disk environment, first-passage cells were cultured under 2% O<sub>2</sub> until they reached ~80% confluence. Gene knockdown and knockout targeting <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, and <span class="html-italic">RICTOR</span> were performed using both siRNA-mediated RNAi and CRISPR–Cas9, respectively. After the cells were transfected for 24 h, the suppression of mTOR, RAPTOR, and RICTOR and autophagy were evaluated by Western blotting. The cell number was counted. Cell viability was measured using the CCK-8 assay to evaluate the toxicity associated with RNAi and CRISPR–Cas9. Additionally, to mimic the clinically relevant low-nutrient and inflammatory disease conditions, following siRNA or CRISPR–Cas9 treatment for 24 h, the cells were stimulated with pro-inflammatory IL-1β in serum-free DMEM for an additional 24 h. Subsequent analyses included evaluating the apoptosis, pyroptosis, senescence, and matrix metabolism using Western blotting, TUNEL staining for apoptosis, and SA-β-gal staining for senescence.</p>
Full article ">Figure 3
<p>RNAi and CRISPR–Cas9 enhance the selective suppression of mTOR, RAPTOR, and RICTOR in human disk NP cells. (<b>A</b>) Western blot analysis for brachyury, CD24, and tubulin in the total protein extracts from five different batches of human disk NP cells in DMEM with 10% FBS. (<b>B</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA with each of two different sequences (Seq. 1 and Seq. 2) in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>C</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid with each of the three different guide RNA sequences (Seq. 1, Seq. 2, and Seq. 3) in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>D</b>) Fluorescence for phase contrast (gray), GFP (green), DAPI (blue), and merged signals in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA containing a GFP sequence in DMEM with 10% FBS to assess the transfection efficiency of the GFP-positive cells relative to the total DAPI-positive cells. (<b>E</b>) Morphological appearance of human disk NP cells 24 h post-transfection with <span class="html-italic">RAPTOR</span> siRNA or <span class="html-italic">RAPTOR</span> CRISPR–Cas9 plasmid in DMEM with 10% FBS to assess the number of adherent cells treated relative to the control. (<b>F</b>) CCK-8 assay in human disk NP cells 24 h post-transfection with control siRNA, control CRISPR–Cas9 plasmid, lipofection only, <span class="html-italic">RAPTOR</span> siRNA, or <span class="html-italic">RAPTOR</span> CRISPR–Cas9 plasmid in DMEM with 10% FBS to assess the viability of the cells treated relative to the control. Cells were counted in duplicated five random low-power fields (100×). Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). In (<b>A</b>), the immunoblots shown are all results from experiments with similar outcomes (<span class="html-italic">n</span> = 5). In (<b>B</b>–<b>E</b>), the immunoblots and cellular images shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 4
<p>Selective suppression of RAPTOR/mTORC1 inhibits autophagy and p70/S6K but differentially induces Akt activation in human disk NP cells. (<b>A</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA with the sequence showing the highest suppression efficiency in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>B</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid with the sequence presenting the highest suppression efficiency in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>C</b>) Western blot analysis for Akt, phosphorylated Akt (p-Akt), p70/S6K, phosphorylated p70/S6K (p-p70/S6K), and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in DMEM with 10% FBS. (<b>D</b>) Western blot analysis for Akt, p-Akt, p70/S6K, p-p70/S6K, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in DMEM with 10% FBS. (<b>E</b>) Western blot analysis for LC3, p62/SQSTM1, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>F</b>) Western blot analysis for LC3, p62/SQSTM1, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. Statistical analysis was performed using the paired <span class="html-italic">t</span>-test or one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots shown represent the typical results from experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 5
<p>Selective suppression of RAPTOR/mTORC1 inhibits apoptosis in human disk NP cells. (<b>A</b>) Western blot analysis for PARP, cleaved PARP, cleaved caspase-9, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>B</b>) Western blot analysis for PARP, cleaved PARP, cleaved caspase-9, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>C</b>) Fluorescence for TUNEL (green), DAPI (blue), and merged signals in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of TUNEL-positive cells relative to the total DAPI-positive cells. (<b>D</b>) Fluorescence for TUNEL (green), DAPI (blue), and merged signals in human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of TUNEL-positive cells relative to the total DAPI-positive cells. Cells were counted in duplicated five random low-power fields (100×). Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots and cellular images shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 6
<p>Selective suppression of RAPTOR/mTORC1 inhibits pyroptosis in human disk NP cells. (<b>A</b>) Western blot analysis for caspase-1, cleaved caspase-1, GSDMD, N-terminal GSDMD, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the expression levels of the target protein relative to tubulin. (<b>B</b>) Western blot analysis for caspase-1, cleaved caspase-1, GSDMD, N-terminal GSDMD, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the expression levels of the target protein relative to tubulin. Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 7
<p>Selective suppression of RAPTOR/mTORC1 inhibits senescence in human disk NP cells. (<b>A</b>) Western blot analysis for p16/INK4A, p21/WAF1/CIP1, p53, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>B</b>) Western blot analysis for p16/INK4A, p21/WAF1/CIP1, p53, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>C</b>) Colorimetric assay for the SA-β-gal signals (blue, indicated by black arrowheads) in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of SA-β-gal-positive cells relative to the total cells. (<b>D</b>) Colorimetric assay for the SA-β-gal signals (blue, indicated by black arrowheads) in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of SA-β-gal-positive cells relative to the total cells. Cells were counted in duplicated five random low-power fields (100×). Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots and cellular images shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 8
<p>Selective suppression of RAPTOR/mTORC1 increases matrix anabolism through decreased catabolic enzymes in human disk NP cells. (<b>A</b>) Western blot analysis for aggrecan, COL2A1, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>B</b>) Western blot analysis for aggrecan, COL2A1, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>C</b>) Western blot analysis for MMP-3, MMP-13, TIMP-1, and TIMP-2 in the supernatant protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>D</b>) Western blot analysis for MMP-3, MMP-13, TIMP-1, and TIMP-2 in the supernatant protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. The immunoblots shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 9
<p>RNAi facilitates transient <span class="html-italic">RAPTOR</span> gene knockdown but CRISPR–Cas9 provides extensive <span class="html-italic">RAPTOR</span> gene knockout in human disk NP cells. Western blot analysis for RAPTOR and tubulin in the total protein extracts from five different batches of human disk NP cells at 0, 24, 48, 72, 120, and 168 h post-transfection with <span class="html-italic">RAPTOR</span> siRNA or CRISPR–Cas9 plasmid in 10% FBS-supplemented DMEM with a media change every 48 h to assess the time-course expression levels of the RAPTOR protein relative to tubulin. Statistical analysis was performed using two-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are represented as the mean ± standard deviation (<span class="html-italic">n</span> = 5). The immunoblots shown are all results from the experiments with similar outcomes (<span class="html-italic">n</span> = 5).</p>
Full article ">
11 pages, 1594 KiB  
Article
Heparanase 2 Modulation Inhibits HSV-2 Replication by Regulating Heparan Sulfate
by James Hopkins, Ipsita Volety, Farreh Qatanani and Deepak Shukla
Viruses 2024, 16(12), 1832; https://doi.org/10.3390/v16121832 - 26 Nov 2024
Viewed by 394
Abstract
The host enzyme heparanase (HPSE) facilitates the release of herpes simplex virus type 2 (HSV-2) from target cells by cleaving the viral attachment receptor heparan sulfate (HS) from infected cell surfaces. HPSE 2, an isoform of HPSE, binds to but does not possess [...] Read more.
The host enzyme heparanase (HPSE) facilitates the release of herpes simplex virus type 2 (HSV-2) from target cells by cleaving the viral attachment receptor heparan sulfate (HS) from infected cell surfaces. HPSE 2, an isoform of HPSE, binds to but does not possess the enzymatic activity needed to cleave cell surface HS. Our study demonstrates that HSV-2 infection significantly elevates HPSE 2 protein levels, impacting two distinct stages of viral replication. We show that higher HPSE 2 negatively affects HSV-2 replication which may be through the regulation of cell surface HS. By acting as a competitive inhibitor of HPSE, HPSE 2 may be interfering with HPSE’s interactions with HS. We demonstrate that the enhanced expression of HPSE 2, either via viral infection or plasmid transfection, reduces HPSE’s ability to cleave HS, thereby hindering viral egress. Conversely, low HPSE 2 levels achieved through siRNA transfection allow HPSE to cleave more HS, reducing viral entry. Altogether, we propose a hypothetical model in which the modulation of HPSE 2 impedes HSV-2 replication by regulating HS availability on the cell surface. This dual role of HPSE 2 in viral replication and potential tumor suppression underscores its significance in cellular processes and viral pathogenesis. Full article
(This article belongs to the Special Issue Viruses and Eye Diseases)
Show Figures

Figure 1

Figure 1
<p>HPSE 2 increases in response to HSV-2 infection. (<b>A</b>). Increase in HPSE 2 mRNA levels: vaginal epithelial cells were infected with HSV-2 333, and samples were collected at 0, 12, 24, 36, and 48 h post-infection. The relative fold change over uninfected control is shown. (<b>B</b>). Representative Western blot of HPSE 2 expression after HSV-2 333 infection at an MOI of 1 with samples taken at 0, 12, 24, 36, and 48 h post-infection. (<b>C</b>). Representative immunofluorescence microscopy images of HPSE 2 stain. HSV-2 333 GFP was used to infect cells at a MOI of 0.1, then images were taken at 0, 12, 24, 36, and 48 h post-infection. The top row is HPSE 2, shown in red; the middle row is HSV-2 GFP; and the last row includes Hoechst, shown in blue, HPSE 2 stain, shown in red, and HSV-2, shown in green. Scale bar, 40 μm. Asterisks on plotted graph denote a significant difference as determined by 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, NS: not significant).</p>
Full article ">Figure 2
<p>Overexpression of HPSE 2 leads to lower viral replication (<b>A</b>). Representative Western blot of expression of viral protein gD after overexpression of HPSE 2 (upper panel). The quantified protein levels of gD are plotted, indicating decreased gD protein levels when isoforms are expressed compared to vector control (lower panel). VK2 cells were infected with HSV-2 333 24 h after transfection with plasmids overexpressing HPSE 2 a, b, and c and a control. Samples were collected at 0 for the control and 12, 24, and 48 h post-infection for all categories. (<b>B</b>). Quantification of plaque assay of media collected from cells that were overexpressing HPSE 2 a, b, and c and then were infected with HSV-2 at an MOI of 1. Samples were collected at 24 h post-infection. (<b>C</b>). Quantification of plaque assay of cell lysates collected from cells that were overexpressing HPSE 2 a, b, and c and then were infected with HSV-2 at an MOI of 1. Samples were collected at 24 h post-infection. (<b>D</b>). Quantification of infection after overexpressing HPSE 2 a, b, and c flow cytometry experiments, with samples collected 12, 24, and 36 h post-infection. All plotted results are presented as mean ± SEM of three independent experiments (n = 3). Asterisks denote a significant difference as determined by 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, NS: not significant).</p>
Full article ">Figure 3
<p>Knockdown of HPSE leads to lowerHSV-2 replication (<b>A</b>). Representative Western blot of expression of viral protein gD after knockdown of HPSE 2 and treatment with OGT 2115 after infection with HSV-2 333. The samples were collected 24 h post-infection (left panel). The quantified protein levels of gD are plotted, indicating decreased gD protein levels when isoforms are expressed compared to vector control (right panel). (<b>B</b>). Quantification of Western blot of expression of viral protein gD after knockdown of HPSE 2. (<b>C</b>). Quantification of a plaque assay of cell lysate collected from HPSE 2 knockdown cells or cells treated with OGT 2115, which then were infected with HSV-2 at an MOI of 1. Samples were collected 24 h post-infection. (<b>D</b>). Quantification of a plaque assay of media collected from HPSE 2 knockdown cells or cells treated with OGT 2115, which then were infected with HSV-2 at an MOI of 1. Samples were collected 24 h post-infection. (<b>E</b>). Quantification of infection after treatment with OGT 2115 or HPSE 2 knockdown in flow cytometry experiments. All plotted results are presented as mean ± SEM of three independent experiments (n = 3). Asterisks denote a significant difference as determined by 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; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, NS: not significant).</p>
Full article ">Figure 4
<p>Mechanism of HPSE 2 inhibition of HSV-2 infection. (<b>A</b>). Representative immunofluorescent microscopy images of HPSE 2 surface stain. HSV-2 333 GFP was used to infect cells at an MOI of 0.1, then images were taken 0, 12, 24, 36, and 48 h post-infection. The top row is HPSE 2 in red, the middle row is HSV-2 in GFP, and the last row includes Hoechst in blue, HPSE 2 stain in red, and HSV-2 in green. (<b>B</b>). Representative immunofluorescent images of surface HPSE 2c-myc stain. HSV-2 333 was used to infect cells at an MOI of 1 for 12 h. Upper left is myc stain only in uninfected sample, upper right is Hoescht and myc stain merged for uninfected sample, lower left is myc stain only for infected sample 12 h post-infection, and lower right is Hoechst and myc stain merged for infected sample 24 h post-infection. (<b>C</b>). Representative immunofluorescent images of surface heparan sulfate stain. HSV-2 333 was used to infect cells at an MOI of 1 for 24 h. Left column shows control cells, with uninfected cells on top and infected cells on the bottom. Right column shows cells overexpressing HPSE 2c, with uninfected cells on top and infected cells on the bottom. (<b>D</b>). Quantification of cell surface heparan sulfate flow cytometry experiments by flow with cells overexpressing HPSE 2. (<b>E</b>). Representative immunofluorescent images of surface heparan sulfate stain. HSV-2 333 was used to infect cells at an MOI of 1 for 24 h. Left column shows control cells, with uninfected cells on top and infected cells on the bottom. Right column shows cells with HPSE 2c knockdown, with uninfected cells on top and infected cells on the bottom. (<b>F</b>). Quantification of cell surface heparan sulfate flow cytometry experiments used cells with HPSE 2 knockdown. Scale bar, 40 μm. All plotted results are presented as mean ± SEM of three independent experiments (n = 3). Asterisks denote a significant difference as determined by 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, NS: not significant).</p>
Full article ">
14 pages, 3474 KiB  
Article
Enhancing the Chemosensitivity of MKN-45 Gastric Cancer Cells to Docetaxel via B7H6 Suppression: A Novel Therapeutic Strategy
by Elif Sibel Aslan, Nermin Akcali, Cuneyd Yavas, Sajjad Eslamkhah, Savas Gur and Lutfiye Karcioglu Batur
Life 2024, 14(12), 1546; https://doi.org/10.3390/life14121546 - 26 Nov 2024
Viewed by 502
Abstract
Purpose: Although chemotherapy is one of the standard treatments for gastric cancer, the disease’s resistance mechanisms continue to limit the survival rates. B7H6 (NCR3LG1), an immune checkpoint belonging to the B7 family, is significantly overexpressed in gastric cancer. This work investigated [...] Read more.
Purpose: Although chemotherapy is one of the standard treatments for gastric cancer, the disease’s resistance mechanisms continue to limit the survival rates. B7H6 (NCR3LG1), an immune checkpoint belonging to the B7 family, is significantly overexpressed in gastric cancer. This work investigated the possibility of using B7H6 suppression to improve the effectiveness of the widely used chemotherapy medication docetaxel. Materials and Methods: In this study, MKN-45 gastric cancer cells were transfected for 24 h with siRNA targeting B7H6, and then, docetaxel was added at optimal inhibitory doses (IC25 and IC50). To assess the impact of this combination therapy, cellular viability, proliferation, and migration were assessed using MTT assay, colony-forming unit assay, and wound-healing assay, respectively. Additionally, apoptosis and cell cycle status were evaluated by flow cytometry. Moreover, using qRT-PCR, the gene expression of B7H6 and indicators associated with apoptosis was also examined. Results: The sensitivity of MKN-45 cells to docetaxel was greatly increased by the siRNA-mediated knockdown of B7H6, resulting in a decrease in the drug’s IC50 value. When compared to each therapy alone, the combination of B7H6 siRNA plus docetaxel at IC50 levels exhibited a significant increase in apoptosis rate. The volume of cells arrested at the sub-G1 and G2-M phase was shown to rise when B7H6 siRNA transfection was combined with docetaxel. Furthermore, the combination treatment significantly decreased the ability of cells to migrate and form colonies. Conclusions: B7H6 suppression increases the susceptibility of MKN-45 gastric cancer cells to docetaxel treatment, resulting in decreased cellular proliferation and increased rates of apoptosis. The present work underscores the possibility of enhancing treatment results in gastric cancer by merging conventional chemotherapy with gene-silencing approaches. Full article
(This article belongs to the Section Physiology and Pathology)
Show Figures

Figure 1

Figure 1
<p>Analysis of the B7H6 immune checkpoint gene expression in gastric cancer cell lines. The MKN-45 cell line has the most significant level of expression (*** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>The expression of the <span class="html-italic">B7H6</span> gene in MKN-45 cells was inhibited in a dose-dependent manner using particular siRNAs at concentrations of 40, 60, 80, and 100 pmol for 24 h. The data indicate a significant reduction in <span class="html-italic">B7H6</span> expression, with the most prominent decline found at a concentration of 60 pmol siRNA (*** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>The MTT test findings demonstrate the suppressive effects of docetaxel on the vitality of MKN-45 cells. At concentrations of 15 μg/mL and 9.8 μg/mL, docetaxel decreased cell viability to 50% (IC50) and 25% (IC25), respectively. Following transfection, the IC50 and IC25 values exhibited a drop to 5.6 μg/mL and 2.35 μg/mL, respectively, suggesting an improved responsiveness to docetaxel.</p>
Full article ">Figure 4
<p>Wound-healing assay showing the impact of <span class="html-italic">B7H6</span> siRNA and docetaxel on MKN-45 cell migration at 0, 24, and 48 h. Cells treated with the combination of <span class="html-italic">B7H6</span> siRNA and docetaxel, especially at IC50, exhibited significantly reduced migration compared to those treated with either treatment alone.</p>
Full article ">Figure 5
<p>Colony-forming unit assay showing colony formation was reduced with <span class="html-italic">B7H6</span> siRNA, further decreased with docetaxel, and nearly eliminated with the combination treatment.</p>
Full article ">Figure 6
<p>(<b>A</b>) Apoptosis in MKN-45 cells treated with docetaxel at IC25 and IC50 doses. When docetaxel was applied alone at IC25 and IC50, the apoptosis rates observed were 13.89% and 31.8%, respectively. This indicates that while docetaxel independently promotes apoptosis, the effect is limited compared to combination treatments. (<b>B</b>) Apoptosis in MKN-45 cells treated with a combination of <span class="html-italic">B7H6</span> siRNA and docetaxel at IC25 and IC50 doses. The combination treatment significantly enhanced apoptosis rates, reaching 36.4% for the IC25 dose and 63.8% for the IC50 dose. This substantial increase highlights the synergistic effect of <span class="html-italic">B7H6</span> siRNA in enhancing the chemosensitivity of MKN-45 cells to docetaxel. Statistical analysis confirmed that these differences were highly significant (**** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>(<b>A</b>) Flow cytometry plots for the control group and cells treated with IC25 and IC50 doses of docetaxel. (<b>B</b>) Flow cytometry plots for cells treated with siRNA alone and in combination with IC25 and IC50 doses of docetaxel. Flow cytometry analysis of MKN-45 cells treated with docetaxel and <span class="html-italic">B7H6</span> siRNA demonstrates significant changes in cell cycle distribution across various treatment groups, including control, IC25, IC50, siRNA alone, and combinations of siRNA with IC25 and IC50 doses of docetaxel. Notably, treatment with docetaxel, especially when combined with <span class="html-italic">B7H6</span> siRNA, led to marked increases in sub-G1 (apoptotic) and G2-M phase arrest, indicative of enhanced apoptosis and cell cycle disruption. The combined treatment with IC25 and IC50 doses of docetaxel and siRNA significantly elevated the sub-G1 and G2-M phase cell populations compared to treatment with docetaxel alone.</p>
Full article ">Figure 8
<p>qRT-PCR analysis of apoptosis-related gene expression in MKN-45 cells. The combined treatment with <span class="html-italic">B7H6</span> siRNA and docetaxel significantly modulated the expression of key apoptotic markers. <span class="html-italic">Bax</span> (<b>A</b>) and caspase-3 (<b>B</b>) levels were notably increased, indicating enhanced pro-apoptotic activity, while <span class="html-italic">Bcl-2</span> expression (<b>C</b>) was significantly reduced, suggesting a decrease in anti-apoptotic signaling. These changes compared to individual treatments and control groups highlight the potentiation of apoptosis induced by the combination therapy. Asterisks (*, ***, ****) in the graph represent varying levels of statistical significance, with * indicating <span class="html-italic">p</span> &lt; 0.05, *** indicating <span class="html-italic">p</span> &lt; 0.001 and **** indicating <span class="html-italic">p</span> &lt; 0.0001 reflecting progressively higher levels of confidence in the observed differences.</p>
Full article ">
15 pages, 1932 KiB  
Article
Oxysophocarpine Prevents the Glutamate-Induced Apoptosis of HT–22 Cells via the Nrf2/HO–1 Signaling Pathway
by Ruiying Yuan, Dan Gao, Guibing Yang, Dongzhi Zhuoma, Zhen Pu, Yangzhen Ciren, Bin Li and Jianqing Yu
Curr. Issues Mol. Biol. 2024, 46(11), 13035-13049; https://doi.org/10.3390/cimb46110777 - 16 Nov 2024
Viewed by 706
Abstract
Oxysophocarpine (OSC), a quinolizidine alkaloid, shows neuroprotective potential, though its mechanisms are unclear. The aim of the present study was to investigate the neuroprotective effects of OSC through the nuclear factor erythroid 2−related factor 2 (Nrf2)/ heme oxygenase−1 (HO–1) signaling pathway using the [...] Read more.
Oxysophocarpine (OSC), a quinolizidine alkaloid, shows neuroprotective potential, though its mechanisms are unclear. The aim of the present study was to investigate the neuroprotective effects of OSC through the nuclear factor erythroid 2−related factor 2 (Nrf2)/ heme oxygenase−1 (HO–1) signaling pathway using the HT–22 cell line. Assessments of cell viability were conducted utilizing the 3−(4,5−dimethylthiazol−2−yl)−2,5−diphenyltetrazolium bromide (MTT) assay. Assessments of oxidative stress (OS) were conducted through the quantification of reactive oxygen species (ROS). The integrity of the mitochondrial membrane potential (MMP) was scrutinized using fluorescent probe technology. Apoptosis levels were quantified using terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) staining. The trafficking of Nrf2 within the cell nucleus was examined through immunofluorescence analysis. Furthermore, Western blotting (WB) was applied to evaluate the expression levels of proteins implicated in apoptosis and the Nrf2/HO–1 pathway. To further probe the influence of OSC on the overexpression of antioxidant enzymes, cells were subjected to transfection with HO–1 siRNA. The results showed that OSC inhibited glutamate-induced OS, as evidenced by reduced cell viability and ROS levels. Furthermore, the apoptotic condition induced by glutamate in HT–22 cells was significantly reduced following OSC treatment. More interestingly, the Nrf2/HO–1 signaling pathway was upregulated following OSC treatment. These results suggest that OSC can exert neuroprotective effects by regulating the Nrf2/HO–1 pathway to inhibit neuronal cell apoptosis, potentially aiding in the treatment of neurodegenerative diseases. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structure of oxysophocarpine.</p>
Full article ">Figure 2
<p>The effects of OSC on cell viability in HT–22 cells were assessed. HT–22 cells were exposed to varying concentrations (1.25, 2.5, 5, 10, 20 μM) of OSC for a period of 12 h. Cell viability was determined using the MTT assay. Each bar in the graph represents the mean ± standard deviation (SD), derived from three independent experiments (<span class="html-italic">n</span> = 3). “ns” stands for “not significant”, The bars marked with ## indicate a statistically significant difference compared to the control group (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>The influence of OSC on glutamate-induced cytotoxicity and ROS production in HT–22 cells was assessed. Prior to a 24 h exposure to glutamate at a concentration of 20 mM, HT–22 cells were subjected to pretreatment with a range of OSC concentrations (1.25, 2.5, 5, 10 μM). Panel (<b>A</b>) illustrates the assessment of cell viability utilizing the MTT assay, while Panel (<b>B</b>) depicts the quantification of ROS production using the DCF Fluorescence intensity. Trolox, administered at 50 μM, served as a benchmark for a positive control. The data are presented as a percentage relative to untreated cell populations, with each bar signifying the mean ± SD derived from triplicate experiments. Statistical significance is denoted as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 in contrast to the untreated control group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 in contrast to the group exposed solely to 20 mM glutamate. The presence or absence of treatments is indicated by “+” and “−”, respectively. ROS, reactive oxygen species.</p>
Full article ">Figure 4
<p>The impact of OSC on the modulation of the MMP and the expression of the apoptotic proteins BCL–2/BAX in glutamate-exposed HT–22 cells was investigated. HT–22 cells were pretreated with a range of concentrations (1.25, 2.5, 5, 10 μM) of OSC, followed by a 24 h exposure to glutamate at a concentration of 20 mM. (<b>A</b>) The MMP was evaluated using JC−1 staining, which was observed under a microscope at 200× magnification. Green fluorescence indicated mitochondrial depolarization, whereas red fluorescence represented normal polarization. (<b>B</b>) The levels of BCL–2/BAX were quantified through Western blotting (WB), with the expression levels normalized against actin as a loading control. The data, represented as mean values ± SD, were derived from three independent experiments (n = 3). Statistical significance is denoted as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 indicate significant differences compared to the untreated control; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent substantial differences from the group treated with glutamate alone (20 mM). The symbols “+” and “−” represent the inclusion or exclusion of the respective treatments.</p>
Full article ">Figure 5
<p>This study investigates the impact of OSC on the apoptotic response in HT-22 cells following glutamate exposure. (<b>A</b>) The apoptotic rate in HT–22 cells, subjected to 20 mM glutamate for 24 h with a prior treatment of OSC at concentrations of 1.25, 2.5, 5, and 10 μM, was ascertained using the TUNEL staining method. Apoptotic cells were identified by green fluorescence under a 200× microscope magnification. (<b>B</b>) The levels of apoptosis-related proteins, including cleaved caspase–3, caspase–3, cleaved caspase–9, and caspase–9, were assessed via WB analysis. The expression data were normalized against actin, a constitutively expressed protein. The results are expressed as a percentage relative to the control cells, which were not treated. Each bar graph displays the mean ± SD from three independent experiments (n = 3). Statistical significance is indicated as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 indicate significant differences from the untreated control; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 suggest substantial differences from the group treated solely with 20 mM glutamate. The inclusion or exclusion of treatments is indicated by “+” and “−”, respectively.</p>
Full article ">Figure 6
<p>Influence of OSC on Nrf2 translocation dynamics in HT–22 cells. The study examines the effect of OSC on the subcellular distribution of Nrf2 in HT–22 cells following exposure to a concentration of 10 μM for intervals of 0.5, 1, 1.5, or 2 h. (<b>A</b>,<b>B</b>) Nrf2 protein levels in both cytosolic and nuclear compartments were ascertained by WB analysis. This approach allows for the assessment of Nrf2 translocation from the cytoplasm to the nucleus in response to OSC treatment. (<b>C</b>) The visualization and quantification of Nrf2 translocation were further accomplished using immunofluorescence microscopy, providing a qualitative representation of protein movement within the cellular context. For the normalization of protein levels, cytosolic Nrf2 was referenced against actin, while nuclear Nrf2 was calibrated against lamin B1, ensuring the accuracy of the comparative analysis. Data are depicted as the mean ± SD derived from three independent experiments (n = 3). Statistical significance is represented by the following notations: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 indicate significant differences when compared to the control group without treatment.</p>
Full article ">Figure 7
<p>Modulation of HO–1 protein expression by OSC in HT–22 Cells. (<b>A</b>) Cells were exposed to a range of OSC concentrations (1.25, 2.5, 5, 10 μM) for a duration of 12 h to determine the dose-dependent effect on HO–1 expression. Cobalt protoporphyrin (CoPP), at a concentration of 20 μM, served as a positive control to validate the response. (<b>B</b>) To explore the time course of HO–1 induction, cells were treated with a fixed concentration of 10 μM OSC for varying periods. The protein expression of HO–1 was quantified using WB analysis, a method that allows for the detection and quantification of specific proteins. The results were normalized relative to actin, a reference protein, to adjust for any variations in protein loading. The data presentation follows the standard format, where each bar graph segment illustrates the mean ± SD from triplicate samples (n = 3), ensuring the reproducibility and reliability of the findings. Statistical significance is denoted by the symbols <span class="html-italic"><sup>#</sup> p</span> &lt; 0.05 and <span class="html-italic"><sup>##</sup> p</span> &lt; 0.01, which indicate significant differences in HO–1 expression levels when compared to the control group without treatment. The presence or absence of treatment is indicated by “+” and “−” signs, respectively.</p>
Full article ">Figure 8
<p>Impact of HO–1 knockdown on HT–22 cell response to OSC and glutamate challenge. This study delineates the consequences of HO–1 suppression in HT–22 cells under conditions designed to mimic OS. (<b>A</b>,<b>D</b>) The survival of HT–22 cells, following pretreatment with 10 μM OSC in conjunction with or without 50 μM SnPP and si−HO–1, was subsequently challenged with 20 mM glutamate for 24 h. The quantitative assessment of cell viability was performed using the MTT assay. (<b>B</b>,<b>E</b>) The production of ROS was evaluated through DCF fluorescence measurement, providing a quantitative assessment of intracellular ROS levels. (<b>C</b>) Representative WB images illustrate the levels of HO–1 protein expression in the treated cells, offering a visual confirmation of the HO–1 knockdown efficacy. Data are presented as the mean ± SD from three independent experiments (n = 3), ensuring the statistical robustness of the findings. Statistical significance is indicated as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 denote significant differences relative to the untreated control group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate substantial differences when compared to the group treated solely with 20 mM glutamate; <sup>%</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>%%</sup> <span class="html-italic">p</span> &lt; 0.01 signify substantial differences in comparison to the group treated with 10 μM OSC. The presence or absence of specific treatments is denoted by the symbols “+” and “−”.</p>
Full article ">
15 pages, 4067 KiB  
Article
p21Waf1/Cip1 Is a Novel Downstream Target of 40S Ribosomal S6 Kinase 2
by Alakananda Basu and Zhenyu Xuan
Cancers 2024, 16(22), 3783; https://doi.org/10.3390/cancers16223783 - 10 Nov 2024
Viewed by 614
Abstract
Background/Objectives: The ribosomal S6 kinase 2 (S6K2) acts downstream of the mechanistic target of rapamycin complex 1 and is a homolog of S6K1 but little is known about its downstream effectors. The objective of this study was to use an unbiased transcriptome [...] Read more.
Background/Objectives: The ribosomal S6 kinase 2 (S6K2) acts downstream of the mechanistic target of rapamycin complex 1 and is a homolog of S6K1 but little is known about its downstream effectors. The objective of this study was to use an unbiased transcriptome profiling to uncover how S6K2 promotes breast cancer cell survival. Methods: RNA-Seq analysis was performed to identify novel S6K2 targets. Cells were transfected with siRNAs or plasmids containing genes of interest. Western blot analyses were performed to quantify total and phosphorylated proteins. Apoptosis was monitored by treating cells with different concentrations of doxorubicin. Results: Silencing of S6K2, but not S6K1, decreased p21 in MCF-7 and T47D breast cancer cells. Knockdown of Akt1 but not Akt2 decreased p21 in MCF-7 cells whereas both Akt1 and Akt2 knockdown attenuated p21 in T47D cells. While Akt1 overexpression enhanced p21 and partially reversed the effect of S6K2 deficiency on p21 downregulation in MCF-7 cells, it had little effect in T47D cells. S6K2 knockdown increased JUN mRNA and knockdown of cJun enhanced p21. Low concentrations of doxorubicin increased, and high concentrations decreased p21 levels in T47D cells. Silencing of S6K2 or p21 sensitized T47D cells to doxorubicin via c-Jun N-terminal kinase (JNK)-mediated downregulation of Mcl-1. Conclusions: S6K2 knockdown enhanced doxorubicin-induced apoptosis by downregulating the cell cycle inhibitor p21 and the anti-apoptotic protein Mcl-1 via Akt and/or JNK. Full article
(This article belongs to the Section Molecular Cancer Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>RNA-Seq analysis. (<b>A</b>). Venn graph of DEGs detected using both Cuffdiff and DESeq following S6K2 KD. FDR &lt; 0.05 was applied for each method. (<b>B</b>). The representative gene ontology terms of functional annotation clusters, which are significantly enriched in 118 shared DEGs (FDR &lt; 0.05). (<b>C</b>). Densitometric quantification of <span class="html-italic">CDKN1A</span> mRNA normalized with GAPDH control. The asterisk (*) indicates a significant difference from control siRNA-transfected cells (<span class="html-italic">p</span> &lt; 0.05) using paired Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>T47D (<b>A</b>,<b>B</b>) or MCF-7 (<b>C</b>,<b>D</b>) cells were transfected with or without control non-targeting siRNA or SMARTpool (SP) S6K1 or S6K2 siRNA. Western blot analyses were performed with indicated antibodies. The intensity of p21 was determined using ImageJ and normalized with respect to loading control. Each bar represents mean ± S.E. <span class="html-italic">p</span> values were calculated using a paired Student’s <span class="html-italic">t</span> test. (<b>E</b>). Different concentrations of cell lysates from MCF-7 cells transfected with an empty vector pcDNA3 (PC) or a vector containing S6K2 construct were subjected to Western blot analyses with indicated antibodies.</p>
Full article ">Figure 3
<p>T47D (<b>A</b>,<b>B</b>) or MCF-7 (<b>C</b>,<b>D</b>) cells were transfected with indicated siRNAs and Western blot analyses were performed with indicated antibodies. Each bar represents the mean ± S.E of four independent experiments. <span class="html-italic">p</span> values were calculated using paired Student’s <span class="html-italic">t</span> test of control versus individual siRNA as described under <a href="#cancers-16-03783-f002" class="html-fig">Figure 2</a>. ***, <span class="html-italic">p</span> ≤ 0.0005; **, <span class="html-italic">p</span> ≤ 0.005; *, <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 4
<p>T47D (<b>A</b>,<b>B</b>) or MCF-7 (<b>C</b>,<b>D</b>) cells were transfected with indicated siRNAs and Western blot analyses were performed with indicated antibodies. Each bar represents mean ± S.E of at least six independent experiments. <span class="html-italic">p</span> values were calculated using paired Student’s <span class="html-italic">t</span> test.</p>
Full article ">Figure 5
<p>T47D (<b>A</b>,<b>B</b>) or MCF-7 (<b>C</b>,<b>D</b>) cells were transfected with control or S6K2 siRNA and then infected with or without adenoviral vectors containing Akt1. Western blot analyses were performed with indicated antibodies. Each bar represents mean ± S.E of six independent experiments. <span class="html-italic">p</span> values calculated using paired <span class="html-italic">t</span> test of control versus Akt1 overexpressing cells: T47D, <span class="html-italic">p</span> = 0.0007; MCF-7, <span class="html-italic">p</span> = 0.0005; Light gray bar, control siRNA; black bar, S6K2 siRNA.</p>
Full article ">Figure 6
<p>T47D cells were transfected with indicated siRNAs. Western blot analyses were performed with indicated antibodies (<b>A</b>,<b>C</b>,<b>E</b>). The intensities of cJun (<b>B</b>) and p21 (<b>D</b>) were determined using ImageJ and normalized with respect to loading controls. Each bar represents mean ± S.E. <span class="html-italic">p</span> values were calculated using paired Student’s <span class="html-italic">t</span> test.</p>
Full article ">Figure 7
<p>T47D cells were transfected with control non-targeting siRNA or S6K2 siRNA and then treated with indicated concentrations of doxorubicin (Dox). Western blot analyses were performed with indicated antibodies. The band corresponding to cleaved caspase-7 was quantified using ImageJ and the intensities of bands normalized with loading controls are shown.</p>
Full article ">Figure 8
<p>T47D cells were transfected with control non-targeting siRNA or p21 siRNA and then treated with indicated concentrations of doxorubicin. Western blot analyses were performed with indicated antibodies. The bands corresponding to p21, cleaved caspase-3, caspase-7, and PARP were quantified using ImageJ, and the intensities of bands normalized with loading controls are shown.</p>
Full article ">Figure 9
<p>T47D cells were transfected with control non-targeting siRNA, S6K2 and/or c-Jun siRNA and then treated with or without 0.3 and 1.0 µM (<b>A</b>) or 10 µM (<b>B</b>) doxorubicin. Western blot analysis was performed with indicated antibodies. The band corresponding to cleaved caspase-7 or PARP was quantified using ImageJ and the intensities of bands were normalized with tubulin.</p>
Full article ">Figure 10
<p>T47D cells were transfected with control non-targeting siRNA or JNK1 siRNA and then treated with indicated concentrations of doxorubicin. Western blot analyses were performed with indicated antibodies.</p>
Full article ">
14 pages, 2205 KiB  
Article
miR-29a Downregulates PIK3CA Expression and Inhibits Cervical Cancer Cell Dynamics: A Comparative Clinical Analysis
by Hyorim Jeong, Kangchan Choi, Dasom Hwang, Sunyoung Park, Yong Serk Park and Hyeyoung Lee
Curr. Issues Mol. Biol. 2024, 46(11), 12704-12717; https://doi.org/10.3390/cimb46110754 - 8 Nov 2024
Viewed by 889
Abstract
HPV/pap tests are widely used for cervical cancer screening, playing a crucial role in early diagnosis and guiding future treatment options. However, approximately 50% of cervical cancer patients are diagnosed at an advanced stage, which is associated with higher recurrence rates and poorer [...] Read more.
HPV/pap tests are widely used for cervical cancer screening, playing a crucial role in early diagnosis and guiding future treatment options. However, approximately 50% of cervical cancer patients are diagnosed at an advanced stage, which is associated with higher recurrence rates and poorer survival outcomes than early-stage diagnoses. This underscores the need for effective treatments for advanced-stage cervical cancer. Among the various oncogenes implicated in cancer, PIK3CA expression is known to cause cervical cancer, suggesting that inhibiting PIK3CA may impede cervical cancer progression. In this study, we transfected PIK3CA-overexpressing tumor cells (SiHa, C33A, and HeLa) with miR-29a, a microRNA extensively studied as a therapeutic candidate for oncogene suppression in various tumor types. We conducted RT-qPCR and Western blot analyses to assess changes in PIK3CA expression at the RNA and protein levels. Wound healing and cell migration assays were used to evaluate the effects of miR-29a on cell division and migration in HeLa cells. We confirmed a reduction in PIK3CA expression at both RNA and protein levels following miR-29a transfection. After transfecting miR-29a into HeLa cells, we observed a reduction in cell division and migration, as demonstrated by wound healing and cell migration assays. Additionally, we found that miR-29a binds to the 3′-UTR region of PIK3CA, leading to a reduction in its gene expression. Furthermore, we correlated the concentration of miR-29a in clinical histologic biopsy samples from cervical cancer patients with disease progression. These findings indicate that miR-29a can slow the progression of cervical cancer by targeting PIK3CA and potentially aid in its treatment. miR-29a shows promise as a therapeutic agent for inhibiting oncogene expression and controlling cervical cancer progression, especially in advanced-stage cases. Full article
Show Figures

Figure 1

Figure 1
<p>Differential expression of miR-29a and PIK3CA in HPV-negative and HPV-positive cervical cancer cell lines. (<b>a</b>) Relative expression levels of miR-29a in three cervical cancer cell lines: C33A (HPV-negative), SiHa (HPV type 16 positive), and HeLa (HPV type 18 positive), as determined by RT-qPCR. C33A cells exhibit the highest expression of miR-29a, whereas HeLa cells show the lowest levels. Error bars represent the standard deviation from three independent experiments. (<b>b</b>) Relative expression levels of PIK3CA in the same cell lines, as quantified by RT-qPCR. C33A cells show the highest PIK3CA expression, with both SiHa and HeLa displaying lower but comparable expression levels. Statistical significance is denoted as * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Inhibition of PIK3CA expression by miR-29a in cervical cancer cell lines. (<b>a</b>) Relative expression levels of PIK3CA mRNA in C33A (HPV-negative), SiHa (HPV type 16 positive), and HeLa (HPV type 18 positive) cervical cancer cell lines following transfection with 200 pmole of miR-29a, as determined by RT-qPCR. (<b>b</b>) Western blot analysis showing PIK3CA protein levels in HeLa cells transfected with either scramble miRNA (control) or miR-29a. Error bars represent the standard deviation from three independent experiments. Statistical significance is indicated as * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Effect of miR-29a on HeLa cell proliferation and migration. (<b>a</b>) Wound healing assay results for HeLa cells treated with scramble miRNA (control) and miR-29a demonstrate that miR-29a significantly inhibits cell growth and division. Images were taken at the indicated time points to assess the closure of the wound area. (<b>b</b>) Cell migration assay results for HeLa cells treated with scramble miRNA and miR-29a. The black specks (highlighted by white arrows) indicate the migrated cells. Statistical significance is denoted as *** <span class="html-italic">p</span> &lt; 0.001. Error bars represent the standard deviation from three independent experiments.</p>
Full article ">Figure 4
<p>Specificity of miR-29a for PIK3CA 3′-UTR binding validated by luciferase activity. (<b>a</b>) Schematic representation of the pmiRGLO-3′UTR plasmids used for the luciferase assay, showing both the wild-type and mutant versions of the PIK3CA 3′-UTR. (<b>b</b>) Luminescence results from co-transfection of HeLa cells with either scramble miRNA or miR-29a and the pmiRGLO-3′UTR plasmids. The data indicate that miR-29a significantly reduces luciferase activity when the wild-type 3′-UTR is present, confirming specific binding to PIK3CA. Error bars represent the mean ± standard deviation (S.D.) of three independent experiments. Statistical significance is denoted as *** <span class="html-italic">p</span> &lt; 0.001, ns means not significant.</p>
Full article ">Figure 5
<p>Analysis of miR-29a and PIK3CA expression levels in normal and cervical cancer samples. (<b>a</b>) Box plot comparing the relative expression levels of miR-29a between normal and cervical cancer patients. (<b>b</b>) Box plot comparing the relative expression levels of PIK3CA mRNA between normal and cervical cancer patients. (<b>c</b>) Analyses of tumor size, the relative expression levels of miR-29a, and PIK3CA in patients infected with high-risk HPV types (16 and 18). (<b>d</b>) Scatter plot showing the negative correlation between miR-29a and PIK3CA mRNA expression levels, based on the TCGA dataset. Error bars represent the standard deviation from three independent experiments.</p>
Full article ">Figure 6
<p>Analysis of miR-29a and PIK3CA mRNA expression levels in cervical cancer and non-cervical cancer tissues. (<b>a</b>) Box plot comparing the relative expression levels of miR-29a and PIK3CA mRNA in cervical tissue samples from patients with and without cervical cancer. The data demonstrate a significant difference in the expression levels of miR-29a and PIK3CA between the two groups. (<b>b</b>) Regression analysis plot showing the relationship between miR-29a and PIK3CA mRNA expression levels, indicating a potential correlation between these two markers. Statistical significance is denoted as *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
12 pages, 2111 KiB  
Article
The Functional Role of the Long Non-Coding RNA LINCMD1 in Leiomyoma Pathogenesis
by Tsai-Der Chuang, Nhu Ton, Shawn Rysling and Omid Khorram
Int. J. Mol. Sci. 2024, 25(21), 11539; https://doi.org/10.3390/ijms252111539 - 27 Oct 2024
Viewed by 779
Abstract
Existing evidence indicates that LINCMD1 regulates muscle differentiation-related gene expression in skeletal muscle by acting as a miRNA sponge, though its role in leiomyoma development is still unknown. This study investigated LINCMD1′s involvement in leiomyoma by analyzing paired myometrium and leiomyoma tissue samples [...] Read more.
Existing evidence indicates that LINCMD1 regulates muscle differentiation-related gene expression in skeletal muscle by acting as a miRNA sponge, though its role in leiomyoma development is still unknown. This study investigated LINCMD1′s involvement in leiomyoma by analyzing paired myometrium and leiomyoma tissue samples (n = 34) from patients who had not received hormonal treatments for at least three months prior to surgery. Myometrium smooth muscle cells (MSMCs) were isolated, and gene expression of LINCMD1 and miR-135b was assessed via qRT-PCR, while luciferase assays determined the interaction between LINCMD1 and miR-135b. To examine the effects of LINCMD1 knockdown, siRNA transfection was applied to a 3D MSMC spheroid culture, followed by qRT-PCR and Western blot analyses of miR-135b, APC, β-Catenin and COL1A1 expression. The results showed that leiomyoma tissues had significantly reduced LINCMD1 mRNA levels, regardless of patient race or MED12 mutation status, while miR-135b levels were elevated compared to matched myometrium samples. Luciferase assays confirmed LINCMD1′s role as a sponge for miR-135b. LINCMD1 knockdown in MSMC spheroids increased miR-135b levels, reduced APC expression, and led to β-Catenin accumulation and higher COL1A1 expression. These findings highlight LINCMD1 as a potential therapeutic target to modulate aberrant Wnt/β-Catenin signaling in leiomyoma. Full article
(This article belongs to the Special Issue The Role of Non‐coding RNAs in Human Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>An inverse relationship in the expression of LINCMD1 and miR-135b in leiomyomas. qRT-PCR analysis showing the expression levels of lncRNA LINCMD1 (<b>A</b>) and miR-135b (<b>B</b>) in 34 paired samples of myometrium (Myo) and leiomyomas (Lyo). Relative fold change in LINCMD1 (<b>C</b>) and miR-135b (<b>D</b>) expression (Lyo/paired Myo) is displayed based on MED12 mutation status, comparing MED12 wild type (n = 10) and MED12 mutated (n = 24) samples. Data are presented as mean ± SEM, with statistical significance indicated as * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>siRNA-mediated knockdown of LINCMD1 in myometrium spheroid cells for 96 h significantly reduced LINCMD1 expression levels (<b>A</b>) and led to an increased expression of miR-135b (<b>B</b>). Data (n = 4) are presented as mean ± SEM, with *** <span class="html-italic">p</span> &lt; 0.01 indicating statistical significance.</p>
Full article ">Figure 3
<p>LINCMD1 directly targets miR-135b, and the resulting changes in miR-135b regulate the β-Catenin signaling pathway. (<b>A</b>) Sequence alignment showing the coordinated positions of LINCMD1 with miR-135b. (<b>B</b>) Relative luciferase activity in myometrium spheroid cells transfected with Renilla and Firefly luciferase reporter pEZX-MT01 (Control) or pEZX-MT01 (LINCMD1). The ratio of Firefly to Renilla luciferase activity was measured after 48 h and expressed as relative luciferase activity compared to NC, which was independently set to 1 (n = 6). (<b>C</b>) Western blot analysis of APC, non-phosphorylated β-Catenin at serine 45, total β-Catenin, and COL1A1 following the transfection of myometrium spheroid cells with control pre-miR oligonucleotides (NC) or pre-miR-135b for 96 h. The relative band intensities are presented in a bar graph (<b>D</b>, n = 3) as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>An inverse relationship between APC expression and β-Catenin activity in leiomyomas. (<b>A</b>) qRT-PCR analysis showing the expression levels of APC and β-Catenin in 34 paired myometrium (Myo) and leiomyoma (Lyo) samples. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Their relative expression is shown as fold change (Lyo/paired Myo) based on MED12 mutation status. (<b>C</b>) Western blot analysis of APC, non-phosphorylated β-Catenin at serine 45, total β-Catenin, and COL1A1 in tissue extracts (n = 8) from myometrium (M) and paired leiomyoma (L). The relative band intensities are displayed in (<b>D</b>), with data presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>The LINCMD1/miR-135/APC/β-Catenin signaling axis in leiomyomas contributes to ECM accumulation. (<b>A</b>) qRT-PCR analysis showing mRNA expression levels of APC, β-Catenin, and COL1A1 in primary myometrium spheroid cells following transfection with LINCMD1 siRNA for 96 h (n = 4, * <span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Representative Western blot analysis of β-Catenin, non-phosphorylated β-Catenin at serine 45, total β-Catenin, and COL1A1 after LINCMD1 siRNA transfection in primary myometrium spheroid cells for 96 h. Relative band intensities are shown in (<b>C</b>), with data presented as mean ± SEM (n = 4, * <span class="html-italic">p</span> &lt; 0.05). (<b>D</b>) A representative gel showing levels of APC, non-phosphorylated β-Catenin at serine 45, total β-Catenin, and COL1A1 after co-transfection of LINCMD1 siRNA with either anti-NC or anti-miR-135b in primary myometrium spheroid cells for 96 h. (<b>E</b>) Bar plot depicting the mean relative band intensities, presented as mean ± SEM (n = 4, * <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 ">
28 pages, 8346 KiB  
Article
Autophagy Regulator Rufy 4 Promotes Osteoclastic Bone Resorption by Orchestrating Cytoskeletal Organization via Its RUN Domain
by Eiko Sakai, Minoru Saito, Yu Koyanagi, Yoshitsugu Takayama, Fatima Farhana, Yu Yamaguchi and Takayuki Tsukuba
Cells 2024, 13(21), 1766; https://doi.org/10.3390/cells13211766 - 25 Oct 2024
Viewed by 903
Abstract
Rufy4, a protein belonging to the RUN and FYVE domain-containing protein family, participates in various cellular processes such as autophagy and intracellular trafficking. However, its role in osteoclast-mediated bone resorption remains uncertain. In this study, we investigated the expression and role of the [...] Read more.
Rufy4, a protein belonging to the RUN and FYVE domain-containing protein family, participates in various cellular processes such as autophagy and intracellular trafficking. However, its role in osteoclast-mediated bone resorption remains uncertain. In this study, we investigated the expression and role of the Rufy4 gene in osteoclasts using small interfering RNA (siRNA) transfection and gene overexpression systems. Our findings revealed a significant increase in Rufy4 expression during osteoclast differentiation. Silencing Rufy4 enhanced osteoclast differentiation, intracellular cathepsin K levels, and formation of axial protrusive structures but suppressed bone resorption. Conversely, overexpressing wild-type Rufy4 in osteoclasts hindered differentiation while promoting podosome formation and bone resorption. Similarly, overexpression of a Rufy4 variant lacking the RUN domain mimics the effects of Rufy4 knockdown, significantly increasing intracellular cathepsin K levels, promoting osteoclastogenesis, and elongated axial protrusions formation, yet inhibiting bone resorption. These findings indicate that Rufy4 plays a critical role in osteoclast differentiation and bone resorption by regulating the cytoskeletal organization through its RUN domain. Our study provides new insights into the molecular mechanisms governing osteoclast activity and underscores Rufy4’s potential as a novel therapeutic target for bone disorders characterized by excessive bone resorption. Full article
Show Figures

Figure 1

Figure 1
<p>Increased <span class="html-italic">Rufy4</span> expression during osteoclast differentiation and impaired osteoclast formation following <span class="html-italic">Rufy4</span> knockdown. (<b>a</b>) List of upregulated transcripts in <span class="html-italic">Nrf2</span>-deficient osteoclasts compared to <span class="html-italic">Keap1</span>-deficient cells. (<b>b</b>) <span class="html-italic">Rufy4</span> mRNA expression was assessed by quantitative real-time PCR during osteoclast differentiation in RAW-D cells treated with 100 ng/mL RANKL. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>c</b>) Schematic representation of the <span class="html-italic">Rufy4</span> knockdown experiment schedule. (<b>d</b>) <span class="html-italic">Rufy4</span> knockdown efficacy was assessed by measuring the <span class="html-italic">Rufy4</span> mRNA levels. RAW-D cells were transfected with control or Rufy4-specific siRNA probes on days 1 and 2 in the presence of 100 ng/mL RANKL. Cells were collected on day 3, and RNA was prepared. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>e</b>) Tartrate-resistant acid phosphatase (TRAP) staining of control and <span class="html-italic">Rufy4</span>-knockdown osteoclasts. Control or Rufy4-knockdown RAW-D cells were stimulated with RANKL (100 ng/mL). On day 4, cells were fixed and stained with TRAP. Scale bar, 50 μm. (<b>f</b>) The number of TRAP-positive multinucleated cells was counted. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p><span class="html-italic">Rufy4</span> knockdown promotes osteoclast differentiation but impairs bone resorption. (<b>a</b>) <span class="html-italic">Rufy4</span> knockdown markedly increases the expression of osteoclast marker genes. Control or <span class="html-italic">Rufy4</span>-knockdown RAW-D cells were cultured with 100 ng/mL RANKL for 3 days. After the isolation of mRNA, quantitative real-time PCR was performed. Data are presented as the mean ± SD from 3 independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>b</b>) <span class="html-italic">Rufy4</span> knockdown upregulates osteoclast marker proteins. Control or <span class="html-italic">Rufy4</span>-knockdown RAW-D cells were cultured with 100 ng/mL RANKL for 3 days and harvested. Lysates were subjected to western blot analysis with specific antibodies against Cathepsin K and c-Src. LAMP1, Cathepsin D, and β-actin (control). Representative immunoblots are shown, and the quantification results are presented as the mean ± SD from 3 independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>c</b>) <span class="html-italic">Rufy4</span> knockdown impairs bone resorption. Control and <span class="html-italic">Rufy4</span>-knockdown RAW-D cells were seeded on Osteo Assay plates and cultured with 500 ng/mL RANKL for 7–10 days. Images of resorption pits are shown. Scale bar, 200 μm. (<b>d</b>) Resorption areas were calculated using the ImageJ software. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p><span class="html-italic">Rufy4</span> knockdown promotes large axial protrusive structure formation. Control RAW-D cells or <span class="html-italic">Rufy4</span> knockdown RAW-D cells were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed and stained with phalloidin (green) and α-tubulin antibody (magenta) and then visualized by confocal microscopy (A–E). (<b>a</b>) Control osteoclasts exhibited actin accumulation along the cell periphery (A). Tubulin was distributed throughout the cytoplasm (B). Arrows indicate protrusive structures in the tail-like region (C). High magnification image of the dotted square d in panel C. Arrowheads indicate the presence of hair-like structures (D). The cell contour was traced by blue, yellow, and red lines corresponding to the Type 1, 2, and 3 lines (E). Only the traced lines for (E) are visualized in (F). Representative fluorescence images of control osteoclasts (G–I). Traced lines for (G), (H), and (I) are shown in (J–L). (<b>b</b>) <span class="html-italic">Rufy4</span>-knockdown osteoclasts exhibited actin distribution along the cell periphery. (A) Tubulin was distributed throughout the cytoplasm (B). High-magnification image of square d and square e in panel C (D and E, respectively). Elongated axial protrusive structures (D, arrows) and actin accumulation (E, arrows) are visualized. The cell contour was traced by blue, yellow, and red lines corresponding to the Type 1, 2, and 3 lines (F). Representative fluorescence images of <span class="html-italic">Rufy4</span>-knockdown osteoclasts (G–I). Traced lines for (G–I) are shown in (J–L). Scale bar, 20 μm. (<b>c</b>) The relative length of type 1 or type 2 plus 3 curves per whole cell perimeter was compared between the control and <span class="html-italic">Rufy4</span>-knockdown osteoclasts. The total number of analyzed cells was 62 and 63 for the control and <span class="html-italic">Rufy4</span>-knockdown groups, respectively. All values are presented as mean ± SD (** <span class="html-italic">p</span> &lt; 0.01, compared to the control).</p>
Full article ">Figure 3 Cont.
<p><span class="html-italic">Rufy4</span> knockdown promotes large axial protrusive structure formation. Control RAW-D cells or <span class="html-italic">Rufy4</span> knockdown RAW-D cells were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed and stained with phalloidin (green) and α-tubulin antibody (magenta) and then visualized by confocal microscopy (A–E). (<b>a</b>) Control osteoclasts exhibited actin accumulation along the cell periphery (A). Tubulin was distributed throughout the cytoplasm (B). Arrows indicate protrusive structures in the tail-like region (C). High magnification image of the dotted square d in panel C. Arrowheads indicate the presence of hair-like structures (D). The cell contour was traced by blue, yellow, and red lines corresponding to the Type 1, 2, and 3 lines (E). Only the traced lines for (E) are visualized in (F). Representative fluorescence images of control osteoclasts (G–I). Traced lines for (G), (H), and (I) are shown in (J–L). (<b>b</b>) <span class="html-italic">Rufy4</span>-knockdown osteoclasts exhibited actin distribution along the cell periphery. (A) Tubulin was distributed throughout the cytoplasm (B). High-magnification image of square d and square e in panel C (D and E, respectively). Elongated axial protrusive structures (D, arrows) and actin accumulation (E, arrows) are visualized. The cell contour was traced by blue, yellow, and red lines corresponding to the Type 1, 2, and 3 lines (F). Representative fluorescence images of <span class="html-italic">Rufy4</span>-knockdown osteoclasts (G–I). Traced lines for (G–I) are shown in (J–L). Scale bar, 20 μm. (<b>c</b>) The relative length of type 1 or type 2 plus 3 curves per whole cell perimeter was compared between the control and <span class="html-italic">Rufy4</span>-knockdown osteoclasts. The total number of analyzed cells was 62 and 63 for the control and <span class="html-italic">Rufy4</span>-knockdown groups, respectively. All values are presented as mean ± SD (** <span class="html-italic">p</span> &lt; 0.01, compared to the control).</p>
Full article ">Figure 4
<p>Overexpression of eGFP-<span class="html-italic">Rufy4</span> inhibits osteoclast formation and stimulates bone resorption. (<b>a</b>) Quantitative RT-PCR analysis of <span class="html-italic">Rufy4</span> mRNA expression in RAW-D cells expressing either eGFP or eGFP-<span class="html-italic">Rufy4</span>. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>b</b>) Western blot analysis of RAW-D cells expressing eGFP or eGFP-<span class="html-italic">Rufy4</span>. Cell lysates were subjected to western blot analysis using anti-GFP or anti-GAPDH antibodies. (<b>c</b>) TRAP staining of eGFP- and eGFP-<span class="html-italic">Rufy4</span>-overexpressing osteoclasts. eGFP- or eGFP-<span class="html-italic">Rufy4</span>-overexpressing RAW-D cells were stimulated with 100 ng/mL RANKL for 3 days and then fixed and stained for TRAP. Scale bar, 50 μm. (<b>d</b>) Number of TRAP-positive multinucleated cells. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>e</b>) Bone resorption area of eGFP or eGFP-<span class="html-italic">Rufy4</span>-overexpressing osteoclasts. Cells were seeded onto Osteo Assay Stripwell Plates containing 500 ng/mL RANKL. The images show the bone resorption area for each osteoclast. Scale bar, 20 μm. (<b>f</b>) The bone resorption area was determined using the ImageJ software. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>g</b>) Western blot analysis of eGFP or eGFP-<span class="html-italic">Rufy4</span>-overexpressing RAW-D cells cultured with 100 ng/mL RANKL for 2 or 3 days. Cells were harvested on the indicated day, and lysates were subjected to western blot analysis with specific antibodies against various proteins involved in osteoclast differentiation and actin polymerization, with GAPDH as a control. Representative immunoblots are shown, and the quantification results are presented as the mean ± SD from 3 independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4 Cont.
<p>Overexpression of eGFP-<span class="html-italic">Rufy4</span> inhibits osteoclast formation and stimulates bone resorption. (<b>a</b>) Quantitative RT-PCR analysis of <span class="html-italic">Rufy4</span> mRNA expression in RAW-D cells expressing either eGFP or eGFP-<span class="html-italic">Rufy4</span>. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>b</b>) Western blot analysis of RAW-D cells expressing eGFP or eGFP-<span class="html-italic">Rufy4</span>. Cell lysates were subjected to western blot analysis using anti-GFP or anti-GAPDH antibodies. (<b>c</b>) TRAP staining of eGFP- and eGFP-<span class="html-italic">Rufy4</span>-overexpressing osteoclasts. eGFP- or eGFP-<span class="html-italic">Rufy4</span>-overexpressing RAW-D cells were stimulated with 100 ng/mL RANKL for 3 days and then fixed and stained for TRAP. Scale bar, 50 μm. (<b>d</b>) Number of TRAP-positive multinucleated cells. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>e</b>) Bone resorption area of eGFP or eGFP-<span class="html-italic">Rufy4</span>-overexpressing osteoclasts. Cells were seeded onto Osteo Assay Stripwell Plates containing 500 ng/mL RANKL. The images show the bone resorption area for each osteoclast. Scale bar, 20 μm. (<b>f</b>) The bone resorption area was determined using the ImageJ software. Data are presented as the mean ± SD from 3 independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>g</b>) Western blot analysis of eGFP or eGFP-<span class="html-italic">Rufy4</span>-overexpressing RAW-D cells cultured with 100 ng/mL RANKL for 2 or 3 days. Cells were harvested on the indicated day, and lysates were subjected to western blot analysis with specific antibodies against various proteins involved in osteoclast differentiation and actin polymerization, with GAPDH as a control. Representative immunoblots are shown, and the quantification results are presented as the mean ± SD from 3 independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>A RUN-domain deletion mutant of <span class="html-italic">Rufy4</span> stimulates osteoclast formation but impairs bone resorption. (<b>a</b>) Schematic representation of <span class="html-italic">Rufy4</span> wild-type and RUN-domain deletion mutants. RUN, OmpH, and FYVE domains are shown in red, yellow, and blue, respectively. Recombinant Rufy4 was expressed as an N-terminal 1 × FLAG-tagged fusion protein. (<b>b</b>) Western blot analysis of RAW-D cells expressing the control (empty vector), <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutants. Yellow asterisk indicates non-specific binding of antibody (<b>c</b>) Cell viability was assessed using the Cell Counting Kit-8. Data are presented as mean ± SD from 3 independent experiments. (<b>d</b>) TRAP staining of osteoclasts overexpressing the control, <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutants. Control, <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutant-overexpressing RAW-D cells were stimulated with 100 ng/mL RANKL for 3 days. The cells were then fixed and stained for TRAP. Scale bar, 100 μm. (<b>e</b>) Number of TRAP-positive multinucleated cells. Data are presented as mean ± SD from three independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>f</b>) Bone resorption areas of osteoclasts overexpressing control, <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutant, seeded onto Osteo Assay Stripwell Plates containing 500 ng/mL RANKL. The images show the bone resorption area for each osteoclast. Scale bar, 100 μm. (<b>g</b>) The bone resorption area was determined using the ImageJ software. Data are presented as mean ± SD from three independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>h</b>) The RUN-domain deletion mutant overexpressing osteoclasts exhibited a significantly increased expression of osteoclast marker genes. RAW-D cells overexpressing the control or <span class="html-italic">Rufy4</span> wild-type and RUN-domain deletion mutant were cultured with 100 ng/mL RANKL for 3 days. Following mRNA isolation, quantitative real-time PCR was performed. Data are presented as mean ± SD from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>i</b>) Western blot analysis of control (empty vector)-, <span class="html-italic">Rufy4</span> wild-type-, and RUN-domain deletion mutant-overexpressing RAW-D cells cultured with 100 ng/mL RANKL for 3 days. Cells were harvested on the indicated day, and lysates were subjected to western blot analysis with specific antibodies against various proteins involved in osteoclast differentiation and actin polymerization, with GAPDH as a control. Representative immunoblots are shown, and the quantification results are presented as mean ± SD from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5 Cont.
<p>A RUN-domain deletion mutant of <span class="html-italic">Rufy4</span> stimulates osteoclast formation but impairs bone resorption. (<b>a</b>) Schematic representation of <span class="html-italic">Rufy4</span> wild-type and RUN-domain deletion mutants. RUN, OmpH, and FYVE domains are shown in red, yellow, and blue, respectively. Recombinant Rufy4 was expressed as an N-terminal 1 × FLAG-tagged fusion protein. (<b>b</b>) Western blot analysis of RAW-D cells expressing the control (empty vector), <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutants. Yellow asterisk indicates non-specific binding of antibody (<b>c</b>) Cell viability was assessed using the Cell Counting Kit-8. Data are presented as mean ± SD from 3 independent experiments. (<b>d</b>) TRAP staining of osteoclasts overexpressing the control, <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutants. Control, <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutant-overexpressing RAW-D cells were stimulated with 100 ng/mL RANKL for 3 days. The cells were then fixed and stained for TRAP. Scale bar, 100 μm. (<b>e</b>) Number of TRAP-positive multinucleated cells. Data are presented as mean ± SD from three independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>f</b>) Bone resorption areas of osteoclasts overexpressing control, <span class="html-italic">Rufy4</span> wild-type, and RUN-domain deletion mutant, seeded onto Osteo Assay Stripwell Plates containing 500 ng/mL RANKL. The images show the bone resorption area for each osteoclast. Scale bar, 100 μm. (<b>g</b>) The bone resorption area was determined using the ImageJ software. Data are presented as mean ± SD from three independent experiments (** <span class="html-italic">p</span> &lt; 0.01). (<b>h</b>) The RUN-domain deletion mutant overexpressing osteoclasts exhibited a significantly increased expression of osteoclast marker genes. RAW-D cells overexpressing the control or <span class="html-italic">Rufy4</span> wild-type and RUN-domain deletion mutant were cultured with 100 ng/mL RANKL for 3 days. Following mRNA isolation, quantitative real-time PCR was performed. Data are presented as mean ± SD from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>i</b>) Western blot analysis of control (empty vector)-, <span class="html-italic">Rufy4</span> wild-type-, and RUN-domain deletion mutant-overexpressing RAW-D cells cultured with 100 ng/mL RANKL for 3 days. Cells were harvested on the indicated day, and lysates were subjected to western blot analysis with specific antibodies against various proteins involved in osteoclast differentiation and actin polymerization, with GAPDH as a control. Representative immunoblots are shown, and the quantification results are presented as mean ± SD from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p><span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts exhibit extensive F-actin assembly, whereas RUN-domain-deficient mutant-expressing osteoclasts show enhanced formation of protrusive structures. (<b>a</b>) RAW-D cells overexpressing the control (A–D, A′–D′), FLAG-tagged <span class="html-italic">Rufy4</span> wild-type (E–H, E′–H′), and FLAG-tagged <span class="html-italic">Rufy4</span> RUN-domain deletion mutant (I–L, I′–L′) were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed, stained with phalloidin and DAPI, and visualized using confocal microscopy. Control multinucleated osteoclasts exhibit F-actin assembly (A–D). <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts showed abundant F-actin assemblies, such as belts (E–H), whereas <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts showed elongated axial protrusive structures (I–L). Cell contour was traced by blue for Type 1 lines, yellow for Type 2 lines, and red for Type 3 lines, using phalloidin images. Scale bar, 20 μm. (<b>b</b>) The relative length of type 1 or type 2 plus 3 curves per whole cell perimeter was compared. The total number of analyzed cells was 54, 66, and 68 for control and <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts and <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts, respectively. All values are presented as the mean ± SD (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>c</b>) RAW-D cells overexpressing the <span class="html-italic">Rufy4</span> RUN domain deletion mutant were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed and stained with phalloidin (green) and α-tubulin antibody (magenta) and then visualized by confocal microscopy. Polymerization of F-actin (A, yellow arrowheads) and tubulin (C, white arrows) is observed. (B) Higher magnification image of the dotted square b in A. (D) Higher magnification image of the dotted square d in A. Scale bar, 20 μm. (<b>d</b>) RAW-D cells overexpressing the <span class="html-italic">Rufy4</span> RUN domain deletion mutant were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed and stained with phalloidin (green), LAMP1 (magenta), and DAPI (blue) and visualized using confocal microscopy (A) (B–E) Higher magnification images of the dotted square (b–e) in (A), respectively. LAMP1 was observed at the cell periphery and tips of the moving tail and bridge (arrows). Scale bar, 20 μm. (<b>e</b>) RAW-D cells overexpressing the control (A), FLAG-tagged <span class="html-italic">Rufy4</span> wild-type (B), and FLAG-tagged <span class="html-italic">Rufy4</span> RUN-domain deletion mutant (C) were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed, stained with phalloidin (green), anti-vinculin antibody (magenta), and DAPI (blue), and visualized using confocal microscopy. High-magnification images of the dotted squares are shown on the left side of each image. Scale bar, 20 μm. (<b>f</b>) Blot analyses of GTP-bound RhoA in osteoclasts overexpressing control, FLAG-tagged <span class="html-italic">Rufy4</span> wild-type, and FLAG-tagged <span class="html-italic">Rufy4</span> RUN-domain deletion mutant. The cells were cultured with 100 ng/mL RANKL. After 3 days, cell extracts with the same amounts of protein were subjected to Rho GTPase pull-down experiments. GDP and GTPγS were used as the negative and positive controls, respectively. (<b>g</b>) Western blot analysis of control-, <span class="html-italic">Rufy4</span> wild-type-, and RUN-domain deletion mutant-overexpressing RAW-D cells cultured with 100 ng/mL RANKL for 2 or 3 days. Cells were harvested, and lysates were subjected to western blot analysis with specific antibodies against various proteins involved in actin polymerization and podosome formation, with GAPDH as a control. Representative immunoblots are shown, and the quantification results are presented as mean ± SD from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6 Cont.
<p><span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts exhibit extensive F-actin assembly, whereas RUN-domain-deficient mutant-expressing osteoclasts show enhanced formation of protrusive structures. (<b>a</b>) RAW-D cells overexpressing the control (A–D, A′–D′), FLAG-tagged <span class="html-italic">Rufy4</span> wild-type (E–H, E′–H′), and FLAG-tagged <span class="html-italic">Rufy4</span> RUN-domain deletion mutant (I–L, I′–L′) were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed, stained with phalloidin and DAPI, and visualized using confocal microscopy. Control multinucleated osteoclasts exhibit F-actin assembly (A–D). <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts showed abundant F-actin assemblies, such as belts (E–H), whereas <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts showed elongated axial protrusive structures (I–L). Cell contour was traced by blue for Type 1 lines, yellow for Type 2 lines, and red for Type 3 lines, using phalloidin images. Scale bar, 20 μm. (<b>b</b>) The relative length of type 1 or type 2 plus 3 curves per whole cell perimeter was compared. The total number of analyzed cells was 54, 66, and 68 for control and <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts and <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts, respectively. All values are presented as the mean ± SD (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>c</b>) RAW-D cells overexpressing the <span class="html-italic">Rufy4</span> RUN domain deletion mutant were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed and stained with phalloidin (green) and α-tubulin antibody (magenta) and then visualized by confocal microscopy. Polymerization of F-actin (A, yellow arrowheads) and tubulin (C, white arrows) is observed. (B) Higher magnification image of the dotted square b in A. (D) Higher magnification image of the dotted square d in A. Scale bar, 20 μm. (<b>d</b>) RAW-D cells overexpressing the <span class="html-italic">Rufy4</span> RUN domain deletion mutant were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed and stained with phalloidin (green), LAMP1 (magenta), and DAPI (blue) and visualized using confocal microscopy (A) (B–E) Higher magnification images of the dotted square (b–e) in (A), respectively. LAMP1 was observed at the cell periphery and tips of the moving tail and bridge (arrows). Scale bar, 20 μm. (<b>e</b>) RAW-D cells overexpressing the control (A), FLAG-tagged <span class="html-italic">Rufy4</span> wild-type (B), and FLAG-tagged <span class="html-italic">Rufy4</span> RUN-domain deletion mutant (C) were cultured on glass coverslips with 100 ng/mL RANKL. After 3 days, the cells were fixed, stained with phalloidin (green), anti-vinculin antibody (magenta), and DAPI (blue), and visualized using confocal microscopy. High-magnification images of the dotted squares are shown on the left side of each image. Scale bar, 20 μm. (<b>f</b>) Blot analyses of GTP-bound RhoA in osteoclasts overexpressing control, FLAG-tagged <span class="html-italic">Rufy4</span> wild-type, and FLAG-tagged <span class="html-italic">Rufy4</span> RUN-domain deletion mutant. The cells were cultured with 100 ng/mL RANKL. After 3 days, cell extracts with the same amounts of protein were subjected to Rho GTPase pull-down experiments. GDP and GTPγS were used as the negative and positive controls, respectively. (<b>g</b>) Western blot analysis of control-, <span class="html-italic">Rufy4</span> wild-type-, and RUN-domain deletion mutant-overexpressing RAW-D cells cultured with 100 ng/mL RANKL for 2 or 3 days. Cells were harvested, and lysates were subjected to western blot analysis with specific antibodies against various proteins involved in actin polymerization and podosome formation, with GAPDH as a control. Representative immunoblots are shown, and the quantification results are presented as mean ± SD from three independent experiments (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>A RUN-domain deletion mutant of <span class="html-italic">Rufy4</span> exhibits long-lasting bridge and tail formation. RAW-D cells expressing the control (empty vector, (<b>a</b>)), <span class="html-italic">Rufy4</span> wild-type (<b>b</b>), and RUN-domain deletion mutant (<b>c</b>) were cultured on 35 mm plastic culture dishes with 100 ng/mL RANKL for 48 h, and live-cell phase contrast imaging was performed every 6 min for 12 h. Time-lapse images showing bridges and tails (arrows) are representative of five independent experiments. (<b>d</b>) The formation and duration of bridges and tails were quantified using live-cell video data. The total numbers of analyzed cells for the duration time of the cytokinetic bridge were 35, 30, and 65; for the duration time of moving tail, they were 63, 45, and 66; and for duration time of multinuclear cell bridge were 35, 45, and 35 for control and <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts, and <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts, respectively. Data are presented as mean ± SD (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7 Cont.
<p>A RUN-domain deletion mutant of <span class="html-italic">Rufy4</span> exhibits long-lasting bridge and tail formation. RAW-D cells expressing the control (empty vector, (<b>a</b>)), <span class="html-italic">Rufy4</span> wild-type (<b>b</b>), and RUN-domain deletion mutant (<b>c</b>) were cultured on 35 mm plastic culture dishes with 100 ng/mL RANKL for 48 h, and live-cell phase contrast imaging was performed every 6 min for 12 h. Time-lapse images showing bridges and tails (arrows) are representative of five independent experiments. (<b>d</b>) The formation and duration of bridges and tails were quantified using live-cell video data. The total numbers of analyzed cells for the duration time of the cytokinetic bridge were 35, 30, and 65; for the duration time of moving tail, they were 63, 45, and 66; and for duration time of multinuclear cell bridge were 35, 45, and 35 for control and <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts, and <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts, respectively. Data are presented as mean ± SD (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7 Cont.
<p>A RUN-domain deletion mutant of <span class="html-italic">Rufy4</span> exhibits long-lasting bridge and tail formation. RAW-D cells expressing the control (empty vector, (<b>a</b>)), <span class="html-italic">Rufy4</span> wild-type (<b>b</b>), and RUN-domain deletion mutant (<b>c</b>) were cultured on 35 mm plastic culture dishes with 100 ng/mL RANKL for 48 h, and live-cell phase contrast imaging was performed every 6 min for 12 h. Time-lapse images showing bridges and tails (arrows) are representative of five independent experiments. (<b>d</b>) The formation and duration of bridges and tails were quantified using live-cell video data. The total numbers of analyzed cells for the duration time of the cytokinetic bridge were 35, 30, and 65; for the duration time of moving tail, they were 63, 45, and 66; and for duration time of multinuclear cell bridge were 35, 45, and 35 for control and <span class="html-italic">Rufy4</span> wild-type-overexpressing osteoclasts, and <span class="html-italic">Rufy4</span> RUN-domain deletion mutant-overexpressing osteoclasts, respectively. Data are presented as mean ± SD (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
22 pages, 5573 KiB  
Article
Circular RNA hsa_circ_0008726 Targets the hsa-miR-206-3p/KLF4 Axis to Modulate 4,4′-Methylene Diphenyl Diisocyanate-Glutathione Conjugate-Induced Chemokine Transcription in Macrophages
by Chen-Chung Lin, Brandon F. Law and Justin M. Hettick
Cells 2024, 13(20), 1725; https://doi.org/10.3390/cells13201725 - 18 Oct 2024
Viewed by 983
Abstract
Exposure to 4,4′-methylene diphenyl diisocyanate (MDI) in the workplace may lead to the development of occupational asthma (OA). However, the specific mechanism(s) by which MDI induces OA are poorly understood. Previous reports have demonstrated that MDI and MDI-glutathione (GSH) conjugate exposure downregulates endogenous [...] Read more.
Exposure to 4,4′-methylene diphenyl diisocyanate (MDI) in the workplace may lead to the development of occupational asthma (OA). However, the specific mechanism(s) by which MDI induces OA are poorly understood. Previous reports have demonstrated that MDI and MDI-glutathione (GSH) conjugate exposure downregulates endogenous human/murine (hsa/mmu)-microRNA(miR)-206-3p, resulting in the activation of mmu/hsa-miR-206-3p-regulated signaling pathways in macrophages. Circular RNAs (circRNAs) regulate many important biological processes by targeting endogenous miRs; however, whether MDI/MDI-GSH exposure may influence circRNA expressions is unknown. Several circRNAs have been identified that regulate hsa-miR-206-3p. We hypothesize that MDI-GSH conjugate exposure induces endogenous circRNA(s) to regulate hsa-miR-206-3p in macrophages. The expression of candidate hsa-miR-206-3p-binding circRNAs was determined from MDI-GSH conjugate-treated differentiated THP-1 macrophages using RT-qPCR. MDI-GSH exposures induced hsa_circ_0008726 and its host gene transcript DNAJB6, whereas other circRNA(s) examined were either not detected or unchanged. RNA-induced silencing complex-immunoprecipitation (RISC-IP) experiments confirm that hsa-miR-206-3p can bind to hsa_circ_0008726. The expressions of endogenous hsa-miR-206-3p, hsa-miR-206-3p-regulated KLF4, and KLF4-activated M2 macrophage-associated markers and chemokines were up-/down-regulated by transfection of hsa_circ_0008726 siRNAs or hsa_circ_0008726 overexpression plasmid in macrophages, respectively. These results suggest MDI-GSH exposure downregulates hsa-miR-206-3p via induction of endogenous hsa_circ_0008726/DNAJB6, resulting in the upregulation of hsa-miR-206-3p-mediated regulations in macrophages. Full article
(This article belongs to the Special Issue Advances in the Biogenesis, Biology, and Functions of Noncoding RNAs)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><b>MDI-GSH conjugate treatment induces endogenous circRNA <span class="html-italic">hsa_circ_0008726</span> in differentiated/enhanced THP-1 macrophages.</b> Total RNA was isolated from the indicated MDI-GSH conjugate-treated differentiated/enhanced THP-1 macrophages by the <span class="html-italic">mirVana</span><sup>™</sup> miR isolation kit, reverse transcribed, and subjected to SYBR green-based or TaqMan stem-loop miR RT-qPCR. Endogenous miR/circRNA expressions of (<b>A</b>) <span class="html-italic">hsa-miR-206-3p</span>, (<b>B</b>) <span class="html-italic">hsa_circ_0000199</span>, (<b>C</b>) <span class="html-italic">hsa_circ_0001264</span>, (<b>D</b>) <span class="html-italic">hsa_circ_0001982</span>, (<b>E</b>) <span class="html-italic">hsa_circ_0004662</span>, (<b>F</b>) <span class="html-italic">hsa_circ_0007428</span>, (<b>G</b>) <span class="html-italic">hsa_circ_0008726</span>, (<b>H</b>) <span class="html-italic">hsa_circ_0056618</span>, (<b>I</b>) <span class="html-italic">hsa_circ_0057558</span>, (<b>J</b>) <span class="html-italic">hsa_circ_0058141</span>, and (<b>K</b>) <span class="html-italic">hsa_circ_0072088</span> were determined 24 h after MDI-GSH conjugate treatments (N = 3; bars, SEM). MDI: 4,4′-methylene diphenyl diisocyanate. GSH: Glutathione. (* <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><b>Circular RNA <span class="html-italic">hsa_circ_0008726</span> is presented in THP-1 macrophages, and MDI-GSH conjugates upregulate endogenous <span class="html-italic">hsa_circ_0008726</span> parental host gene transcript <span class="html-italic">DNAJB6</span>.</b> (<b>A</b>) Characteristics of <span class="html-italic">hsa_circ_0008726</span> obtained from the Circular RNA Interactome database. (<b>B</b>) Illustration shows exon numbers and designed convergent and divergent primer sites on the mature <span class="html-italic">DNAJB6</span> transcripts. RNAse R degrades linear RNA species, including the <span class="html-italic">DNAJB6</span> transcript. CircRNA <span class="html-italic">hsa_circ_0008726</span> is back spliced from exon 3–5 of the <span class="html-italic">DNAJB6</span> transcript. (<b>C</b>) Total RNA was isolated from THP-1 macrophages by the <span class="html-italic">mirVana<sup>™</sup></span> miR isolation kit and treated with or without RNAse R, further purified using the <span class="html-italic">mirVana<sup>™</sup></span> miR isolation kit, reverse transcribed, and subjected to RT-PCR using convergent or divergent primers. RT-NTC: Templates from a cDNA synthesis reaction without adding reverse transcriptase. PCR-NTC: Use only water to replace cDNA templates during PCR reaction. (<b>D</b>) Total RNA was isolated from MDI-GSH-treated differentiated/enhanced THP-1 macrophages at indicated concentrations for 24 h by the <span class="html-italic">mirVana<sup>™</sup></span> miR isolation kit, reverse transcribed, and subjected to TaqMan RT-qPCR assays. Endogenous levels of <span class="html-italic">DNAJB6</span> were determined at 24 h after MDI-GSH conjugate treatment (N = 3; bars, SEM). MDI: 4,4′-methylene diphenyl diisocyanate. GSH: Glutathione (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p><b>Human circular RNA <span class="html-italic">hsa_circ_0008726</span> is a target of <span class="html-italic">hsa-miR-206-3p.</span></b> (<b>A</b>) Alignment of the <span class="html-italic">hsa_circ_0008726</span> sequence regions of potential <span class="html-italic">hsa-miR-206-3p</span> binding sites. (<b>B</b>) Differentiated/enhanced THP-1 macrophages were transfected with 25 nM of indicated miR-mimic or nontargeting miR-mimic control (miR-mimic-Ctl) for 24 h. The cells were collected and immunoprecipitated using the panAGO or isotype IgG antibody after 24 h transfection. RNA was isolated, and the fold enrichment of <span class="html-italic">hsa_circ_0008726</span> was measured (N = 3; bars, SEM). (<b>C</b>,<b>D</b>) THP-1 macrophages were transfected with 25 nM of either miR-mimic/inhibitor-206-3p, miR-mimic/inhibitor-381-3p, or nontargeting miR-mimic/inhibitor control for 24 h. Total RNA was isolated from the indicated miR-mimics/inhibitors transfected THP-1 macrophages by the <span class="html-italic">mirVana<sup>™</sup></span> miR isolation kit, reverse transcribed, and subjected to RT-qPCR. The endogenous <span class="html-italic">hsa_circ_0008726</span> levels from indicated (<b>C</b>) miR-mimics or (<b>D</b>) miR-inhibitors transfected THP-1 macrophages were determined by SYBR Green RT-qPCR assays (N = 3; bars, SEM). (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p><b>Transfection of <span class="html-italic">hsa_circ_0008726</span> siRNA knocks down endogenous <span class="html-italic">hsa_circ_0008726</span> levels and upregulates <span class="html-italic">hsa-miR-206-3p</span> in differentiated/enhanced THP-1 macrophages.</b> Differentiated/enhanced THP-1 macrophages were transfected with 25 nM of either si-<span class="html-italic">hsa_circ_0008726</span>#1, si-<span class="html-italic">hsa_circ_0008726</span>#2 siRNA, or nontargeting siRNA control (siCtl). After 24 h, the endogenous levels of (<b>A</b>) <span class="html-italic">hsa_circ_0008726</span> and (<b>B</b>) <span class="html-italic">hsa-miR-206-3p</span> were measured by RT-qPCR (N = 3; bars, SEM). (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p><b>CircRNA <span class="html-italic">hsa_circ_0008726</span> as a downstream effector to MDI-GSH conjugate exposure for regulating <span class="html-italic">hsa-miR-206-3p/KLF4</span> and KLF4-mediated M2 macrophage-associated markers and chemokines in macrophages.</b> Differentiated/enhanced THP-1 macrophages were transfected with 25 nM of either si-<span class="html-italic">hsa_circ_0008726</span>#1 or nontargeting siRNA control (siCtl) for 24 h, followed by treatment either with or without 10 µM MDI-GSH conjugate for 24 h. Total RNA was isolated from macrophages with indicated treatments/transfections by the <span class="html-italic">mirVana<sup>™</sup></span> miR isolation kit, reverse transcribed, and subjected to SYBR green-based or TaqMan stem-loop miR RT-qPCR. The endogenous levels of (<b>A</b>) <span class="html-italic">hsa_circ_0008726</span> and (<b>B</b>) <span class="html-italic">hsa-miR-206-3p</span> as well as the M2 macrophage-associated transcription factor (<b>C</b>) <span class="html-italic">KLF4</span>, markers (<b>D</b>) <span class="html-italic">CD206</span>, (<b>E</b>) <span class="html-italic">TGM2</span>, (<b>F</b>) <span class="html-italic">CCL17</span>, (<b>G</b>) <span class="html-italic">CCL22</span>, and (<b>H</b>) <span class="html-italic">CCL24</span> mRNA levels were determined in total RNA isolated from macrophages as indicated treatments (N = 3; bars, SEM). (* <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 when compared to vehicle-treated macrophages with transfection of or nontargeting siRNA control (siCtl); <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001, when compared to macrophages treated with 10 µM MDI-GSH conjugate as well as with transfection of indicated either si-<span class="html-italic">hsa_circ_0008726</span>#1 or nontargeting siRNA control (siCtl)).</p>
Full article ">Figure 6
<p><b>Circular RNA <span class="html-italic">hsa_circ_0008726</span> overexpression increases M2 macrophage associate markers and chemokines in differentiated/enhanced THP-1 macrophages.</b> Differentiated/enhanced THP-1 macrophages were transfected with 2.5 µg of either pcDNA3.1<sup>(+)</sup>_Circ_Mini-<span class="html-italic">hsa_circ_0008726</span> or pcDNA3.1<sup>(+)</sup>_Circ_Mini vector plasmids for 48 h. Total RNA was isolated from plasmids transfected THP-1 macrophages by the <span class="html-italic">mirVana<sup>™</sup></span> miR isolation kit, reverse transcribed, and subjected to SYBR green or TaqMan RT-qPCR. The transgene of (<b>A</b>) <span class="html-italic">hsa_circ_0008726</span> and (<b>B</b>) <span class="html-italic">hsa-miR-206-3p</span> as well as the endogenous M2 macrophage-associated markers (<b>C</b>) <span class="html-italic">KLF4</span>, (<b>D</b>) <span class="html-italic">CD206</span>, (<b>E</b>) <span class="html-italic">TGM2</span>, (<b>F</b>) <span class="html-italic">CCL17</span>, (<b>G</b>) <span class="html-italic">CCL22</span>, and (<b>H</b>) <span class="html-italic">CCL24</span> mRNA levels were determined by RT-qPCR (N = 3; bars, SEM). (* <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><b>CircRNA <span class="html-italic">hsa_circ_0008726</span> plays an important role for the secretion of chemokines CCL17, CCL22, and CCL24 and regulates T-cell and eosinophil chemotaxis/migration in macrophages.</b> Cell-free conditioned media were obtained from THP-1 macrophages transfected with either the <span class="html-italic">hsa_circ_0008726</span> overexpression plasmid or the empty vector for 48 h. The secreted protein levels of (<b>A</b>) CCL17, (<b>B</b>) CCL22, and (<b>C</b>) CCL24 in conditioned media from either <span class="html-italic">hsa_circ_0008726</span> overexpressed THP-1 macrophages or empty vector transfected THP-1 macrophages were determined by ELISA according to the manufacturer’s instructions. The isolated conditioned media were used as chemoattractants to attract (<b>D</b>) Jurkat T-cell clone E6-1 or differentiated (<b>E</b>) HL-60 C_15 eosinophils. T-cell and eosinophil migration responding to the conditioned media was measured after 6 h. Percent of cells migrated towards the bottom chamber are shown (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 8
<p><b>Proposed mechanisms by which MDI-GSH conjugate exposure induces M2 macrophage-associated markers and chemokine <span class="html-italic">CCL17</span>, <span class="html-italic">CCL22</span>, and <span class="html-italic">CCL24</span> via <span class="html-italic">hsa_circ_0008726/hsa-miR-206-3p</span>-regulated KLF4 activation in macrophages.</b> MDI: 4,4′-methylene diphenyl diisocyanate; TFs: transcription factors; CDS: coding sequences; KLF4: Krüppel-like factor 4. Note: Some illustrated schematics were obtained from <span class="html-italic">motifolio</span> templates (Motifolio Inc., Ellicott City, MD, USA).</p>
Full article ">
18 pages, 5966 KiB  
Article
Co-Regulation Mechanism of Host p53 and Fos in Transcriptional Activation of ILTV Immediate-Early Gene ICP4
by Zheyi Liu, Xuefeng Li, Lu Cui, Shufeng Feng, Zongxi Han, Yu Zhang, Shengwang Liu and Hai Li
Microorganisms 2024, 12(10), 2069; https://doi.org/10.3390/microorganisms12102069 - 16 Oct 2024
Viewed by 748
Abstract
Infectious laryngotracheitis virus (ILTV) exhibits a cascade expression pattern of encoded genes, and ICP4 is the only immediate-early gene of ILTV, which plays a crucial role in initiating the subsequent viral genes. Therefore, studying the transcriptional regulation mechanism of ICP4 holds promise for [...] Read more.
Infectious laryngotracheitis virus (ILTV) exhibits a cascade expression pattern of encoded genes, and ICP4 is the only immediate-early gene of ILTV, which plays a crucial role in initiating the subsequent viral genes. Therefore, studying the transcriptional regulation mechanism of ICP4 holds promise for effectively blocking ILTV infection and spread. Host transcriptional factors p53 and Fos are proven to regulate a variety of viral infections, and our previous studies have demonstrated their synergistic effects in regulating ILTV infection. In this study, we constructed eukaryotic expression vectors for p53 and Fos as well as their specific siRNAs and transfected them into a chicken hepatoma cell line. The results showed that knocking down p53 or Fos significantly inhibited ICP4 transcription, while overexpressing p53 or Fos had an opposite effect. A further CoIP and ChIP-qPCR assay suggested p53 and Fos physically interacted with each other, and jointly bound to the upstream transcriptional regulatory region of ICP4. To elucidate the specific mechanisms of p53 and Fos in regulating ICP4 transcription, we designed p53 and Fos protein mutants by mutating their DNA binding domains, which significantly reduced their binding ability to DNA without affecting their interaction. The results showed that Fos directly bound to the promoter region of ICP4 as a binding target of p53, and the p53–Fos protein complex acted as a transcriptional co-regulator of ICP4. Studying the transcriptional process and regulatory pattern of ICP4 is of great significance for understanding the molecular mechanism of ILTV infection, and thus for finding effective methods to control and prevent it. Full article
(This article belongs to the Special Issue State-of-the-Art Veterinary Microbiology in China (2023, 2024))
Show Figures

Figure 1

Figure 1
<p>Effects of p53 and Fos overexpression and knockdown on the transcription of <span class="html-italic">ICP4.</span> (<b>A</b>–<b>E</b>) 24 h after instantaneous transfection of pCAG-p53-Flag and pCAG-Fos-HA into LMH cells, the overexpression efficiency of p53 and Fos was analyzed on mRNA level by RT-qPCR (<b>A</b>,<b>B</b>) and on protein level by Western blot (<b>C</b>,<b>D</b>) and immunofluorescence (<b>E</b>). The scale bar indicates 150 µm. (<b>F</b>,<b>G</b>) The knockdown efficiency of p53 and Fos was analyzed on mRNA level by RT-qPCR 24 h after sip53 and siFos transfection into LMH cells. (<b>H</b>,<b>I</b>) 24 h after transfection with pCAG-Flag, pCAG-p53-Flag or pCAG-HA, pCAG-Fos-HA, LMH cells were infected with ILTV (MOI = 1). The transcription level of ILTV immediate-early gene <span class="html-italic">ICP4</span> was detected at the indicated time points by absolute quantitative PCR with standard curve method. (<b>J</b>,<b>K</b>) 24 h after transfection with sicontrol, sip53 or siFos, LMH cells were infected with ILTV (MOI = 1). The transcription level of ILTV immediate-early gene <span class="html-italic">ICP4</span> was quantitatively detected at the indicated time points. Data in (<b>A</b>,<b>B</b>,<b>F</b>–<b>K</b>) are presented as the mean ± SD, <span class="html-italic">n</span> = 3. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate the levels of significance.</p>
Full article ">Figure 2
<p>p53 promotes <span class="html-italic">Fos</span> transcription and directly binds to the promoter region of <span class="html-italic">Fos</span>. (<b>A</b>,<b>B</b>) Effect of overexpression or knockdown of p53 on <span class="html-italic">Fos</span> transcription was assayed by RT-qPCR. (<b>C</b>,<b>D</b>) Effect of overexpression or knockdown of Fos on <span class="html-italic">p53</span> transcription was assayed by RT-qPCR. (<b>E</b>) Prediction of the putative p53 DNA binding sites within the 2000 bp upstream region of <span class="html-italic">Fos</span> gene using our previous ChIP-sequencing data. (<b>F</b>) The binding level of p53 on the predicted sites was validated by ChIP-qPCR. (<b>G</b>) The transcriptional activity of p53 was assayed by dual-luciferase reporter assay. Data in (<b>A</b>–<b>D</b>,<b>F</b>,<b>G</b>) are represented as mean ± standard deviation, <span class="html-italic">n</span> = 3. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate the levels of significance.</p>
Full article ">Figure 3
<p>p53 indirectly binds to <span class="html-italic">ICP4</span> promoter as a transcriptional regulator. (<b>A</b>) Prediction of the putative Fos DNA binding sites within the 2000 bp upstream region of <span class="html-italic">ICP4</span> gene using Jasper database. (<b>B</b>) The binding level of Fos on the predicted sites was validated by ChIP-qPCR. (<b>C</b>) The transcriptional activity of Fos was assayed by dual-luciferase reporter assay. (<b>D</b>) Co-IP of p53 and Fos in LMH cells with or without ILTV infection (MOI = 1) using antibodies specifically recognizing HA or Flag. IP: Immunoprecipitation; IB: Immunoblotting. (<b>E</b>) The binding level of p53 on the predicted sites was validated by ChIP-qPCR. Data in (<b>B</b>,<b>C</b>,<b>E</b>) are represented as mean ± standard deviation, <span class="html-italic">n</span> = 3. ** <span class="html-italic">p</span> &lt; 0.01 indicates the levels of significance.</p>
Full article ">Figure 4
<p>p53 directly binding with DNA is not necessary for transcriptional regulation of <span class="html-italic">ICP4.</span> (<b>A</b>) Mutation of chicken p53 at the conserved region of DNA binding domain. (<b>B</b>) 24 h after instantaneous transfection of pCAG-p53-Flag and pCAG-pm1-Flag into LMH cells, the overexpression efficiency was analyzed by Western blot. Tubulin was used as the inner control. (<b>C</b>) The binding of wtp53 and pm1 to the classical p53 target genes was detected by ChIP-qPCR. (<b>D</b>) The effects of wtp53 and pm1 on the transcription of classical p53 target genes were assayed by RT-qPCR. (<b>E</b>) Co-IP of p53 and Fos in LMH cells with or without p53 mutation using antibodies specifically recognizing HA or Flag. IP: Immunoprecipitation; IB: Immunoblotting. (<b>F</b>) The binding of wtp53 and pm1 to <span class="html-italic">ICP4</span> promoter was detected by ChIP-qPCR. (<b>G</b>) The binding of Fos to <span class="html-italic">ICP4</span> promoter upon co-overexpression of wtp53 or pm1 was detected by ChIP-qPCR. (<b>H</b>) The transcriptional activities of wtp53 and pm1 were assayed by dual-luciferase reporter assay. (<b>I</b>) LMH cells were transfected with pCAG-Flag, pCAG-p53-Flag, or pCAG-pm1-Flag, and 24 h later infected with ILTV (MOI = 1). The transcription level of <span class="html-italic">ICP4</span> was quantitatively detected by absolute quantitative PCR with standard curve method. Data in (<b>C</b>,<b>D</b>,<b>F</b>–<b>I</b>) are represented as mean ± standard deviation, <span class="html-italic">n</span> = 3. ** <span class="html-italic">p</span> &lt; 0.01 indicate the levels of significance.</p>
Full article ">Figure 5
<p>Fos directly binding with DNA is required for the transcriptional regulation of <span class="html-italic">ICP4</span> by p53. (<b>A</b>) Mutation of chicken Fos at the conserved region of DNA binding domain. (<b>B</b>) 24 h after instantaneous transfection of pCAG-Fos-HA and pCAG-Fm1-HA into LMH cells, the overexpression efficiency was analyzed by Western blot. Tubulin was used as the inner control. (<b>C</b>) The binding of wide-type Fos and Fm1 to the classical Fos target genes was detected by ChIP-qPCR. (<b>D</b>) The effect of wide-type Fos and Fm1 on the transcription of classical Fos target genes was assayed by RT-qPCR. (<b>E</b>) Co-IP of p53 and Fos in LMH cells with or without Fos mutation using antibodies specifically recognizing HA or Flag. IP: Immunoprecipitation; IB: Immunoblotting. (<b>F</b>) The binding of Fos and Fm1 to <span class="html-italic">ICP4</span> promoter was detected by ChIP-qPCR. (<b>G</b>) The binding of p53 to <span class="html-italic">ICP4</span> promoter upon co-overexpression of wide-type Fos and Fm1 was detected by ChIP-qPCR. (<b>H</b>) The transcriptional activities of wide-type Fos and Fm1 were assayed by dual-luciferase reporter assay. (<b>I</b>) LMH cells were transfected with pCAG-HA, pCAG-Fos-HA, or pCAG-Fm1-HA, and 24 h later infected with ILTV (MOI = 1). The transcription level of <span class="html-italic">ICP4</span> was quantitatively detected by absolute quantitative PCR with standard curve method. Data in (<b>C</b>,<b>D</b>,<b>F</b>–<b>I</b>) are represented as mean ± standard deviation, <span class="html-italic">n</span> = 3. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 indicate the levels of significance.</p>
Full article ">Figure 6
<p>Diagram of host p53 and Fos in transcriptional activation of ILTV immediate-early gene <span class="html-italic">ICP4</span>.</p>
Full article ">Figure 7
<p>Prediction of the interaction between p53 monomer and Fos monomer.</p>
Full article ">Figure 8
<p>Prediction of interaction between p53 tetramer and Fos monomer.</p>
Full article ">
15 pages, 2987 KiB  
Article
MiR-181a Negatively Regulates Claudin-3 to Facilitate Lateolabrax maculatus Iridovirus Replication in Lateolabrax maculatus Astroglia Cells
by Yanping Ma, Jingjing Xu, Le Hao, Gang Wang, Wen Huang and Zhenxing Liu
Viruses 2024, 16(10), 1589; https://doi.org/10.3390/v16101589 - 9 Oct 2024
Viewed by 746
Abstract
Lateolabrax maculatus iridovirus (LMIV) is a variant strain of red sea bream iridovirus (RSIV), causing serious economic losses in aquaculture. Claudins (CLDNs) are major components of tight junctions (TJs) forming an important line of defense against pathogens. Our pilot miRNA-mRNA joint analysis indicated [...] Read more.
Lateolabrax maculatus iridovirus (LMIV) is a variant strain of red sea bream iridovirus (RSIV), causing serious economic losses in aquaculture. Claudins (CLDNs) are major components of tight junctions (TJs) forming an important line of defense against pathogens. Our pilot miRNA-mRNA joint analysis indicated the degradation of CLDN3, as well as its interaction with miR-181a during LMIV infection. To elucidate the miR-181a/CLDN3/LMIV interactions, in vitro assays were carried out on LMB-L cells. We first confirmed that LMIV infection could decrease the expression of CLDN3, accompanied by the enhancement of permeability, suggesting the dysfunction of TJs. Contrary to the inhibition of CLDN3, the activation of miR-181a was proved, presenting a negative correlation between miR-181a and CLDN3 (Pearson r = −0.773 and p < 0.01). In addition, the influence of CLDN3 on LMIV replication was analyzed by knockdown and over-expression of CLDN3. When CLDN3 was silenced in LMB-L cells with siCLDN3-623 at 9 days post transfection (dpt), LMIV copies and titers were significantly up-regulated by 1.59-fold and 13.87-fold, respectively. By contrast, LMIV replication in LMB-L cells was reduced by 60% and 71%, post transfection with pcDNA3.1-CLDN3 over-expressed plasmid at 6 dpt and 9 dpt, respectively. Ultimately, the regulatory relationship between miR-181a and CLDN3 was further validated by dual luciferase reporter assays. Taking into account the above-described results, we proposed a “miR-181a/CLDN3/LMIV” regulatory relationship. This study provides a new insight for understanding the mechanism of LMIV replication. Full article
(This article belongs to the Section Animal Viruses)
Show Figures

Figure 1

Figure 1
<p>Analysis of spotted sea bass CLDN3 protein sequences. (<b>a</b>) Structural comparison of CLDN3 was performed between spotted sea bass and mouse by online swiss-model. Typical domains are indicated by arrows. (<b>b</b>) Phylogenetic trees of full-length CLDN3s were constructed using the neighbor-joining procedure in MEGA software 11.0. The numbers on the forks indicate the percentage of bootstrap support from 1000 replicates. Genbank accession numbers are given in front of species names. Spotted sea bass CLDN3 is highlighted by the blank triangle.</p>
Full article ">Figure 2
<p>Expression of CLDN3, miR-181a, and replication of LMIV. CLDN3 expression was down-regulated and miR-181a expression was up-regulated in LMB-L cells post LMIV infection. The non-infected cells (0 d) served as negative controls. Total RNA, including mRNA and microRNA, was collected for expression assessment of CLDN3 and miR-181a at different time points; meanwhile, the DNA of LMIV-infected cells was used for the calculation of LMIV copies. (<b>a</b>) Virus copy numbers increased in LMB-L cells after exposure to LMIV. (<b>b</b>,<b>c</b>) The relative expression of CLDN3 (<b>b</b>) and miR-181a (<b>c</b>) was determined in LMIV-infected LMB-L cells, normalized to RNA-polymerase II or the U6 gene, respectively, exhibiting an inverse trend between the miR-181a levels and the expression of CLDN3. The data are presented as a mean ± SD (<span class="html-italic">n</span> = 9); ns, not significant, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Over-expression of CLDN3 inhibited LMIV replication in LMB-L cells. (<b>a</b>) The relative expression of CLDN3 in pcDNA3.1-CLDN3-transfected cells, normalized to that in pcDNA3.1-transfected cells, was assessed at 12 h.p.t, 24 h.p.t, and 48 h.p.t. (<b>b</b>) Over-expression analysis of CLDN3 in pcDNA3.1-CLDN3 plasmid-transfected cells by indirect immunofluorescence and Western blotting (Scale bar = 100 μm). (<b>c</b>) CPE induced by LMIV infection was mitigated by CLDN3 over-expression, with cell morphology and cytoplasmic vacuolization changed. Scale bars represent 100 μm; Black arrow indicates infected-detached cells, White arrow indicates infeted-loose cells. (<b>d</b>) After 48 h of transfection with pcDNA3.1-CLDN3 or pcDNA3.1 (control), transfected cells were exposed to LMIV and viral loads were quantified by the absolute quantitative qPCR method at 1, 3, 6, and 9 d.p.i. Over-expression of CLDN3 significantly inhibited LMIV replication in LMB-L cells at 6 d.p.i. and 9 d.p.i. The values are shown as means ± SD (<span class="html-italic">n</span> = 9) and analyzed by one-way ANOVA test; ns, not significant, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Knockdown of CLDN3 using siRNA facilitated LMIV replication in LMB-L cells. (<b>a</b>) The relative expression of CLDN3 was detected post siCLDN3-176 and siCLDN3-623 transfection in LMB-L cells, normalized to siCtrl-transfected cells. SiCLDN3-623 led to lower CLDN3 expression levels, compared with siCLDN3-176. (<b>b</b>) Knockdown analysis of CLDN3 in siRNA-transfected cells by indirect immunofluorescence and Western blotting (Scale bar = 100 μm). (<b>c</b>) CPE induced by LMIV infection was aggravated by siCLDN3-623 transfection, with cell morphology and cytoplasmic vacuolization changed. Scale bars represent 100 μm; Black arrow, infected-detached cells. (<b>d</b>) LMIV copies were quantified post knockdown of CLDN3. Forty-eight hours after transfection with siCLDN3-623, the cells were subjected to LMIV infection, followed by virus quantification with qPCR. This showed that siCLDN3-623 transfection significantly improved LMIV copies in LMB-L cells at 3, 5, 7, and 9 d.p.i. The data are presented as mean ± SD; ns, not significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Assessment of permeability by FITC-Dextran. (<b>a</b>) The fluorescence ratios of the upper wells to the lower wells was measured, revealing that the function of the tight junction was stable after 4 days of culture. (<b>b</b>) LMIV infection increased the permeability of LMB-L cells. The relative fluorescence values were determined at 24 h, 48 h, and 72 h post LMIV infection, respectively; ns, not significant, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>The analysis of correlation between miR-181a and CLDN3. (<b>a</b>) The KEGG enrichment analysis of mRNA sequencing after LMIV infection; The red box indicates Tight junction signaling pathway. (<b>b</b>) Schematic diagrams of microRNA target prediction show the miR-181a binding site in the 3′-UTR of CLDN3.</p>
Full article ">Figure 7
<p>The analysis of interaction between miR-181a and the 3′-UTR of CLDN3 by dual-luciferase reporter assays. HEK293 cells were co-transfected with miR181a mimic, miR-negative, and different plasmids (pmirGLO-CLDN3-WT-3′-UTR and pmirGLO-CLDN3-Mut-3′-UTR). Firefly luciferase/Renilla luciferase activity was measured after 24 h of transfection, and significantly reduced relative luciferase activity was observed in cells transfected with miR-181a mimic plus pmirGLO-CLDN3-WT-3′-UTR. The values are shown as means ± SD (<span class="html-italic">n</span> = 9). The significance levels were set as * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
18 pages, 2885 KiB  
Article
Effect of Degree of Substitution and Molecular Weight on Transfection Efficacy of Starch-Based siRNA Delivery System
by Amir Regev, Chen Benafsha, Riki Goldbart, Tamar Traitel, Moshe Elkabets and Joseph Kost
Polysaccharides 2024, 5(4), 580-597; https://doi.org/10.3390/polysaccharides5040037 - 7 Oct 2024
Viewed by 872
Abstract
RNA interference (RNAi) is a promising approach for gene therapy in cancers, but it requires carriers to protect and deliver therapeutic small interfering RNA (siRNA) molecules to cancerous cells. Starch-based carriers, such as quaternized starch (Q-Starch), have been shown to be biocompatible and [...] Read more.
RNA interference (RNAi) is a promising approach for gene therapy in cancers, but it requires carriers to protect and deliver therapeutic small interfering RNA (siRNA) molecules to cancerous cells. Starch-based carriers, such as quaternized starch (Q-Starch), have been shown to be biocompatible and are able to form nanocomplexes with siRNA, but significant electrostatic interactions between the carrier and siRNA prevent its release at the target site. In this study, we aim to characterize the effects of the degree of substitution (DS) and molecular weight (Mw) of Q-Starch on the gene silencing capabilities of the Q-Starch/siRNA transfection system. We show that reducing the DS reduces the electrostatic interactions between Q-Starch and siRNA, which now decomplex at more physiologically relevant conditions, but also affects additional parameters such as complex size while mostly maintaining cellular uptake capabilities. Notably, reducing the DS renders Q-Starch more susceptible to enzymatic degradation by α-amylase during the initial Q-Starch pretreatment. Enzymatic cleavage leads to a reduction in the Mw of Q-Starch, resulting in a 25% enhancement in its transfection capabilities. This study provides a better understanding of the effects of the DS and Mw on the polysaccharide-based siRNA delivery system and indicates that the polysaccharide Mw may be the key factor in determining the transfection efficacy of this system. Full article
(This article belongs to the Special Issue Latest Research on Polysaccharides: Structure and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Reaction scheme for starch quaternization by CHMAC. (<b>B</b>) Q-Starch<sub>(DS)</sub> nitrogen content as function of amount of quaternization reagent. (<b>C</b>) FTIR spectrum of native starch and different Q-Starches. (<b>D</b>) Magnification of relevant peak in FTIR spectrum of (<b>C</b>).</p>
Full article ">Figure 2
<p>(<b>A</b>) Representative mechanism for Q-Starch<sub>(DS)</sub>/siRNA complex formation through self-assembly. (<b>B</b>) Q-Starch<sub>(0.12)</sub>/siRNA complex formation evaluated by agarose gel electrophoresis at increasing N/P molar ratios. (<b>C</b>) Zeta potential of Q-Starch<sub>(DS)</sub>/siRNA complexes at increasing DS at N/P molar ratios of 1, 2, and 3.</p>
Full article ">Figure 3
<p>(<b>A</b>) An illustration depicting the ionic strength agarose gel electrophoresis experimental procedure; free siRNA is observed after complex exposure to increasing ionic strength. (<b>B</b>) The agarose gel electrophoresis of Q-Starch<sub>(0.12)</sub>/siRNA complexes formed at an N/P molar ratio of 2 at increasing ionic strengths. (<b>C</b>) NaCl concentration in which decomplexation (free siRNA) was first observed in each Q-Starch<sub>(DS)</sub> at an N/P ratio of 2. (<b>D</b>) Particle size distribution for Q-Starch<sub>(DS)</sub>/siRNA complexes at an N/P molar ratio of 2. (<b>E</b>) The mean particle size for the different Q-Starch<sub>(DS)</sub>/siRNA complexes at an N/P molar ratio of 2. (<b>F</b>) Two representative cryoTEM images of Q-Starch<sub>(0.12)</sub>/siRNA complexes (marked with a red arrows) at an N/P molar ratio of 2.</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>) An illustration depicting the ionic strength agarose gel electrophoresis experimental procedure; free siRNA is observed after complex exposure to increasing ionic strength. (<b>B</b>) The agarose gel electrophoresis of Q-Starch<sub>(0.12)</sub>/siRNA complexes formed at an N/P molar ratio of 2 at increasing ionic strengths. (<b>C</b>) NaCl concentration in which decomplexation (free siRNA) was first observed in each Q-Starch<sub>(DS)</sub> at an N/P ratio of 2. (<b>D</b>) Particle size distribution for Q-Starch<sub>(DS)</sub>/siRNA complexes at an N/P molar ratio of 2. (<b>E</b>) The mean particle size for the different Q-Starch<sub>(DS)</sub>/siRNA complexes at an N/P molar ratio of 2. (<b>F</b>) Two representative cryoTEM images of Q-Starch<sub>(0.12)</sub>/siRNA complexes (marked with a red arrows) at an N/P molar ratio of 2.</p>
Full article ">Figure 4
<p>Cellular uptake and gene silencing of Q-Starch<sub>(DS)</sub>/siRNA complexes. (<b>A</b>) Mean cy5 intensity following 24 h of incubation with Q-Starch<sub>(DS)</sub> complexes as assessed via FACSAria III; (<b>B</b>) EGFR gene mRNA expression following 72 h of incubation with Q-Starch<sub>(DS)</sub> complexes as assessed via RT-PCR (ns: no statistical significance).</p>
Full article ">Figure 5
<p>(<b>A</b>) Illustration depicting expected experimental results for iodine test following Q-Starch<sub>(DS)</sub> cleavage, together with experimental results for Q-Starch<sub>(0.12)</sub>. (<b>B</b>) Required α-amylase concentration for Q-Starch<sub>(DS)</sub> cleavage at different DSs via iodine starch test. (n = 3) (<b>C</b>) Shear viscosity as function of applied shear rate for Q-Starch<sub>(0.59)</sub> and Q-Starch<sub>(0.30)</sub> before and after cleavage with α-amylase enzyme. (<b>D</b>) Agarose gel electrophoresis results of Q-Starch <sub>(0.30)</sub>/siRNA and Q-Starch <sub>(0.59)</sub>/siRNA complexes with or without pre-cleavage of Q-Starch with α-amylase enzyme at 1:1 wt. %. All complexes were formed at N/P molar ratio of 2. (<b>E</b>) RT-PCR results of EGFR gene mRNA expression in Cal33 cells, 72 h post-transfection with Q-Starch<sub>(0.59)</sub>/siRNA complexes; four examined groups include cells treated with Q-Starch<sub>(DS)</sub>/siRNA<sup>NC5</sup> and Q-Starch<sub>(DS)</sub>/siRNA<sup>EGFR</sup> complexes at siRNA concentration of 50 nM, N/P molar ratio of 2, with and without pre-cleavage of Q-Starch<sub>(0.59)</sub> with α-amylase 24 h before transfection. Average + SEM (n = 3). ns: not statistically significant (<span class="html-italic">p</span> &gt; 0.05), *: statistically significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
20 pages, 4922 KiB  
Article
The 419th Aspartic Acid of Neural Membrane Protein Enolase 2 Is a Key Residue Involved in the Axonal Growth of Motor Neurons Mediated by Interaction between Enolase 2 Receptor and Extracellular Pgk1 Ligand
by Bing-Chang Lee, Jui-Che Tsai, Yi-Hsin Huang, Chun-Cheng Wang, Hung-Chieh Lee and Huai-Jen Tsai
Int. J. Mol. Sci. 2024, 25(19), 10753; https://doi.org/10.3390/ijms251910753 - 6 Oct 2024
Viewed by 1112
Abstract
Neuron-specific Enolase 2 (Eno2) is an isozyme primarily distributed in the central and peripheral nervous systems and neuroendocrine cells. It promotes neuronal survival, differentiation, and axonal regeneration. Recent studies have shown that Eno2 localized on the cell membrane of motor neurons acts as [...] Read more.
Neuron-specific Enolase 2 (Eno2) is an isozyme primarily distributed in the central and peripheral nervous systems and neuroendocrine cells. It promotes neuronal survival, differentiation, and axonal regeneration. Recent studies have shown that Eno2 localized on the cell membrane of motor neurons acts as a receptor for extracellular phosphoglycerate kinase 1 (ePgk1), which is secreted by muscle cells and promotes the neurite outgrowth of motor neurons (NOMN). However, interaction between Eno1, another isozyme of Enolase, and ePgk1 failed to return the same result. To account for the difference, we constructed seven point-mutations of Eno2, corresponding to those of Eno1, and verified their effects on NOMN. Among the seven Eno2 mutants, eno2-siRNA-knockdown NSC34 cells transfected with plasmid encoding the 419th aspartic acid mutated into serine (Eno2-[D419S]) or Eno2-[E420K] showed a significant reduction in neurite length. Moreover, the Eno2-ePgk1-interacted synergic effect on NOMN driven by Eno2-[D419S] was more profoundly reduced than that driven by Eno2-[E420K], suggesting that D419 was the more essential residue involved in NOMN mediated by Eno2-ePgk1 interaction. Eno2-ePgk1-mediated NOMN appeared to increase the level of p-Cofilin, a growth cone collapse marker, in NSC34 cells transfected with Eno2-[D419S] and incubated with ePgk1, thereby inhibiting NOMN. Furthermore, we conducted in vivo experiments using zebrafish transgenic line Tg(mnx1:GFP), in which GFP is tagged in motor neurons. In the presence of ePgk1, the retarded growth of axons in embryos injected with eno2-specific antisense morpholino oligonucleotides (MO) could be rescued by wobble-eno2-mRNA. However, despite the addition of ePgk1, the decreased defective axons and the increased branched neurons were not significantly improved in the eno2-[D419S]-mRNA-injected embryos. Collectively, these results lead us to suggest that the 419th aspartic acid of mouse Eno2 is likely a crucial site affecting motor neuron development mediated by Eno2-ePgk1 interaction, and, hence, mutations result in a significant reduction in the degree of NOMN in vitro and axonal growth in vivo. Full article
(This article belongs to the Special Issue Molecular Research on Neuronal Cell Death and Neurogenesis)
Show Figures

Figure 1

Figure 1
<p>Construction of Eno2 mutant protein expression plasmid. A two-step PCR method was used to generate the expression plasmid, with the mutation of aspartic acid (D) to serine (S) at position 419 (Eno2-[D419S]) as an example. Four primers (p1 to p4) were used for the PCR, where p1 and p4 represent the forward and reverse flanking primers of the <span class="html-italic">eno2</span> fragment, and p2 and p3 are complementary mutagenic primers carrying the mutation sequence. The plasmid pCS2-Eno2-wb- Flag containing the mouse <span class="html-italic">eno2</span> gene fused with a Flag reporter gene and with wobble-modified nucleotides (wb) was used as the template in the first PCR step. Primers p1 and p2 were used to produce the upstream fragment of the gene containing the mutation site (X), while p3 and p4 were used to produce the downstream fragment. In the second PCR step, the products from the first step were mixed, and a small amount of p1 and p4 primers were added to generate the complete mutant gene fragment. After PCR completion, the mutant gene fragment and the pCS2 plasmid were digested with EcoRI and XhoI restriction enzymes to produce sticky-ended inserts and vectors. Finally, the insert and vector were ligated to create the pCS2-derived expression plasmid for the Eno2 mutant protein fused with the Flag reporter gene. Blue color: <span class="html-italic">eno2</span> cDNA; orange color: reporter Flag; red color: mutation site; grey color: plasmid pCS2<sup>+</sup> backbone.</p>
Full article ">Figure 2
<p>Western blot analysis demonstrated that the recombinant Eno2-wb protein was expressed in the Eno2-knockdown cells transfected with siRNA. <span class="html-italic">NSC34</span> cells were transfected with <span class="html-italic">eno2</span> siRNA to knock down the endogenous Eno2, followed by transfection DNA encoding Eno2 but consisting of nucleotides modified at the wobble position (Eno2-wb) to avoid the block by introducing siRNA. The expression level of Eno2 protein in the cells was detected by the Eno2 antibody. Lane 1 was the control group, and its Eno2 expressional level was set at 1, while lanes 2–4 represented Eno2 expressed within NSC cells treated without (−) or with (+) <span class="html-italic">eno2</span> siRNA or <span class="html-italic">eno2-wb</span>. The expressional level of Eno2 was quantified relative to that of α-Tubulin that served as an internal control. The number listed below each lane indicates the fold change of Eno2 intensity of each treatment compared to the control group, set as 1.</p>
Full article ">Figure 3
<p>The effect of various point mutations in Eno2 on promoting the neurite growth of motor neurons. (<b>A</b>,<b>B</b>) Morphological differentiation of cultured NSC34 neural cells: (<b>A</b>) one-day incubation; (<b>B</b>) five-day incubation. Axonal neurites derived from cultured NSC34 cells were observed, and the neurite length was measured. (<b>C</b>) Statistical analysis of the average length of neurites. NSC34 cells transfected with siRNA (pCS2<sup>+</sup>) to inhibit endogenous mouse <span class="html-italic">eno2</span> served as a negative control. After transfection with <span class="html-italic">eno2</span> siRNA, <span class="html-italic">eno2</span> DNA with modified wobble-nucleotides (<span class="html-italic">eno2-wb</span>) was transfected in NSC34. The resultant neurite length was normalized to a value of 1.0, which served as the positive control. The average neurite length of motor neurons was measured after transfection with siRNA, followed by transfection with plasmid harboring different point-mutated <span class="html-italic">eno2-wb</span> DNA. The fold change in neurite length was calculated relative to the positive control set as 1.0. Then, each experimental group was independently compared with the control group using <span class="html-italic">t</span>-test for statistical analysis (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; ns indicates no significant difference). The <span class="html-italic">t</span> values and degrees of freedom for groups of <span class="html-italic">-[L410I]</span>, <span class="html-italic">-[M411L]</span>, <span class="html-italic">-[D419S]</span>, <span class="html-italic">-[E420K], -[R422K], -[H426R]</span> and <span class="html-italic">-[N427S]</span> were 0.5754 and 2, 2.309 and 2, 6.362 and 3, 2.064 and 3, 5.413 and 2, 4.255 and 2, and 4.496 and 3, respectively.</p>
Full article ">Figure 4
<p>The synergistic effect of promoting motor neurons by the addition of Pgk1 and transfection of various mutations of Eno2. Statistical analysis of the average length of neurites. NSC34 neural cells transfected with siRNA (pCS2+) to inhibit endogenous mouse Eno2 without the addition of extracellular Pgk1 (ePgk1) served as a negative control. As the positive control, NSC34 cells were transfected with <span class="html-italic">eno2</span> siRNA, followed by transfection of wobble-modified <span class="html-italic">eno2</span> DNA (<span class="html-italic">eno2-wb</span>) combined with Pgk1 immersion. The resultant average neurite length was normalized as 1.0. Similarly, plasmid harboring a point-mutated <span class="html-italic">eno2</span> DNA, as indicated, was transfected in siRNA-treated cells, followed by Pgk1 immersion. The average neurite length of motor neurons was measured, and fold change compared to the positive control was calculated. Each experimental group was independently compared with the control group using Student’s <span class="html-italic">t</span>-test for statistical significance (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; ns indicates no significant difference). The t values and degrees of freedom for groups of pCS2<sup>+</sup>, <span class="html-italic">-[M411L]</span>, <span class="html-italic">-[D419S]</span> and <span class="html-italic">–[E420K]</span> were 4.972 and 5, 1.151 and 2, 9.281 and 2, and 6.189 and 2, respectively.</p>
Full article ">Figure 5
<p>The rescue effect of mutant <span class="html-italic">eno2</span> mRNA on Eno2 knockdown in zebrafish embryos. Microinjection was performed at the one-cell stage in the transgenic line <span class="html-italic">Tg(mnx1:GFP)</span>. Embryos were collected at 30 hpf and observed for motor neurons in the 6th to 17th somites, as shown in Figure (<b>A</b>,<b>A’</b>) with (<b>A</b>) under visible light and (<b>A’</b>) under fluorescence. (<b>B</b>) Western blot analysis comparing Eno2 protein levels (indicated by arrows) in embryos with and without MO injection, using rabbit anti−Eno2 antibodies for detection. (<b>C</b>–<b>E</b>) Three phenotypes: normal, mild defect, and severe defect with arrows indicating defect locations. Scale bar: 100 μm. (<b>F</b>,<b>G</b>) Frequency of defects. Control groups were designated as (1) the control group without treatment; (2) the eno2 MO group in which injection with eno2-specific antisense morpholino oligonucleotide (MO) served as a negative control; and (3) the MO plus <span class="html-italic">eno2</span> mRNA group in which co-injection of eno2 MO and <span class="html-italic">eno2</span>-wb mRNA served as a positive control. Experimental groups were (<b>F</b>) the MO plus <span class="html-italic">eno2-wb-[D419S]</span> mRNA group: co-injection of eno2 MO combined with <span class="html-italic">eno2-wb-[D419S]</span> mRNA; and (<b>G</b>) the MO plus <span class="html-italic">eno2[E420K]</span> mRNA group: co-injection of eno2 MO combined with <span class="html-italic">eno2-wb-[E420K]</span> mRNA. Both figures represent the averages of three independent experiments. White indicates the percentage of severe defects, gray indicates the percentage of mild defects, and black indicates the overall defect percentage. Statistical significance was analyzed using one-way ANOVA (* indicates <span class="html-italic">p</span> &lt; 0.05; ns indicates no significant difference). F statistics of (<b>F</b>) were 1.401 as numerator and 2.802 as denominator, while (<b>G</b>) were 1.917 as numerator and 3.833 as denominator.</p>
Full article ">Figure 6
<p>The effect of overexpressing mutant Eno2 on the occurrence rate of branched axons of motor neurons in zebrafish embryos immersed with recombinant Pgk1. (<b>A</b>,<b>B</b>) Two phenotypes of the caudal primary (CaP) axons of motor neurons in the 30 hpf transgenic line <span class="html-italic">Tg(mnx1:GFP)</span> embryos. (<b>A</b>) Examples of normal phenotype and (<b>B</b>) branched axon phenotype (branching sites indicated by white arrowheads). Scale bar: 50 μm. Experimental manipulations were performed at the one-cell stage. The first group of zebrafish embryos served as the non-injected control group, the second group was injected with mouse <span class="html-italic">eno2-wb</span> mRNA, and the third group was injected with mutated <span class="html-italic">eno2-wb</span> mRNA. At 30 hpf, the percentage of embryos exhibiting the branched axonal phenotype among CaP neurons was calculated for each group. (<b>C</b>) The results of embryos injected with <span class="html-italic">eno2-wb-[D419]</span> mRNA and (<b>D</b>) the results of embryos injected with <span class="html-italic">eno2-wb-[E420K]</span> mRNA. Each panel represents the average of three independent experiments with statistical analysis performed using one-way ANOVA (*, <span class="html-italic">p</span> &lt; 0.05; ns indicates no significant difference). F statistics of (<b>C</b>) were 1.239 as numerator and 2.478 as denominator, while (<b>D</b>) were 1 as numerator and 2.001 as denominator.</p>
Full article ">Figure 7
<p>Using Western blot to analyze the level of phosphorylated Cofilin (p-Cofilin) expressed in NSC34 cells. (<b>A</b>) Western blot analysis. NSC34 cells were transfected with <span class="html-italic">eno2</span>-siRNA to knock down endogenous Eno2, then rescued with different DNA material, as indicated, in the absence (−) or presence (+) of Pgk1 protein. pCS<sup>+</sup>: plasmid with an siRNA insertion; <span class="html-italic">eno2-wb</span>: wild type Eno2 but consisting of nucleotides modified at wobble position (<span class="html-italic">eno2-wb</span>) to avoid being blocked by introducing siRNA; and <span class="html-italic">D419S</span>: mutant Eno2-wb consisting of an aspartic acid mutated into serine at the 419th position of Eno2. The expressional level of p-Cofilin at Ser3 was quantified relative to that of α-Tubulin, which served as an internal control. The number listed below each lane indicates the fold change of the p-Cofilin intensity of each treatment compared to the control group, set as 1. (<b>B</b>) Quantitative and statistical analyses. Data were averaged from three independent experiments and presented as mean ± SD (<span class="html-italic">n</span> = 3). One-way ANOVA, followed by Tukey’s multiple comparison test, was used to perform statistical analysis (** <span class="html-italic">p</span> &lt; 0.005). F statistics were 4 as numerator and 10 as denominator.</p>
Full article ">Figure 8
<p>Molecular docking model to illustrate the key amino acid involved in Eno2-ePgk1 interaction. (<b>A</b>,<b>B</b>) Simulated model to illustrate how Eno2 (PDB ID: 5TD9) interacted with ePgk1 (PDB ID: 2ZGV) is presented in blue and orange, respectively. Drawing of partial enlargement of critical amino acid residues involved in Eno2-ePgk1 interaction was presented, including (<b>A</b>) wild-type Eno2 D419 and (<b>B</b>) mutant Eno2 D419S. The segment of the 345th to 360th amino acids of Pgk1, the segment of the 404th to 431st amino acid of Eno2, and the mutant Eno2 D419th were shown in Dodger blue, chocolate, and orange, respectively. (<b>C</b>,<b>D</b>) Drawing of partial enlargement of the surface polarity distribution of (<b>C</b>) Eno2 D419 and (<b>D</b>) Eno2 D419S. Dotted circles highlighted the D419th residue on Eno2. Lipophilicity was presented as low (hydrophilicity, in cyan) to high (hydrophobicity, in brown). Dotted zone indicates the 419th site. (<b>E</b>,<b>F</b>) Drawing of partial enlargement of the surface charge distribution of (<b>E</b>) Eno2 D419 and (<b>F</b>) Eno2 D419S. Chargeability from negative charge (in red) to positive charge (in blue). Dotted circles marked the 419th residue of Eno2.</p>
Full article ">
22 pages, 5283 KiB  
Article
Understanding the Dosage-Dependent Role of Dicer1 in Thyroid Tumorigenesis
by María Rojo-Pardillo, Ludivine Godefroid, Geneviève Dom, Anne Lefort, Frederick Libert, Bernard Robaye and Carine Maenhaut
Int. J. Mol. Sci. 2024, 25(19), 10701; https://doi.org/10.3390/ijms251910701 - 4 Oct 2024
Viewed by 932
Abstract
Tumors originating from thyroid follicular cells are the most common endocrine tumors, with rising incidence. Despite a generally good prognosis, up to 20% of patients experience recurrence and persistence, highlighting the need to identify the underlying molecular mechanisms. Dicer1 has been found to [...] Read more.
Tumors originating from thyroid follicular cells are the most common endocrine tumors, with rising incidence. Despite a generally good prognosis, up to 20% of patients experience recurrence and persistence, highlighting the need to identify the underlying molecular mechanisms. Dicer1 has been found to be altered in papillary thyroid cancer (PTC). Studies suggest that Dicer1 functions as a haploinsufficient tumor suppressor gene: partial loss promotes tumorigenesis, while complete loss prevents it. To investigate the effects of partial or total Dicer1 loss in PTC in vitro, we generated stable Dicer1 (+/−) cell lines from TPC1 using CRISPR-Cas9, though no Dicer1 (−/−) lines could be produced. Therefore, siRNA against Dicer1 was transfected into Dicer1 (+/−) cell lines to further decrease its expression. Transcriptomic analysis revealed changes in proliferation and cell locomotion. BrdU staining indicated a slow-down of the cell cycle, with fewer cells in S phase and more in G0-G1-phase. Additionally, transwell assays showed decreased invasion and migration after Dicer1 knockdown by siRNA. Moreover, Dicer1 overexpression led to decreased proliferation, invasion, and increased apoptosis. Our findings deepen the understanding of Dicer1’s role in thyroid cancer, demonstrating that both complete elimination and overexpression of Dicer1 inhibit thyroid oncogenesis, highlighting Dicer1 as a promising target for novel therapeutic strategies. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Dicer1</span> mRNA expression was decreased in the human PTC and in the PTC-derived cell line TPC1. (<b>A</b>) Analysis of <span class="html-italic">Dicer1</span> mRNA expression in The Cancer Genome Atlas (TCGA) in normal samples (red, Solid Tissue Normal, n = 59), PTC (blue, primary tumors, n = 505), and metastases (purple, n = 8). The expression (log2 (RPM+1)) refers to the logarithm base 2 of the Reads Per Million (RPM) value increased by one. Statistically significant differences were determined using the Kruskal–Wallis test *** <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Analysis of <span class="html-italic">Dicer1</span> mRNA expression using RT-qPCR in 9 PTC (blue) and their adjacent normal tissue (red, NT). Statistically significant differences were determined using the Mann–Whitney test ** <span class="html-italic">p</span> &lt; 0.01. (<b>C</b>) Analysis of <span class="html-italic">Dicer1</span> mRNA expression in the PTC-derived cell line TPC1 (blue, n = 7) compared with a pool of normal human thyroid tissues (red, NT, n = 8). The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney test *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Generation of D11, WB1, and TD5 <span class="html-italic">Dicer1</span> (+/−) cell lines using CRISPR-Cas9. (<b>A</b>) DNA was extracted from isolated clones after transfection, amplified, cloned into a plasmid, and sequenced using Sanger sequencing. TPC1 presented two wild-type alleles (+); D11, WB1, and TD5 displayed a combination of one wild-type allele (+) and one knockout allele (−), with deletions of 20, 7, and 185 nucleotides, respectively. (<b>B</b>) Western blot analysis of Dicer1 protein expression in TPC1 and in CRISPR-Cas9 generated D11, WB1, and TD5 cell lines. The upper image corresponds to the detection of Dicer1 with a molecular weight of 217 kDa, while the lower panel corresponds to the detection of actin with a molecular weight of 42 kDa. (<b>C</b>) Western blot quantification of Dicer1 protein expression in TPC1 and in CRISPR-Cas9 generated D11, WB1, and TD5 cell lines (n = 8). Data were normalized to Dicer1 expression in TPC1. The columns represent the mean values, and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Kruskal–Wallis test. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>RNAseq analysis did not reveal major transcriptomic changes in heterozygous cell lines. (<b>A</b>) PCA analysis of gene variance between TPC1 (purple), D11 (green), WB1 (blue), and TD5 (red). (<b>B</b>,<b>C</b>) Analysis of common upregulated or downregulated genes between TPC1 and D11 (red), TPC1 and WB1 (green), and TPC1 and TD5 (blue). (<b>D</b>) PNMA2, ICAM1, SMIM10, CASP10, EPHA5, LINC01234, ZNF28, PPP1R3C, and ZNF468 mRNA expression analysis using RT-qPCR in TPC1, D11, WB1, and TD5 cell lines (n = 5). The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Kruskal–Wallis test. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Heterozygous loss of Dicer1 did not impact proliferation, apoptosis, invasion or migration. (<b>A</b>) Cell proliferation was analyzed in TPC1, D11, WB1, and TD5 cell lines (n = 4). Cells were intracellularly stained with BrdU Staining Kit for Flow Cytometry. The percentage of BrdU positive cells (cells in S phase of the cell cycle) was determined via flow cytometry. (<b>B</b>) Apoptosis was analyzed using Caspase-Glo<sup>®</sup> 3/7 Assay in TPC1, D11, WB1, and TD5 cell lines (n = 6). Each point represents the average of 2 wells. The blank control value has been subtracted from each point. (<b>C</b>) Migration analysis in TPC1 (n = 10), D11 (n = 6), WB1 (n = 10), and TD5 (n = 10) cell lines. (<b>D</b>) Invasion analysis in TPC1, D11, WB1, and TD5 cell lines (n = 5). The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Data were normalized to TPC1. Statistically significant differences were determined using the Kruskal–Wallis test.</p>
Full article ">Figure 5
<p>Dicer1 mRNA and protein expression were strongly reduced four days after Dicer1 siRNA transfection. (<b>A</b>) Dicer1 mRNA expression was analyzed using RT-qPCR in the TPC1 (n = 7), D11 (n = 5), WB1 (n = 4), and TD5 (n = 4) cell lines transfected with the negative control (NC) or with Dicer1 siRNA (siDICER). (<b>B</b>) Western blot analysis of Dicer1 protein expression in the same conditions. The upper image corresponds to the detection of Dicer1 with a molecular weight of 217 kDa, while the lower panel corresponds to the detection of vinculin with a molecular weight of 117 kDa. (<b>C</b>) Quantification of Dicer1 protein levels in TPC1 (n = 8), D11 (n = 7), WB1 (n = 6) and TD5 (n = 6) cells. Data were normalized to NC. The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney test. * <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.</p>
Full article ">Figure 6
<p>RNAseq analysis revealed transcriptomic alterations after transfection with Dicer1 siRNA in D11 cells. (<b>A</b>) Heatmap representation of the hierarchical clustering (Pearson &lt; 0.0001) of genes selected with DESeq2 method, FDR 0.05, and fold change 1.5× (following Dicer1 siRNA (siDICER)) transfection vs. control (following transfection with negative control (NC)) (1123 upregulated genes, in red, and 1204 downregulated genes, in green). (<b>B</b>,<b>C</b>) Top 10 upregulated or downregulated pathways network obtained through the enrichment analysis of gene expression (FDR 0.05 and minimum fold change 1.5×) in D11 cells transfected with siDICER compared with controls using The Gene Ontology Biological Process database.</p>
Full article ">Figure 7
<p>Total Dicer1 loss inhibited proliferation. (<b>A</b>) Cell counts were conducted in TPC1 (n = 7), D11 (n = 4), WB1 (n = 5), and TD5 (n = 5) cell lines transfected with the negative control (NC) or with Dicer1 siRNA (siDICER). (<b>B</b>) TPC1 cells transfected with the negative control (NC) or with Dicer1 siRNA (siDICER) were intracellularly stained with Anti-BrdU-APC-A and 7-AAD using the BrdU Staining Kit for Flow Cytometry. For each condition (NC or siDICER), the left panel shows only BrdU positive cells (S phase, P4). On the right panel, BrdU positive cells are stained in blue (S phase, P5) and BrdU negative cells are stained in green (low 7-AAD concentrations, G0/G1 phases, P7) or purple (high 7-AAD concentrations, G2/M phases, P6). (<b>C</b>) Quantification of cells in P5 (S phase) in TPC1 (n = 7), D11 (n = 4), WB1 (n = 5), and TD5 (n = 5) cell lines following negative control (NC) or Dicer1 siRNA (siDICER) transfection. (<b>D</b>) Quantification of cells in P7 (G0/G1 phases) in the same experimental conditions. All experiments were conducted four days after transfection. Data were normalized to the NC. The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney <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 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Cyclin E but not cyclin D1 was downregulated following Dicer1 siRNA transfection. (<b>A</b>) Western blot analysis of cyclin E and cyclin D1 protein expression in TPC1 cells transfected with the negative control (NC) or with Dicer1 siRNA (siDICER). The upper image corresponds to the detection of cyclin E or D1 (48 kDa and 34 kDa, respectively) while the lower panel corresponds to the detection of vinculin (117 kDa). (<b>B</b>) Western blot quantification of cyclin E expression in TPC1 (n = 6), D11 (n = 4), WB1 (n = 6), and TD5 (n = 7) cell lines following negative control (NC) or Dicer1 siRNA (siDICER) transfection. (<b>C</b>) Western blot quantification of cyclin D1 expression in the same experimental conditions (n = 6). Cyclin expression was normalized to vinculin expression and data were normalized to the NC. The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney <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 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>Total loss of Dicer1 resulted in an increase in apoptosis. (<b>A</b>) Apoptosis was analyzed using Caspase-Glo<sup>®</sup> 3/7 Assay in TPC1 (n = 10), D11 (n = 7), WB1 (n = 12), and TD5 (n = 12) cells transfected with the negative control (NC) or with Dicer1 siRNA (siDICER). (<b>B</b>) The percentage of apoptotic cells was determined using a Dead cell apoptosis kit with Annexin V FITC and Propidium Iodide for flow cytometry in TPC1 (n = 5), D11 (n = 6), WB1 (n = 6), and TD5 (n = 6) cells transfected with the negative control (NC) or with Dicer1 siRNA (siDICER). Data were normalized to NC. The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using THE Mann–Whitney <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 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>Dicer1 total loss inhibited migration and invasion. The number of cells migrating (<b>A</b>) or invading (<b>B</b>) was counted 4 days after transfection with the negative control (NC) or Dicer1 siRNA (siDICER). The red arrows point to examples of the cells. (<b>C</b>) Quantification of migrating cells in TPC1 (n = 4), D11 (n = 5), WB1 (n = 5), and TD5 (n = 4) cell lines. (<b>D</b>) Quantification of invasive cells in TPC1 (n = 4), D11 (n = 5), WB1 (n = 5), and TD5 (n = 4) cell lines. Data were normalized to NC. The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney <span class="html-italic">t</span>-test. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 11
<p>Dicer1 knockdown inhibited proliferation as well as migration and invasion in BCPAP and H-Tori3 cells. (<b>A</b>) Analysis of Dicer1 mRNA expression in BCPAP (n = 4) and H-Tori3 (n = 7) cells compared with a pool of normal human thyroid tissues (NT, n = 8). (<b>B</b>) Dicer1 protein expression levels in BCPAP (n = 6) and H-Tori3 (n = 4) cells following negative control (NC) or Dicer1 siRNA transfection (siDICER). (<b>C</b>) Quantification of cells in S phase in H-Tori3 (n = 8) and BCPAP (n = 5) cells. (<b>D</b>) Quantification of cells in G0/G1 phase in H-Tori3 (n = 8) and BCPAP (n = 5). (<b>E</b>) Quantification of migrating cells in H-Tori3 (n = 6) and BCPAP (n = 6). (<b>F</b>) Quantification of invasive cells in H-Tori3 (n = 4) and BCPAP (n = 4). All experiments were conducted four days after transfection with the negative control (NC) or with Dicer1 siRNA (siDICER). Data were normalized to the NC. The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney <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 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 12
<p>Increasing Dicer1 expression inhibited proliferation and invasion while raising apoptosis. All experiments were performed in TPC1 cells transfected with Dicer1 plasmid (Dicer1) 48 h after transfection. (<b>A</b>) Quantification of cells in S phase (n = 6). (<b>B</b>) Apoptosis was analyzed using Caspase-Glo<sup>®</sup> 3/7 Assay (n = 9). (<b>C</b>) The percentage of apoptotic cells was determined using a Dead cell apoptosis kit with Annexin V FITC and Propidium Iodide for flow cytometry (n = 9). (<b>D</b>) Quantification of invasive cells (n = 6). Data were normalized to NC (negative control: cells incubated in the presence of lipofectamine 3000). The columns represent the mean values and the error bars indicate the standard deviation (mean ± SD). Statistically significant differences were determined using the Mann–Whitney <span class="html-italic">t</span>-test. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 13
<p><span class="html-italic">Dicer1</span> exerted dosage-dependent effects in thyroid cells. (<b>A</b>) Very low levels or total absence of <span class="html-italic">Dicer1</span> resulted in reduced (↓) proliferation, invasion and migration, along with increased (↑) apoptosis, suggesting an inhibition of tumorigenesis. Similarly, high levels of <span class="html-italic">Dicer1</span> led to comparable proliferation, invasion, and apoptosis outcomes. Intermediate levels of <span class="html-italic">Dicer1</span> (as in TPC1, D11, WB1, and TD5 cells) led to increased proliferation and reduced apoptosis, potentially promoting tumorigenesis. (<b>B</b>) High levels of <span class="html-italic">Dicer1</span> were obtained via plasmid transfection. TPC1 had intermediate levels of <span class="html-italic">Dicer1</span> and heterozygous <span class="html-italic">Dicer1</span> (+/−) D11, WB1, and TD5 cells showed a 50% decrease in <span class="html-italic">Dicer1</span> expression compared with TPC1 cells. Following <span class="html-italic">Dicer1</span> siRNA transfection, <span class="html-italic">Dicer1</span> was almost completely lost.</p>
Full article ">
Back to TopTop