[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,495)

Search Parameters:
Keywords = Rapamycin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 1822 KiB  
Review
Biochemical Pathways Delivering Distinct Glycosphingolipid Patterns in MDA-MB-231 and MCF-7 Breast Cancer Cells
by Anita Markotić, Jasminka Omerović, Sandra Marijan, Nikolina Režić-Mužinić and Vedrana Čikeš Čulić
Curr. Issues Mol. Biol. 2024, 46(9), 10200-10217; https://doi.org/10.3390/cimb46090608 (registering DOI) - 15 Sep 2024
Abstract
The complex structure of glycosphingolipids (GSLs) supports their important role in cell function as modulators of growth factor receptors and glutamine transporters in plasma membranes. The aberrant composition of clustered GSLs within signaling platforms, so-called lipid rafts, inevitably leads to tumorigenesis due to [...] Read more.
The complex structure of glycosphingolipids (GSLs) supports their important role in cell function as modulators of growth factor receptors and glutamine transporters in plasma membranes. The aberrant composition of clustered GSLs within signaling platforms, so-called lipid rafts, inevitably leads to tumorigenesis due to disturbed growth factor signal transduction and excessive uptake of glutamine and other molecules needed for increased energy and structural molecule cell supply. GSLs are also involved in plasma membrane processes such as cell adhesion, and their transition converts cells from epithelial to mesenchymal with features required for cell migration and metastasis. Glutamine activates the mechanistic target of rapamycin complex 1 (mTORC1), resulting in nucleotide synthesis and proliferation. In addition, glutamine contributes to the cancer stem cell GD2 ganglioside-positive phenotype in the triple-negative breast cancer cell line MDA-MB-231. Thieno[2,3-b]pyridine derivative possesses higher cytotoxicity against MDA-MB-231 than against MCF-7 cells and induces a shift to aerobic metabolism and a decrease in S(6)nLc4Cer GSL-positive cancer stem cells in the MDA-MB-231 cell line. In this review, we discuss findings in MDA-MB-231, MCF-7, and other breast cancer cell lines concerning their differences in growth factor receptors and recent knowledge of the main biochemical pathways delivering distinct glycosphingolipid patterns during tumorigenesis and therapy. Full article
(This article belongs to the Special Issue Latest Review Papers in Molecular Biology 2024)
Show Figures

Figure 1

Figure 1
<p>Ceramide synthesis.</p>
Full article ">Figure 2
<p>Structure of glycosphingolipid Gg3Cer. Acetamide group of N-acetylglucosamine is marked in red. <span class="html-italic">Trans</span> double bond within sphingosine, responsible for lipid raft formation, is marked in green.</p>
Full article ">Figure 3
<p>The interplay of growth factor and estrogen signaling in breast cancer cell proliferation, survival, and migration. Higher activation of PLCγ and mTOR is expected in MCF-7 cells containing ER, HER3, and a low level of HER2, which are absent in MDA-MB-231 cells. Compound <b>1</b>, as an inhibitor of PLCγ, is effective in lowering the percentage of MDA-MB-231 but not MCF-7 CSCs. Abbreviations: AKT, protein kinase B or Akt; Compound <b>1</b> or thieno[2,3-<span class="html-italic">b</span>]pyridine derivative, 3-amino-<span class="html-italic">N</span>-(3-chloro-2-methylphenyl)-5-oxo-5,6,7,8-tetrahydrothieno[2,3-<span class="html-italic">b</span>]quinoline-2-carboxamide; E, estrogen; ER, estrogen receptor; mTORC1 and mTORC2, mechanistic targets of rapamycin complex I and II; PDK1, 3-phosphoinositide-dependent kinase 1; PI3K, phosphoinositide 3-kinase; PIP<sub>3</sub>, phosphatidylinositol 3,4,5-trisphosphate; and PLC gamma, phospholipase C gamma. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 4
<p>Synthesis of ganglioseries GSLs catalyzed by enzymes, written with blue letters, induced in breast CSCs [<a href="#B72-cimb-46-00608" class="html-bibr">72</a>,<a href="#B75-cimb-46-00608" class="html-bibr">75</a>] and of neolactoseries GSLs. Percentages of CSCs positive for red and blue framed GSLs were decreased and increased, respectively, after thieno[2,3-<span class="html-italic">b</span>]pyridine derivative (Compound <b>1</b>, yellow letters) treatment in MDA-MB-231 (red arrow) and MCF-7 (blue arrow) cells, respectively [<a href="#B2-cimb-46-00608" class="html-bibr">2</a>]. Abbreviations: In the ganglioside nomenclature, G = ganglioside, with the corresponding number of the sialic acid residues described with letters (M = mono, D = di), and the numbers denote the number of neutral sugar residues that are required to reach the number 5 (1 = GalGalNAcGalGlc, 2 = GalNAcGalGlc, 3 = GalGlc). Glycosidic residues: Gal = galactose, Glc = glucose, GlcNAc = N-acetylglucosamine, GalNAc = N-acetylgalactosamine. Neutral GSLs: Gg3Cer = gangliotriaosylceramide, Gg4Cer = gangliotetraosylceramide, Lc3Cer = lactotriaosylceramide, nLc4Cer = neolactotetraosylceramide. Acidic GSL: S(6)nLc4Cer = sialyl residue bound by α2-3 glycosidic bond to nLc4Cer. UGCG = UDP-glucose ceramide glycosyltransferase. In the nomenclature of other glycosyltransferases, the letter B = β-glycosidic bond, GALNT = N-acetylgalactosaminyltransferase, ST = sialyltransferase. Numbers within the name of β4-N-acetylgalactosaminyltransferase 1, B4GALNT1, represent the glycosidic bond between carbon C1 of the N-acetylgalactosaminyl residue and C4 of the galactosyl residue of LacCer in Gg3Cer, shown precisely in <a href="#cimb-46-00608-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>The interplay of glycosphingolipid metabolism with other metabolic pathways, which results in distinct metabolite and GSL expression findings in MDA-MB-231 and MCF-7 breast cancer cells after thieno[2,3-<span class="html-italic">b</span>]pyridine derivative treatment. Blue arrows and blue letters indicate the direction of metabolic reactions in MCF-7 cells, while red arrows and red letters indicate reactions in MDA-MB-231 cells. The red arrow from Ac-CoA to the citrate molecule indicates its catabolism in the citric acid cycle for aerobic energy production in the MDA-MB-231 cell line. Black arrows indicate common reactions for both cell lines; * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
15 pages, 2644 KiB  
Article
The Effect of Chronic Immunosuppressive Regimen Treatment on Apoptosis in the Heart of Rats
by Anna Surówka, Michał Żołnierczuk, Piotr Prowans, Marta Grabowska, Patrycja Kupnicka, Marta Markowska, Zbigniew Szlosser and Karolina Kędzierska-Kapuza
Pharmaceuticals 2024, 17(9), 1188; https://doi.org/10.3390/ph17091188 - 10 Sep 2024
Viewed by 260
Abstract
Chronic immunosuppressive therapy is currently the only effective method to prevent acute rejection of a transplanted organ. Unfortunately, the expected effect of treatment brings a number of grave side effects, one of the most serious being cardiovascular complications. In our study, we wanted [...] Read more.
Chronic immunosuppressive therapy is currently the only effective method to prevent acute rejection of a transplanted organ. Unfortunately, the expected effect of treatment brings a number of grave side effects, one of the most serious being cardiovascular complications. In our study, we wanted to investigate how treatment with commonly used immunosuppressive drugs affects the occurrence of programmed cardiac cell death. For this purpose, five groups of rats were treated with different triple immunosuppressive regimens. Cardiac tissue fragments were subjected to the TUNEL assay to visualize apoptotic cells. The expression of Bcl-2 protein, Bax protein, caspase 3 and caspase 9 was also assessed. This study indicates that all immunosuppressive protocols used chronically at therapeutic doses result in an increased percentage of cells undergoing apoptosis in rat heart tissue. The greatest changes were recorded in the TMG (rats treated with tacrolimus, mycophenolate mofetil and glucocorticosteroids) and CMG (rats treated with cyclosporin A, mycophenolate mofetil and glucocorticosteroids) groups. The TRG (rats treated with rapamycin, tacrolimus and glucocorticosteroids) group showed the lowest percentage of apoptotic cells. The internal apoptosis pathway was confirmed only in the TMG group; in the remaining groups, the results indicate programmed cell death via the receptor pathway. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Representative Western blots and densitometric analysis of Bax protein expression levels (normalized to reference protein) in hearts of C—control group without any medication; TRG—rats treated with rapamycin, tacrolimus, glucocorticosteroids; CRG—rats treated with rapamycin, cyclosporin A, glucocorticosteroids; MRG—rats treated with rapamycin, mycophenolate mofetil, glucocorticosteroids; CMG—rats treated with cyclosporin A, mycophenolate mofetil and glucocorticosteroids; TMG—rats treated with tacrolimus, mycophenolate mofetil and glucocorticosteroids; the results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.01 (Mann–Whitney U test).</p>
Full article ">Figure 2
<p>Representative Western blots and densitometric analysis of Bcl-2 protein expression levels (normalized to reference protein) in hearts of C—control group without any medication; TRG—rats treated with rapamycin, tacrolimus, glucocorticosteroids; CRG—rats treated with rapamycin, cyclosporin A, glucocorticosteroids; MRG—rats treated with rapamycin, mycophenolate mofetil, glucocorticosteroids; CMG—rats treated with cyclosporin A, mycophenolate mofetil and glucocorticosteroids; TMG—rats treated with tacrolimus, mycophenolate mofetil and glucocorticosteroids; the results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05 (Mann–Whitney U test).</p>
Full article ">Figure 3
<p>Representative Western blots and densitometric analysis of caspase 3 protein expression levels (normalized to GAPDH) in hearts of C—control group without any medication; TRG—rats treated with rapamycin, tacrolimus, glucocorticosteroids; CRG—rats treated with rapamycin, cyclosporin A, glucocorticosteroids; MRG—rats treated with rapamycin, mycophenolate mofetil, glucocorticosteroids; CMG—rats treated with cyclosporin A, mycophenolate mofetil and glucocorticosteroids; TMG—rats treated with tacrolimus, mycophenolate mofetil and glucocorticosteroids; the results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05 (Mann–Whitney U test).</p>
Full article ">Figure 4
<p>Representative Western blots and densitometric analysis of caspase 9 protein expression levels (normalized to GAPDH) in hearts of C—control group without any medication; TRG—rats treated with rapamycin, tacrolimus, glucocorticosteroids; CRG—rats treated with rapamycin, cyclosporin A, glucocorticosteroids; MRG—rats treated with rapamycin, mycophenolate mofetil, glucocorticosteroids; CMG—rats treated with cyclosporin A, mycophenolate mofetil and glucocorticosteroids; TMG—rats treated with tacrolimus, mycophenolate mofetil and glucocorticosteroids; the results are expressed as means ± SD. * <span class="html-italic">p</span> &lt; 0.05 (Mann–Whitney U test).</p>
Full article ">Figure 5
<p>Representative light micrographs of the TUNEL-positive cells in the heart of rats in the control (<b>a</b>), TRG (<b>b</b>), CRG (<b>c</b>), MRG (<b>d</b>), CMG (<b>e</b>), TMG (<b>f</b>) groups. TUNEL-positive cells with brown-stained nuclei (yellow arrowheads) were observed. C—control group without any medication; TRG—rats treated with rapamycin, tacrolimus, glucocorticosteroids; CRG—rats treated with rapamycin, cyclosporin A, glucocorticosteroids; MRG—rats treated with rapamycin, mycophenolate mofetil, glucocorticosteroids; CMG—rats treated with cyclosporin A, mycophenolate mofetil and glucocorticosteroids; TMG—rats treated with tacrolimus, mycophenolate mofetil and glucocorticosteroids;. TUNEL—terminal deoxynucleotidyl transferase-mediated dUTP nick end-labeling. Scale bar: 50 µm.</p>
Full article ">Figure 6
<p>Bcl-2 family proteins that inhibit and activate apoptosis.</p>
Full article ">Figure 7
<p>Simplified scheme of the mechanism of apoptosis by the extrinsic and intrinsic pathways. <b>1.</b> Initiation of the intrinsic pathway of programmed cell death. <b>2.</b> Fusion of p53 protein with Bcl-2 and Bcl-X<sub>L</sub> proteins. <b>3.</b> Pro-apoptotic proteins creating pores in the membrane of mitochondrion. <b>4.</b> Cytochrome c releasing. <b>5.</b> Cytochrome c binds with apoptotic protease activating factor 1 (Apaf-1) and caspase 9 to create an apoptosome. <b>6.</b> Fusion of a complex of bound anti-apoptotic proteins with Bid, Bik, Bad, Bak proteins. <b>7.</b> Attachment of the complex to the apoptosome. <b>8.</b> The active complex induces the execution caspase cascade, leading to programmed cell death. <b>9.</b> Pro-apoptotic ligands bind to the death receptor (DR). <b>10.</b> Creation of the death-inducing signaling complex (DISC). <b>11.</b> Formed DISC binds Fas-associated death domain (FADD), leading to caspase 8 activation. <b>12</b>. Activation of caspase 3. <b>13.</b> Induction of the execution caspase pathway and apoptosis. <b>14.</b> Proteolysis of Bid protein to the active form t-Bid. <b>15.</b> Active t-Bid protein migration to mitochondrion.</p>
Full article ">Figure 8
<p>Experimental protocol of drugs used in the current research: C—control group without any medication; TRG—rats treated with rapamycin, tacrolimus, glucocorticosteroids; CRG—rats treated with rapamycin, cyclosporin A, glucocorticosteroids; MRG—rats treated with rapamycin, mycophenolate mofetil, glucocorticosteroids; CMG—rats treated with cyclosporin A, mycophenolate mofetil, and glucocorticosteroids; TMG—rats treated with tacrolimus, mycophenolate mofetil, and glucocorticosteroids; n—number.</p>
Full article ">
16 pages, 1337 KiB  
Article
Embryonic Leucine Promotes Early Postnatal Growth via mTOR Signalling in Japanese Quails
by Sawadi F. Ndunguru, Gebrehaweria K. Reda, Brigitta Csernus, Renáta Knop, James K. Lugata, Csaba Szabó, Ádám Z. Lendvai and Levente Czeglédi
Animals 2024, 14(17), 2596; https://doi.org/10.3390/ani14172596 - 6 Sep 2024
Viewed by 345
Abstract
Nutritional cues during embryonic development can alter developmental trajectories and affect postnatal growth. However, the specific mechanisms by which nutrients influence avian growth remain largely unknown. Amino acids can directly interact with the nutrient-sensing pathways, such as the insulin-like growth factor 1 (IGF-1)/mechanistic [...] Read more.
Nutritional cues during embryonic development can alter developmental trajectories and affect postnatal growth. However, the specific mechanisms by which nutrients influence avian growth remain largely unknown. Amino acids can directly interact with the nutrient-sensing pathways, such as the insulin-like growth factor 1 (IGF-1)/mechanistic target of rapamycin (mTOR) pathways, which are known to regulate growth. We examined the effects of embryonic leucine on gene expression and phenotypic growth in Japanese quails by injecting 2.5 mg leucine or saline (control) into Japanese quail eggs on the tenth day of incubation and incubating them under standard conditions. The treatment groups had similar hatching success and size at hatching. However, between 3 and 7 days post-hatching, quails treated with embryonic leucine showed increased growth in body mass and wing, tarsus, head, and intestinal lengths, lasting up to 21 days. The hepatic expression of IGF1, IGF1R, mTOR, and RPS6K1 was upregulated in leucine-treated quails, while the expression of FOXO1 remained unaffected. In conclusion, a subtle increase in embryonic leucine may induce developmental programming effects in Japanese quail by interacting with the IGF-1/mTOR nutrient-sensing pathway to promote growth. This study highlights the role of embryonic amino acids as crucial nutrients for enhancing growth. It provides valuable insight into nutrient intervention strategies during embryonic development to potentially improve poultry growth performance. Full article
(This article belongs to the Special Issue Amino Acid Nutrition in Poultry)
Show Figures

Figure 1

Figure 1
<p>Leucine injection into Japanese quail eggs increased body mass (<b>A</b>) and head (<b>B</b>), tarsus (<b>C</b>) and wing (<b>D</b>) lengths in chicks post-hatch (see <a href="#app1-animals-14-02596" class="html-app">Table S1. Supplementary Materials</a> for detailed sample size). Asterisks indicate a significant difference between the treatment groups (<span class="html-italic">p</span> &lt; 0.05), and error bars indicate mean ± SE.</p>
Full article ">Figure 2
<p>Leucine injection into Japanese quail eggs increased postnatal intestinal length in 21-day-old chicks. Numbers in the bars indicate sample size (n). The asterisk indicates a significant difference between the treatment groups at <span class="html-italic">p</span> &lt; 0.05, and error bars indicate mean ± SE. Numbers 1 and 21 above the bars indicate the age of chicks in days post-hatch.</p>
Full article ">Figure 3
<p>Bar plots present mRNA expression relative to the reference gene. (<b>A</b>) <span class="html-italic">IGF1</span>, (<b>B</b>) <span class="html-italic">IGF1R</span>, (<b>C</b>) <span class="html-italic">mTOR</span>, (<b>D</b>) <span class="html-italic">RPS6K1</span>, (<b>E</b>) <span class="html-italic">FOXO1</span> expression, light grey bars refers to control group, dark grey bars refers to leucine treatment. Numbers in the bars indicate sample size (n). The numbers at the top denote the age of the Japanese quail chicks at which the samples were collected: (1) day-old chicks and (21) 21-day-old chicks. Asterisks indicate significant differences between the treatment groups (<span class="html-italic">p</span> &lt; 0.05), and error bars indicate mean ± SE.</p>
Full article ">
18 pages, 3681 KiB  
Article
Post-Transcriptional Induction of the Antiviral Host Factor GILT/IFI30 by Interferon Gamma
by Taisuke Nakamura, Mai Izumida, Manya Bakatumana Hans, Shuichi Suzuki, Kensuke Takahashi, Hideki Hayashi, Koya Ariyoshi and Yoshinao Kubo
Int. J. Mol. Sci. 2024, 25(17), 9663; https://doi.org/10.3390/ijms25179663 - 6 Sep 2024
Viewed by 245
Abstract
Gamma-interferon-inducible lysosomal thiol reductase (GILT) plays pivotal roles in both adaptive and innate immunities. GILT exhibits constitutive expression within antigen-presenting cells, whereas in other cell types, its expression is induced by interferon gamma (IFN-γ). Gaining insights into the precise molecular mechanism governing the [...] Read more.
Gamma-interferon-inducible lysosomal thiol reductase (GILT) plays pivotal roles in both adaptive and innate immunities. GILT exhibits constitutive expression within antigen-presenting cells, whereas in other cell types, its expression is induced by interferon gamma (IFN-γ). Gaining insights into the precise molecular mechanism governing the induction of GILT protein by IFN-γ is of paramount importance for adaptive and innate immunities. In this study, we found that the 5′ segment of GILT mRNA inhibited GILT protein expression regardless of the presence of IFN-γ. Conversely, the 3′ segment of GILT mRNA suppressed GILT protein expression in the absence of IFN-γ, but it loses this inhibitory effect in its presence. Although the mTOR inhibitor rapamycin suppressed the induction of GILT protein expression by IFN-γ, the expression from luciferase sequence containing the 3′ segment of GILT mRNA was resistant to rapamycin in the presence of IFN-γ, but not in its absence. Collectively, this study elucidates the mechanism behind GILT induction by IFN-γ: in the absence of IFN-γ, GILT mRNA is constitutively transcribed, but the translation process is hindered by both the 5′ and 3′ segments. Upon exposure to IFN-γ, a translation inhibitor bound to the 3′ segment is liberated, and a translation activator interacts with the 3′ segment to trigger the initiation of GILT translation. Full article
(This article belongs to the Special Issue Viral Infections and Immune Responses)
Show Figures

Figure 1

Figure 1
<p>IFN-γ induces GILT protein expression but not its mRNA in HeLa cells. (<b>A</b>) HeLa cells were treated with 0.2 μg/mL of IFN-γ for indicated time period. GILT and actin proteins were analyzed using western blotting. (<b>B</b>) Fluorescent intensities of GILT, FAT10, IDO1, and IFI6 mRNA in both IFN-γ (0.2 μg/mL)-treated and untreated cells were measured with microarray (Kubo et al., 2022) [<a href="#B12-ijms-25-09663" class="html-bibr">12</a>]. Fold inductions by IFN-γ are also indicated. (<b>C</b>) The copy numbers of GAPDH, GILT, and IFI6 mRNAs were quantified using ddPCR. Normalized copy numbers are presented with error bars indicating standard deviations (<span class="html-italic">n</span> = 3). Significance in difference between specified groups is denoted by the <span class="html-italic">p</span>-value from Student’s t-test. (<b>D</b>) Copy numbers of GILT and GAPDH mRNAs in the cytoplasm and nuclei were measured using ddPCR, and the ratios of normalized copy numbers of GILT cDNAs in the cytoplasmic and nuclear fractions to the total copy numbers of GILT cDNA are indicated (<span class="html-italic">n</span> = 3). Error bars show standard deviations.</p>
Full article ">Figure 2
<p>IFN-γ induces GILT protein expression but not its mRNA in TE671 cells. (<b>A</b>) TE671 cells, JEG3 cells, and PBMCs were treated with 0.2 μg/mL of IFN-γ for 3 days, and cell lysates and total RNA samples were extracted. GILT and actin proteins were analyzed using western blotting using their antibodies. (<b>B</b>) Phosphorylated and total STAT1 proteins were analyzed with western blotting, using their specific antibodies. (<b>C</b>) The copy numbers of GAPDH, GILT, and IFI6 mRNAs in TE671 and JEG3 cells were quantified using ddPCR. Normalized copy numbers are presented with error bars indicating standard deviations (<span class="html-italic">n</span> = 3). Significance in difference between specified groups is denoted by the <span class="html-italic">p</span>-value from Student’s t-test.</p>
Full article ">Figure 3
<p>GILT promoter is not activated by IFN-γ. (<b>A</b>) The promoter/enhancer region of the GILT gene was amplified with PCR, and its nucleotide sequence is indicated. Bold and underlined letters show putative GAS and CAT sequences, respectively. (<b>B</b>) HeLa cells were transfected with expression plasmids for nano luciferase (NanoLuc) under the control of the GILT promoter/enhancer and for firefly luciferase (FLuc) under the control of the GAS sequence from the LMP2 gene and treated with IFN-γ. Cell lysates were prepared from the treated cells 3 days after the treatment. NanoLuc and FLuc activities of the cell lysates were measured (<span class="html-italic">n</span> = 3). Luciferase activity of the untreated cells is always set to 1. Relative luciferase activities of the IFN-γ-treated cells to those of untreated cells are indicated. Error bars show standard deviations. The <span class="html-italic">p</span> value between the FLuc activities in untreated and IFN-γ-treated cells is indicated.</p>
Full article ">Figure 4
<p>Stability of GILT protein is not changed by IFN-γ. (<b>A</b>) The translation inhibitor cycloheximide (100 μM final concentration) was added to HeLa cells transduced with an MLV vector expressing GILT and culture for indicated time periods. Cell lysates from the treated cells were analyzed with western blotting using anti-GILT or anti-actin antibody. (<b>B</b>) The intensities of the mature GILT protein detected in the western blotting analysis were measured. The GILT intensities in the untreated GILT-expressing HeLa cells are always set to 1. Relative intensities to the untreated cells are indicated (<span class="html-italic">n</span> = 3). Error bars indicate standard deviations.</p>
Full article ">Figure 5
<p>Impact of the mTOR inhibitor rapamycin on GILT protein expression, GILT promoter activity. (<b>A</b>) HeLa cells were treated with DMSO, rapamycin, and/or IFN-γ for 3 days. Cell lysates from the treated cells were analyzed with western blotting using anti-GILT and anti-actin antibodies. (<b>B</b>) HeLa cells were treated with DMSO or rapamycin for 3 days because rapamycin was dissolved with DMSO. Cell numbers were counted (<span class="html-italic">n</span> = 3). Error bars indicate standard deviations. (<b>C</b>) HeLa cells were transfected with the expression plasmids for NanoLuc and FLuc under the control of the GILT promoter and GAS, respectively. The transfected cells were treated with DMSO, rapamycin, and/or IFN-γ for 3 days as indicated. NanoLuc and FLuc activities were measured (<span class="html-italic">n</span> = 3). Luciferase activities of the DMSO-treated cells are always set to 1. Relative luciferase activities to those of the DMSO-treated cells are indicated. Error bars indicate standard deviations. The <span class="html-italic">p</span> values of Student’s t-test and ANOVA are shown. (<b>D</b>) The copy numbers of GAPDH, GILT, and IFI6 mRNA were measured using ddPCR. Normalized copy numbers of GILT and IFI6 are indicated (<span class="html-italic">n</span> = 3). Error bars show standard deviations. The <span class="html-italic">p</span>-values of ANOVA and Student’s t-test are indicated.</p>
Full article ">Figure 6
<p>Untranslated regions of the GILT mRNA inhibit GILT protein expression. (<b>A</b>) An expression plasmid containing full-length GILT mRNA was obtained (Full GILT). A DNA fragment containing the 5′ UTR and GILT protein-coding region was amplified and ligated into pcDNA3.1 (5′UTR-GILT). A DNA fragment containing the GILT protein-coding region and 3′ UTR was amplified and ligated to pcDNA3.1 (GILT-3′UTR). (<b>B</b>) HeLa cells were transfected with the Renilla luciferase expression plasmid together with the Full GILT, 5′UTR-GILT, or GILT-3′UTR expression plasmid and then were treated with IFN-γ for 24 h. Cell lysates from the treated cells were analyzed with western blotting using anti-GILT and anti-actin antibodies. (<b>C</b>) Renilla luciferase activities of the cell lysates were measaured (<span class="html-italic">n</span> = 3). Error bars show standard deviations.</p>
Full article ">Figure 7
<p>3′ untranslated region of the GILT mRNA inhibits luciferase protein expression but not in the presence of IFN-γ. (<b>A</b>) The 5′ and 3′ UTRs were fused to the 5′ and 3′ ends of the RLuc-coding region, respectively (CMV-5′UTR-RLuc and CMV-RLuc-3′UTR). (<b>B</b>) These expression plasmids were transfected into HeLa cells. RLuc activities were measured. Luminescence levels are indicated with standard deviations (<span class="html-italic">n</span> = 3). (<b>C</b>) The 3′ half region of the full-length GILT cDNA was linked to the 3′ end of the RLuc-coding region (CMV-RLuc-3′GILT). The 5′ half region of the GILT cDNA was fused to the 5′ end of the RLuc (CMV-5′GILT-RLuc). The resulting DNA fragments were ligated to pcDNA3.1. (<b>D</b>) HeLa cells were transfected with CMV-RLuc, CMV-RLuc-3′GILT, or CMV-5′GILT-RLuc expression plasmid and then were treated with IFN-γ for 2 days. RLuc activities of the treated cells were measured. (<b>E</b>) HeLa cells were transfected with CMV-RLuc, CMV-RLuc-3′GILT, or CMV-5′GILT-RLuc expression plasmid and then were treated with IFN-γ and/or rapamycin for 2 days. Relative luciferase activities to the CMV-RLuc-transfected cells in the absence and presence of IFN-γ are indicated (<span class="html-italic">n</span> = 3). Error bars show standard deviations. The <span class="html-italic">p</span> values of the Student’s t-test and ANOVA are demonstrated.</p>
Full article ">Figure 8
<p>STAT1 phosphorylation is required for the initiation of GILT translation by IFN-γ. (<b>A</b>) TE671 cells were treated with IFN-γ and/or FLU, and cell lysates were prepared. Phosphorylated STAT1, total STAT1, and GILT proteins were analyzed using western blotting. (<b>B</b>) The copy numbers of GAPDH, GILT, and IFI6 mRNA were measured using ddPCR. Normalized copy numbers of GILT and IFI6 are indicated (<span class="html-italic">n</span> = 3). Error bars show standard deviations. The <span class="html-italic">p</span> values of the Student’s t-test and ANOVA are demonstrated.</p>
Full article ">Figure 9
<p>Mechanism of GILT protein expression in the absence and presence of IFN-γ.</p>
Full article ">
24 pages, 2124 KiB  
Review
Enhancing Therapeutic Efficacy of FLT3 Inhibitors with Combination Therapy for Treatment of Acute Myeloid Leukemia
by Malia E. Leifheit, Gunnar Johnson, Timothy M. Kuzel, Jeffrey R. Schneider, Edward Barker, Hyun D. Yun, Celalettin Ustun, Josef W. Goldufsky, Kajal Gupta and Amanda L. Marzo
Int. J. Mol. Sci. 2024, 25(17), 9448; https://doi.org/10.3390/ijms25179448 - 30 Aug 2024
Viewed by 874
Abstract
FMS-like tyrosine kinase 3 (FLT3) mutations are genetic changes found in approximately thirty percent of patients with acute myeloid leukemia (AML). FLT3 mutations in AML represent a challenging clinical scenario characterized by a high rate of relapse, even after allogeneic hematopoietic stem cell [...] Read more.
FMS-like tyrosine kinase 3 (FLT3) mutations are genetic changes found in approximately thirty percent of patients with acute myeloid leukemia (AML). FLT3 mutations in AML represent a challenging clinical scenario characterized by a high rate of relapse, even after allogeneic hematopoietic stem cell transplantation (allo-HSCT). The advent of FLT3 tyrosine kinase inhibitors (TKIs), such as midostaurin and gilteritinib, has shown promise in achieving complete remission. However, a substantial proportion of patients still experience relapse following TKI treatment, necessitating innovative therapeutic strategies. This review critically addresses the current landscape of TKI treatments for FLT3+ AML, with a particular focus on gilteritinib. Gilteritinib, a highly selective FLT3 inhibitor, has demonstrated efficacy in targeting the mutant FLT3 receptor, thereby inhibiting aberrant signaling pathways that drive leukemic proliferation. However, monotherapy with TKIs may not be sufficient to eradicate AML blasts. Specifically, we provide evidence for integrating gilteritinib with mammalian targets of rapamycin (mTOR) inhibitors and interleukin-15 (IL-15) complexes. The combination of gilteritinib, mTOR inhibitors, and IL-15 complexes presents a compelling strategy to enhance the eradication of AML blasts and enhance NK cell killing, offering a potential for improved patient outcomes. Full article
(This article belongs to the Special Issue Leukemia: Present and Future)
Show Figures

Figure 1

Figure 1
<p>Structure of the FLT3 receptor in the inactive and active/mutated forms and associated downstream. FLT3 ligand (FLT3L) binding to the FLT3 ligand causes receptor dimerization and activation of downstream signaling pathways. Mutated FLT3 is constitutively active even in the absence of FLT3L. Constitutive activation causes perpetual activation of downstream signaling through the JAK/STAT5/PIM-1, RAS/MEK/MAPK/ERK, and PI3K/Akt/mTOR pathways, causing unchecked survival and growth of AML blasts. The purple and blue circles represent the Extracellular domain of the FLT3 receptor. The yellow circle with a P inside represents a phosphate. Arrows indicate the downstream pathways that are activated. The figure was generated using Biorender (<a href="https://www.biorender.com/" target="_blank">https://www.biorender.com/</a>).</p>
Full article ">Figure 2
<p>Overview of the four complexes found in the mTOR pathway. Specific components of each complex and their potential physiological functions are denoted in the figure [<a href="#B97-ijms-25-09448" class="html-bibr">97</a>]. While all complexes contain mTOR as part of the central structure, they differ in their composition of activating proteins. mTORC1 contains Raptor, Deptor, mLST8, and PRAS40. mTORC2 contains Rictor, Deptor, mLST8, MSIN1, and Protor. Not much is known about mTORC3; however, it contains ETV7 in its structure. mTORC4, like mTORC1 and mTORC2, contains mLST8, but also mEAK-7 and DNA-PKcs as part of its structure. Unchecked signaling through these complexes causes tumorigenesis and unchecked survival and proliferation. The figure was generated using Biorender.</p>
Full article ">Figure 3
<p>(<b>A</b>) Pro-survival signaling pathways in AML and (<b>B</b>) proposed inhibition mechanisms using FLT3 and mTOR inhibitors. FLT3 mutations, such as ITD and TKD, lead to dysregulation of the FLT3 pathway, causing constitutive activation. Gilteritinib inhibits FLT3 at the ITD in the juxtamembrane domain and the TKD mutation in TKD2. Small-molecule inhibitors of the mTOR complexes include dual PI3K/mTORi, allosteric mTORi, and ATP competitive inhibitors of mTORC1 and mTORC2. The red arrow denotes elevated Akt phosphorylation, and the blunt-end red lines represent negative regulation. Black blunted ends represent mechanisms of therapeutic intervention [<a href="#B113-ijms-25-09448" class="html-bibr">113</a>,<a href="#B114-ijms-25-09448" class="html-bibr">114</a>]. These figures were generated using Biorender.</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>) Pro-survival signaling pathways in AML and (<b>B</b>) proposed inhibition mechanisms using FLT3 and mTOR inhibitors. FLT3 mutations, such as ITD and TKD, lead to dysregulation of the FLT3 pathway, causing constitutive activation. Gilteritinib inhibits FLT3 at the ITD in the juxtamembrane domain and the TKD mutation in TKD2. Small-molecule inhibitors of the mTOR complexes include dual PI3K/mTORi, allosteric mTORi, and ATP competitive inhibitors of mTORC1 and mTORC2. The red arrow denotes elevated Akt phosphorylation, and the blunt-end red lines represent negative regulation. Black blunted ends represent mechanisms of therapeutic intervention [<a href="#B113-ijms-25-09448" class="html-bibr">113</a>,<a href="#B114-ijms-25-09448" class="html-bibr">114</a>]. These figures were generated using Biorender.</p>
Full article ">Figure 4
<p>T cell versus NK cell engagement with AML blasts. The HLA-E peptide complex binds to inhibitory receptor NKG2A/CD94 or activating receptor NKG2C/CD94 on the NK surface. The activating receptor NKG2D binds to ULBPs and MICA/MICB. The figure was generated using Biorender.</p>
Full article ">
18 pages, 6192 KiB  
Article
Sodium Tungstate Promotes Neurite Outgrowth and Confers Neuroprotection in Neuro2a and SH-SY5Y Cells
by Nora Montero-Martin, María D. Girón, José D. Vílchez and Rafael Salto
Int. J. Mol. Sci. 2024, 25(17), 9150; https://doi.org/10.3390/ijms25179150 - 23 Aug 2024
Viewed by 509
Abstract
Sodium tungstate (Na2WO4) normalizes glucose metabolism in the liver and muscle, activating the Mitogen-activated protein kinase/extracellular signal-regulated kinase (MAPK/ERK) pathway. Because this pathway controls neuronal survival and differentiation, we investigated the effects of Na2WO4 in mouse [...] Read more.
Sodium tungstate (Na2WO4) normalizes glucose metabolism in the liver and muscle, activating the Mitogen-activated protein kinase/extracellular signal-regulated kinase (MAPK/ERK) pathway. Because this pathway controls neuronal survival and differentiation, we investigated the effects of Na2WO4 in mouse Neuro2a and human SH-SY5Y neuroblastoma monolayer cell cultures. Na2WO4 promotes differentiation to cholinergic neurites via an increased G1/G0 cell cycle in response to the synergic activation of the Phosphatidylinositol 3-kinase (PI3K/Akt) and ERK1/2 signaling pathways. In Neuro2a cells, Na2WO4 increases protein synthesis by activating the mechanistic target of rapamycin (mTOR) and S6K kinases and GLUT3-mediated glucose uptake, providing the energy and protein synthesis needed for neurite outgrowth. Furthermore, Na2WO4 increased the expression of myocyte enhancer factor 2D (MEF2D), a member of a family of transcription factors involved in neuronal survival and plasticity, through a post-translational mechanism that increases its half-life. Site-directed mutations of residues involved in the sumoylation of the protein abrogated the positive effects of Na2WO4 on the MEF2D-dependent transcriptional activity. In addition, the neuroprotective effects of Na2WO4 were evaluated in the presence of advanced glycation end products (AGEs). AGEs diminished neurite differentiation owing to a reduction in the G1/G0 cell cycle, concomitant with lower expression of MEF2D and the GLUT3 transporter. These negative effects were corrected in both cell lines after incubation with Na2WO4. These findings support the role of Na2WO4 in neuronal plasticity, albeit further experiments using 3D cultures, and animal models will be needed to validate the therapeutic potential of the compound. Full article
(This article belongs to the Special Issue Neuropharmacology and Neurodegenerative Diseases 2.0)
Show Figures

Figure 1

Figure 1
<p>Na<sub>2</sub>WO<sub>4</sub> induces neurite outgrowth in Neuro2a (<b>left column</b>) and SH-SY5Y (<b>right column</b>) cells. Neuro2A and SH-SY5Y cells were incubated with increasing concentrations of Na<sub>2</sub>WO<sub>4</sub>, and neurite outgrowth (plotted as the percentage of neurite-bearing cells to total cells) (<b>a</b>), cell cycle, for the G1/G0 and G2 values there are significant differences between control cells and Na<sub>2</sub>WO<sub>4</sub> treated cells at each incubation time (<b>b</b>), and cell viability (plotted as the percentage of viability at time 0) (<b>c</b>) were measured at different time points. Data are mean ± SEM (<span class="html-italic">n</span> = 8). * <span class="html-italic">p</span> &lt; 0.05 of Na<sub>2</sub>WO<sub>4</sub> incubated cells vs control cells.</p>
Full article ">Figure 2
<p>Na<sub>2</sub>WO<sub>4</sub> increased the number of undifferentiated SA-β-gal-positive cells. Cells were differentiated for 48 h in the presence or absence of 1 mM (Neuro2a) or 0.25 mM (SH-SY5Y) Na<sub>2</sub>WO<sub>4</sub> and the percentage of SA-β-gal-positive cells was measured as described in <a href="#sec3-ijms-25-09150" class="html-sec">Section 3</a>. (<b>a</b>) The percentage of Neuro2a SA-β-gal-positive cells. (<b>b</b>) The percentage of SH-SY5Y SA-β-gal-positive cells. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells.</p>
Full article ">Figure 3
<p>Na<sub>2</sub>WO<sub>4</sub> induces the expression of the molecular markers of neuronal differentiation toward a cholinergic phenotype. (<b>a</b>) Neuro2a cells were incubated in a differentiation medium in the absence or presence of 1 mM sodium tungstate for 24 h. (<b>b</b>) SH-SY5Y cells were incubated in a differentiation medium in the absence or presence of 0.25 mM sodium tungstate for 72 h. Acetylcholinesterase expression (AChE) was measured by Western blotting using a specific antibody against AChE. Results were normalized using GAPDH as the loading control. As described in <a href="#sec3-ijms-25-09150" class="html-sec">Section 3</a>, the mRNA levels of the <span class="html-italic">choline O-acetyltransferase</span> (<span class="html-italic">ChAt</span>), <span class="html-italic">tyrosine hydroxylase</span> (<span class="html-italic">Th</span>), and <span class="html-italic">NR4A2</span> genes were measured by qPCR. Results are expressed as mean ± SEM (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05 vs control cells at 0 h. # <span class="html-italic">p</span> &lt; 0.05 between control and Na<sub>2</sub>WO<sub>4</sub> treated cells at the final incubation time.</p>
Full article ">Figure 4
<p>Na<sub>2</sub>WO<sub>4</sub> activates the PI3K/Akt and ERK1/2 signaling pathways in Neuro2a and SH-SY5Y cells. Na<sub>2</sub>WO<sub>4</sub> was added to Neuro2a (<b>a</b>) or SH-SY5Y (<b>b</b>) cells for 24 h or from 0 to 60 min to determine the phosphorylation time course. Western blotting was performed using specific antibodies against phosphorylated and total Akt and ERK1/2. Results are expressed as mean ± SEM (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells.</p>
Full article ">Figure 5
<p>Na<sub>2</sub>WO<sub>4</sub>-induced neurite outgrowth is mediated by the PI3K/Akt and ERK1/2 signaling pathways in Neuro2a cells. Neuro2a cells were pre-treated with PD98059 10 μM or LY294002 20 μM and then treated with 1 mM Na<sub>2</sub>WO<sub>4</sub> for 24 h (<b>a</b>) or 30 min (<b>b</b>). Inhibitors were maintained during the experiment. (<b>a</b>) The percentage of differentiated cells was determined by analyzing cell morphology (<a href="#app1-ijms-25-09150" class="html-app">Supplementary Materials Figure S4</a>). (<b>b</b>) ERK1/2 and Akt activation were measured by Western blotting, as shown in <a href="#ijms-25-09150-f004" class="html-fig">Figure 4</a>. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells; # <span class="html-italic">p</span> &lt; 0.05 vs Na<sub>2</sub>WO<sub>4</sub> treated cells without inhibitors.</p>
Full article ">Figure 6
<p>Na<sub>2</sub>WO<sub>4</sub>-induced neurite outgrowth is concomitant with increased protein synthesis and glucose uptake in Neuro2a cells. Neuro2a cells were incubated with 1 mM Na<sub>2</sub>WO<sub>4</sub> for 24 h. Then, protein synthesis (<b>a</b>) and 2-deoxyglucose uptake (<b>b</b>) were measured as described in <a href="#sec3-ijms-25-09150" class="html-sec">Section 3</a>. Additionally, GLUT3 levels were determined using Western blotting. The phosphorylation status of mTOR and S6K1 was measured after 30 min incubation with 1 mM Na<sub>2</sub>WO<sub>4</sub>. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells.</p>
Full article ">Figure 7
<p>Na<sub>2</sub>WO<sub>4</sub> increased the expression of the transcription factor MEF2D in Neuro2a cells. (<b>a</b>) Neuro2a cells were incubated with 1 mM Na<sub>2</sub>WO<sub>4</sub> for 24 h, and the protein and mRNA expression levels of MEF2D were measured. The transcriptional activity dependent on the MEF2 promoter was evaluated using pMEF2x4 Eb1 Luc, a luciferase reporter system that contains MEF2 binding sites. (<b>b</b>) CHO-k1 or Neuro2a cells were transfected with plasmids that encode for eGFP or a fusion protein between eGFP and MEF2D. Transfected cells treated or not with 1 mM Na<sub>2</sub>WO<sub>4</sub> were then incubated with 100 µM cycloheximide for 24 h, and eGFP fluorescence was measured as described in <a href="#sec3-ijms-25-09150" class="html-sec">Section 3</a>. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs 0 h control cells; # <span class="html-italic">p</span> &lt; 0.05 vs 24 h control cells.</p>
Full article ">Figure 8
<p>Na<sub>2</sub>WO<sub>4</sub> modulates MEF2D protein stability. (<b>a</b>) CHO-k1 cells were transfected with a plasmid that either expressed mutated or wild-type MEF2D proteins and a plasmid that encoded a luciferase reporter system that contained MEF2 binding sites. (<b>b</b>) Neuro2a cells were co-transfected with a plasmid that either expressed mutated or wild-type MEF2D fused to the GAL4 DNA binding domain and a plasmid encoding a luciferase reporter system under the control of 4 GAL4 binding sites. CHO-k1 and Neuro2a cells were incubated with 1 mM Na<sub>2</sub>WO<sub>4</sub> for 24 h and transcriptional activity was evaluated. In addition, CHO-k1 (<b>a</b>) and Neuro2a (<b>b</b>) cells were transfected with plasmids encoding eGFP or a fusion protein between eGFP and wild-type or mutant MEF2D. Transfected cells were treated or not treated with 1 mM Na<sub>2</sub>WO<sub>4</sub> and then incubated with cycloheximide for 24 h, and eGFP fluorescence was measured as described in <a href="#sec3-ijms-25-09150" class="html-sec">Section 3</a>. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells.</p>
Full article ">Figure 9
<p>Na<sub>2</sub>WO<sub>4</sub> reversed the effects of advanced glycation end products (AGEs) on neurites. Neuro2a and SH-SY5Y cells were incubated with 100 µg/mL of AGEs in the absence or presence of Na<sub>2</sub>WO<sub>4</sub>. (<b>a</b>) Percentage neurite outgrowth in Neuro2a or SH-SY5Y cells incubated with the effectors. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells; # <span class="html-italic">p</span> &lt; 0.05 vs AGEs treated cells. (<b>b</b>) Analysis of the cell cycle, for the G1/G0 and G2 values there are significant differences between control cells and Na<sub>2</sub>WO<sub>4</sub>-treated cells.</p>
Full article ">Figure 10
<p>Na<sub>2</sub>WO<sub>4</sub> protects against advanced glycation end products (AGEs) effects on neurites. Neuro2a and SH-SY5Y cells were incubated with 100 µg/mL of AGEs in the absence or presence of Na<sub>2</sub>WO<sub>4</sub>. (<b>a</b>) MEF2D expression in Neuro2a and SH-SY5Y cells. (<b>b</b>) GLUT3 transporter expression in Neuro2a and SH-SY5Y cells. (<b>c</b>) Processed caspase−3 levels in Neuro2a cells incubated with the effectors. Results represent means ± SEM (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05 vs untreated cells; # <span class="html-italic">p</span> &lt; 0.05 vs AGE-treated cells.</p>
Full article ">
20 pages, 3739 KiB  
Article
Hsp70 Negatively Regulates Autophagy via Governing AMPK Activation, and Dual Hsp70-Autophagy Inhibition Induces Synergetic Cell Death in NSCLC Cells
by Bashar Alhasan, Yana A. Gladova, Dmitry V. Sverchinsky, Nikolai D. Aksenov, Boris A. Margulis and Irina V. Guzhova
Int. J. Mol. Sci. 2024, 25(16), 9090; https://doi.org/10.3390/ijms25169090 - 22 Aug 2024
Viewed by 562
Abstract
Proteostasis mechanisms, such as proteotoxic-stress response and autophagy, are increasingly recognized for their roles in influencing various cancer hallmarks such as tumorigenesis, drug resistance, and recurrence. However, the precise mechanisms underlying their coordination remain not fully elucidated. The aim of this study is [...] Read more.
Proteostasis mechanisms, such as proteotoxic-stress response and autophagy, are increasingly recognized for their roles in influencing various cancer hallmarks such as tumorigenesis, drug resistance, and recurrence. However, the precise mechanisms underlying their coordination remain not fully elucidated. The aim of this study is to investigate the molecular interplay between Hsp70 and autophagy in lung adenocarcinoma cells and elucidate its impact on the outcomes of anticancer therapies in vitro. For this purpose, we utilized the human lung adenocarcinoma A549 cell line and genetically modified it by knockdown of Hsp70 or HSF1, and the H1299 cell line with knockdown or overexpression of Hsp70. In addition, several treatments were employed, including treatment with Hsp70 inhibitors (VER-155008 and JG-98), HSF1 activator ML-346, or autophagy modulators (SAR405 and Rapamycin). Using immunoblotting, we found that Hsp70 negatively regulates autophagy by directly influencing AMPK activation, uncovering a novel regulatory mechanism of autophagy by Hsp70. Genetic or chemical Hsp70 overexpression was associated with the suppression of AMPK and autophagy. Conversely, the inhibition of Hsp70, genetically or chemically, resulted in the upregulation of AMPK-mediated autophagy. We further investigated whether Hsp70 suppression-mediated autophagy exhibits pro-survival- or pro-death-inducing effects via MTT test, colony formation, CellTiter-Glo 3D-Spheroid viability assay, and Annexin/PI apoptosis assay. Our results show that combined inhibition of Hsp70 and autophagy, along with cisplatin treatment, synergistically reduces tumor cell metabolic activity, growth, and viability in 2D and 3D tumor cell models. These cytotoxic effects were exerted by substantially potentiating apoptosis, while activating autophagy via rapamycin slightly rescued tumor cells from apoptosis. Therefore, our findings demonstrate that the combined inhibition of Hsp70 and autophagy represents a novel and promising therapeutic approach that may disrupt the capacity of refractory tumor cells to withstand conventional therapies in NSCLC. Full article
(This article belongs to the Special Issue Lung Cancer: From Molecular Mechanisms to Novel Therapeutics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of heat shock exposure and Hsp70 chemical overexpression on autophagy activation in NSCLC tumor cells. (<b>A</b>,<b>C</b>) Human A549 and H1299 tumor cells were exposed to heat shock for 1 h at 43 °C, and incubated for 5 or 18 h to recover. Then, cells were lysed and subjected to immunoblotting with antibodies against Hsp70, p62, and LC3 I/II. (<b>E</b>) A549 tumor cells were treated with ML346 at a concentration of 10 µM for 24 h and then were lysed for immunoblotting. (<b>B</b>,<b>D</b>,<b>F</b>) Values on the charts represent the relative protein expression in the respective sample, which indicates the ratio between the band intensity of the protein of interest and the band intensity of β-Tubulin. Band intensity was measured using the ImageJ software 1.53. The results were considered statistically significant at * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001. ns: non-significant, HS: Heat shock.</p>
Full article ">Figure 2
<p>The knockdown of HSF1 or Hsp70, as well as the use of VER-155008, activates autophagy, while Hsp70 knock-in inhibits autophagy in NSCLC tumor cells. (<b>A</b>) A549 control cells (sh-scrambled) and A549 (shHsp70 or shHSF1) were cultivated until 70–80% confluency, then lysed and subjected to immunoblotting with antibodies against HSP70, ATG5, p62, and LC3 I/II. (<b>C</b>,<b>D</b>) NSCLC cell lines were treated with VER-155008 at the indicated concentrations for 24 h. After that, cells were lysed and subjected to immunoblotting with antibodies against Hsp70, ATG5, p62, and LC3 I/II. (<b>E</b>) A549 tumor cells were treated with VER-155008 at the indicated concentration and then subjected to Hsp70 substrate-binding assay. (<b>H</b>) H1299 tumor cells with shHsp70 or Hsp70 knock-in were cultivated until 70–80% confluency and then subjected to western blotting using antibodies against Hsp70, ATG5, p62, and LC3 I/II. (<b>B</b>,<b>F</b>,<b>G</b>,<b>I</b>) Values on the charts represent the relative protein expression in the respective sample, which indicates the ratio between the band intensity of the protein of interest and the band intensity of β-Tubulin. Band intensity was measured using the ImageJ software 1.53. Values are the means ± SEM from three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001. ns: non-significant.</p>
Full article ">Figure 3
<p>Effects of rapamycin-induced autophagy or autophagy inhibition by SAR405 on Hsp70 activation. (<b>A</b>,<b>C</b>) A549 and H1299 tumor cells were either pretreated or not with 100 nm rapamycin for 2 h and then exposed to heat shock for 1 h at 43 °C, followed by incubation for 18 h to recover. After that, cells were lysed and subjected to immunoblotting with antibodies against Hsp70. (<b>E</b>,<b>H</b>) A549 and H1299 tumor cell lines were treated with SAR405 at specified concentrations for 24 h. Following treatment, the cells were lysed and analyzed by immunoblotting using antibodies against Hsp70, ATG5, p62, and LC3 I/II. (<b>F</b>) A549 tumor cells were treated with SAR405 at the specified concentration, then they were lysed and subjected to Hsp70 substrate-binding assay. (<b>B</b>,<b>D</b>,<b>G</b>,<b>I</b>) Values on the charts represent the relative protein expression in the respective sample, which indicates the ratio between the band intensity of the protein of interest and the band intensity of β-Tubulin. Band intensity was measured using the ImageJ software 1.53. Values are the means ± SEM from three independent experiments; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001. ns: non-significant.</p>
Full article ">Figure 4
<p>Hsp70 regulates autophagy activation by influencing AMPK/mTOR interplay and ULK1/Beclin1 activation. (<b>A</b>) A representative scheme of the regulation of autophagy imitation by AMPK/mTOR activation and the involvement of ULK1 and VPS34 complexes. (<b>B</b>) A549 tumor cells were treated with ML346 at a concentration of 10 µM for 24 h, and then were lysed for immunoblotting with antibodies against p-AMPK, p-mTOR, total ULK1, p-ULK1 Ser555, and p-ULK1 Ser757. (<b>D</b>) A549 control cells (sh-scrambled) and A549 (shHsp70 or shHSF1) or (<b>F</b>) H1299 tumor cells with shHsp70 or Hsp70 knock-in were cultivated until 70–80% confluency, then lysed, and immunoblotting was performed. (<b>H</b>) A549 tumor cells were treated with VER-155008 at the indicated concentrations and then were lysed for immunoblotting. (<b>J</b>) A549 tumor cells were pretreated or not with 100 nM rapamycin for 2 h, then subjected to heat shock at 43 °C for 1 h, followed by an 18 h recovery incubation. Subsequently, the cells were lysed and analyzed by immunoblotting using antibodies against p-AMPK, p-mTOR, total ULK1, p-ULK1 Ser555, and Beclin1. (<b>C</b>,<b>E</b>,<b>G</b>,<b>I</b>,<b>K</b>) Values on the charts represent the relative protein expression in the respective sample, which indicates the ratio between the band intensity of the protein of interest and the band intensity of β-actin. Band intensity was measured using the ImageJ software 1.53. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001. ns: non-significant.</p>
Full article ">Figure 5
<p>Hsp70 inhibition-induced autophagy plays a pro-survival role in NSCLC tumor cells, and combined inhibition of Hsp70 and autophagy with cisplatin synergistically induces apoptosis. (<b>A</b>,<b>B</b>) A549 and H1299 tumor cells were treated with combinations involving the following compounds: cisplatin (10 µM), VER-155008 (10 µM), JG-98 (1 µM), SAR405 (10 µM), CQ (40 µM), and rapamycin (0.1 µM) for 48 h. Afterwards, tumor cell viability was assessed using the MTT test. (<b>C</b>,<b>D</b>) A549 and H1299 tumor cells were treated with combinations of the following compounds: cisplatin (10 µM), JG-98 (1 µM), SAR405 (10 µM), CQ (40 µM), and rapamycin (0.1 µM) for 48 h. Following treatment, cell viability was assessed using the MTT assay. (<b>E</b>–<b>H</b>) A549 and H1299 tumor cells were treated with different combinations of the following compounds: cisplatin (15 µM), VER-155008 (15 µM), SAR405 (15 µM), and rapamycin (0.15 µM) for 48 h. Subsequently, cells were stained with Annexin V/PI for 30 min and directly analyzed using flow cytometry. Values are the means ± SEM from three independent experiments; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001. ns: non-significant.</p>
Full article ">Figure 6
<p>The dual inhibition of Hsp70 and autophagy with cisplatin NSCLC clonogenic potential and 3D tumor spheroids viability. (<b>A</b>) A549 and H1299 tumor cells were treated with combinations of cisplatin (2 µM), VER-155008 (2 µM), SAR405 (2 µM), and rapamycin (20 nM) for 48 h, then the treatments were removed and cells were washed and left to form colonies. After ten days for A549 cells or eight days for H1299 cells, they were fixed with 10% formalin and stained with 0.2% crystal violet. (<b>B</b>–<b>E</b>) A549 and H1299 were seeded on 1% agarose in 96-well plates for 3 days to form spheroids. Afterwards, they were treated with combinations of cisplatin (30 µM), VER-155008 (30 µM), SAR405 (30 µM), JG-98 (3 µM), and rapamycin (0.3 µM) for six days. The spheroids’ morphology was captured (microscope’s magnification ×40), and they were then subjected to a CellTiter-Glo 3D-Spheroid viability assay. Values are the means ± SEM from three independent experiments; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
14 pages, 11156 KiB  
Article
Chromium Affects Mitochondrial Function, Leading to Apoptosis and Autophagy in Turtle Primary Hepatocytes
by Shuqin Lin, Yunjuan Xiao, Jing Lin, Yue Yuan, Haitao Shi, Meiling Hong and Li Ding
Animals 2024, 14(16), 2403; https://doi.org/10.3390/ani14162403 - 19 Aug 2024
Viewed by 441
Abstract
Hexavalent chromium (Cr(VI)), a pervasive industrial contaminant, is highly toxic to both humans and animals. However, its effects on turtles are largely unexplored. Our study aimed to investigate the toxic effects of Cr(VI) on the Reeves’ turtles (Mauremys reevesii) primary hepatocytes. [...] Read more.
Hexavalent chromium (Cr(VI)), a pervasive industrial contaminant, is highly toxic to both humans and animals. However, its effects on turtles are largely unexplored. Our study aimed to investigate the toxic effects of Cr(VI) on the Reeves’ turtles (Mauremys reevesii) primary hepatocytes. We exposed hepatocytes to two concentrations (25 μM and 50 μM) of Cr(VI) for 24 h. The results showed that compared to controls, Cr(VI)-treated cells showed elevated antioxidant enzyme activity (catalase (CAT) and superoxide dismutase (SOD)) and increased reactive oxygen species (ROS) levels. Adenosine triphosphatae (ATP) levels decreased, indicating mitochondrial dysfunction. Additionally, we found significant changes in mitochondrial dynamics related genes, with downregulation of mitofusin 2 (Mfn2) and silent information regulator 1 (SIRT1) and a decrease in sirtuin 3 (SIRT3) and tumor protein 53 (p53) mRNA levels. Annexin V-FITC fluorescence staining-positive cells increased with higher Cr(VI) concentrations, marked by elevated bcl-2-associated X protein (Bax) and cysteinyl aspartate specific proteinase (Caspase3) mRNA levels and reduced B-cell lymphoma-2 (Bcl2) expression. Autophagy-related genes were also affected, with increased microtubule-associated protein 1 light chain 3 (LC3-I), microtubule-associated protein light chain 3II (LC3-II), unc-51-like autophagy-activating kinase 1 (ULK1), and sequestosome 1 (p62/SQSTM1) mRNA levels and decreased mammalian target of rapamycin (mTOR) and Beclin1 expression. Taken together, Cr(VI) promotes cell apoptosis and autophagy in turtle hepatocytes by inducing oxidative stress and disrupting mitochondrial function. These findings highlight the serious health risks posed by Cr(VI) pollution and emphasize the need for protecting wild turtle populations. Full article
(This article belongs to the Special Issue Aquatic Animal Medicine and Pathology)
Show Figures

Figure 1

Figure 1
<p>The primary cell culture and growth curve. (<b>A</b>) After 3 days of culture, it showed a small amount of adherent cell growth. (<b>B</b>) Cells on day 2 after the first passage. (<b>C</b>) Cell morphology after the fifth passage. (<b>D</b>) Cell viability.</p>
Full article ">Figure 2
<p>The effects of Cr(VI) on the cytotoxicity and morphology of hepatocytes. (<b>A</b>) Cytotoxic effect of Cr(VI) on <span class="html-italic">Mauremys reevesii</span> liver cells treated for 24 h. Data expressed as the percentage of cell viability. (<b>B</b>–<b>D</b>) The morphological changes of turtle hepatocytes treated with different concentrations of Cr(VI) for 24 h were observed under a microscope (scale: 100 μm).</p>
Full article ">Figure 3
<p>Effects of Cr(VI) on oxidative stress in turtle hepatocytes. (<b>A</b>) CAT and SOD activity levels. Data are expressed as the mean ± SEM. Values not sharing a common superscript letter differ significantly at <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Cellular ROS detection.</p>
Full article ">Figure 4
<p>The mitochondrial dynamics in turtle hepatocyte. (<b>A</b>) ATP production levels. (<b>B</b>–<b>D</b>) Mitochondrial dynamics-related mRNA level (Mfn2, SIRT1, and SIRT3). Data are expressed as the mean ± SEM. Values not sharing a common superscript letter differ significantly at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Cr(VI) activates the p53 pathway in turtle hepatocytes. Data are expressed as the mean ± SEM. Values not sharing a common superscript letter differ significantly at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Effect of Cr(VI) on the apoptosis of turtle hepatocytes. (<b>A</b>) Fluorescence staining for apoptosis with Annexin V-FITC. (<b>B</b>) Apoptosis-related gene (Bax, Bcl2 and Caspase3) mRNA levels. Data are expressed as the mean ± SEM. Values not sharing a common superscript letter differ significantly at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Effect of Cr(VI) on the autophagy of turtle hepatocytes. Data are expressed as the mean ± SEM. Values not sharing a common superscript letter differ significantly at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
18 pages, 3862 KiB  
Article
A LAT1-Like Amino Acid Transporter Regulates Neuronal Activity in the Drosophila Mushroom Bodies
by Julie Delescluse, Mégane M. Simonnet, Anna B. Ziegler, Kévin Piffaretti, Georges Alves, Yael Grosjean and Gérard Manière
Cells 2024, 13(16), 1340; https://doi.org/10.3390/cells13161340 - 13 Aug 2024
Viewed by 638
Abstract
The proper functioning of neural circuits that integrate sensory signals is essential for individual adaptation to an ever-changing environment. Many molecules can modulate neuronal activity, including neurotransmitters, receptors, and even amino acids. Here, we ask whether amino acid transporters expressed by neurons can [...] Read more.
The proper functioning of neural circuits that integrate sensory signals is essential for individual adaptation to an ever-changing environment. Many molecules can modulate neuronal activity, including neurotransmitters, receptors, and even amino acids. Here, we ask whether amino acid transporters expressed by neurons can influence neuronal activity. We found that minidiscs (mnd), which encodes a light chain of a heterodimeric amino acid transporter, is expressed in different cell types of the adult Drosophila brain: in mushroom body neurons (MBs) and in glial cells. Using live calcium imaging, we found that MND expressed in α/β MB neurons is essential for sensitivity to the L-amino acids: Leu, Ile, Asp, Glu, Lys, Thr, and Arg. We found that the Target Of Rapamycin (TOR) pathway but not the Glutamate Dehydrogenase (GDH) pathway is involved in the Leucine-dependent response of α/β MB neurons. This study strongly supports the key role of MND in regulating MB activity in response to amino acids. Full article
(This article belongs to the Special Issue Molecular Studies of Drosophila Signaling Pathways)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><span class="html-italic">mnd</span> is expressed in the adult <span class="html-italic">Drosophila</span> brain. (<b>a</b>) Schematic representation of the structure of the <span class="html-italic">mnd</span> gene, located on the left arm of the third chromosome (14,985,100–14,991,667 bp). Exons from 1 to 5 are represented by orange boxes and introns by black lines. Three mRNA isoforms were predicted by the ORF analysis from expasy.org: <span class="html-italic">mnd-RD</span> (in blue), <span class="html-italic">mnd-RC</span> (in purple), and <span class="html-italic">mnd-RA</span> (in gray). The 5′ and 3′ untranslated regions are indicated by black boxes. The arrows under the transcripts indicate the primers used for the RT-PCR analysis: green for the common region of all three alternative <span class="html-italic">mnd</span> mRNAs; gray for <span class="html-italic">mnd-RA</span>; purple for <span class="html-italic">mnd-RC</span>; and blue for <span class="html-italic">mnd-RD.</span> Scale bar, 1 kb. (<b>b</b>) <span class="html-italic">mnd</span> expression analyzed by RT-PCR using RNAs extracted from <span class="html-italic">w<sup>1118</sup></span> adult heads and bodies. PCR products were analyzed by electrophoresis on agarose gel. <span class="html-italic">mnd-RC</span> (183 pb) and <span class="html-italic">mnd-RA</span> (111 pb) are both present in the adult heads and bodies, whereas <span class="html-italic">mnd-RD</span> (114 pb) was not detected in heads or bodies. Primers used for RT-PCR are indicated in (<b>a</b>) with colored arrows: green for the common portion of the three alternative <span class="html-italic">mnd</span> mRNAs; gray for <span class="html-italic">mnd-RA</span>; purple for <span class="html-italic">mnd-RC</span>; and blue for <span class="html-italic">mnd-RD.</span> (<b>c</b>) Representative images of double immunostaining with anti-MND (magenta) and anti-GFP (green) on <span class="html-italic">elav-Gal4 &gt; UAS-mCD8::GFP</span> brain (neuronal marker) or on <span class="html-italic">repo-Gal4 &gt; UAS-mCD8::GFP</span> brain (glial cell marker). Red boxes illustrate the area of the MB calyx (upper image) and the cortex glia (lower image) in the brain where the images were recorded. A total of 12 brains for each condition were examined. MND is present in both neurons and glial cells in the adult brain. White indicates the overlap of the two markers on merged images. Scale bar, 50 µm.</p>
Full article ">Figure 2
<p><span class="html-italic">mnd</span> regulatory sequences lead to different expression patterns of MND. <span class="html-italic">mnd</span> regulatory sequences lead to different MND expression patterns, and <span class="html-italic">mnd<sup>24−1</sup></span> tools mimic a part of <span class="html-italic">mnd</span> expression in the MBs. (<b>a</b>) Schematic representation of the structure of the <span class="html-italic">mnd</span> gene. Exons from 1 to 5 are represented by orange boxes and introns by black lines. The upstream regulatory sequences of the <span class="html-italic">mnd</span> gene are represented by the blue and the green box and are designed as <span class="html-italic">mnd<sup>24−1</sup></span> and <span class="html-italic">mnd<sup>25−1</sup></span>, respectively, to generate transgenic driver lines. <span class="html-italic">mnd<sup>24−1</sup></span> is located on the first exon of <span class="html-italic">mnd</span> with the first initial site of transcription, and <span class="html-italic">mnd<sup>25−1</sup></span> is located on the second exon of <span class="html-italic">mnd</span> including the second initial site of transcription. Scale bar, 1 kb. (<b>b</b>) <span class="html-italic">mnd<sup>25−1</sup>-Gal4</span> induces mCD8::GFP expression in the cortex glia. Representative image of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>25−1</sup>-Gal4</span> (<span class="html-italic">mnd<sup>25−1</sup>-Gal4 &gt; UAS-mCD8::GFP</span>) in the whole brain magnified by anti-GFP immunostaining, and in brain cryosection labeled by anti-GFP and anti-nc82 to visualize neuropiles. Eight brains were examined. (<b>c</b>) Collection of representative Z-projections and images of mCD8::GFP driven by <span class="html-italic">mnd<sup>24−1</sup></span> tools and MND labeling. (<b>c1</b>) <span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> induces mCD8::GFP expression in the mushroom body lobes. Representative Z-projections of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> (<span class="html-italic">mnd<sup>24−1</sup>-Gal4 &gt; UAS-mCD8::GFP</span>) in the whole brain, and magnified by anti-GFP immunostaining (<b>c1</b>). Representative Z-projections of double immunostaining in MB lobes: anti-GFP (<b>c1′</b>), anti-MND (<b>c1″</b>), and merge (<b>c1″′</b>). Representative images of mCD8::GFP and MND showing the colocalization in Kenyon cells: anti-GFP (<b>c1*</b>), anti-MND (<b>c1**</b>), and merge (<b>c1***</b>). A total of 34 brains were examined. (<b>c2</b>) <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> induces mCD8::GFP expression in the brain. Representative Z-projection of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA &gt; LexAop-mCD8::GFP</span>) in the whole brain, and magnified by anti-GFP immunostaining (<b>c2</b>), in MB lobes anti-GFP (<b>c2′</b>), anti-MND (<b>c2″</b>), and merge (<b>c2″′</b>). Representative images of mCD8::GFP and MND showing the colocalization in Kenyon cells: anti-GFP (<b>c2*</b>), anti-MND (<b>c2**</b>), and merge (<b>c2***</b>). A total of 25 brains were examined. (<b>c3</b>) Representative Z-projections of mCD8::GFP expression resulting from the genetic intersection between <span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> drivers (<span class="html-italic">mnd<sup>24−1</sup>-Gal4 &gt; UAS &gt; stop &gt; mCD8::GFP/mnd<sup>24−1</sup>-LexA &gt; FLP</span>), and magnified by anti-GFP immunostaining (<b>c3</b>). Both the <span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> drivers are expressed in the MB lobes: anti-GFP (<b>c3′</b>), anti-MND (<b>c3″</b>), and merge (<b>c3″′</b>). Representative images showing colocalization of anti-GFP and anti-MND in Kenyon cells anti-GFP (<b>c3*</b>), anti-MND (<b>c3**</b>), and merge (<b>c3***</b>). A total of 13 brains were examined. For all images, the scale bar is 50 µm.</p>
Full article ">Figure 3
<p><span class="html-italic">mnd</span> is expressed in Kenyon cells forming α/β and ɣ lobes. Genetic intersectional strategy between <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> line, which mimics <span class="html-italic">mnd</span> expression in the brain, and <span class="html-italic">MB-lobe-specific-Gal4</span> driver lines to reveal common cells. (<b>a</b>) Genetic intersectional GFP expression between <span class="html-italic">c739-Gal4</span>, an α/β lobe specific driver, and <span class="html-italic">mnd<sup>24−1</sup>-LexA,</span> and magnified by anti-GFP immunostaining. Representative images of anti-GFP (<b>a1</b>) and anti-MND (<b>a1′</b>) in Kenyon cells (<span class="html-italic">c739-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>), merge (<b>a1″</b>). Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">c739-Gal4</span> (<span class="html-italic">c739-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>) in whole brain (<b>a2</b>). Seven brains were examined. Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-mCD8::GFP</span>) (<b>e</b>). A total of 25 brains were examined. Representative z-projection of mCD8::GFP expression resulting from the intersection of <span class="html-italic">c739-Gal4</span> and <span class="html-italic">mnd<sup>24−1</sup>LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-FLP/c739-Gal4</span> &gt; <span class="html-italic">UAS</span> &gt; <span class="html-italic">stop</span> &gt; <span class="html-italic">mCD8::GFP</span>) (<b>a2e</b>). A total of 11 brains were examined. <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> is expressed in the α/β lobes of the MBs. (<b>b</b>) Genetic intersectional GFP expression between <span class="html-italic">c305a-Gal4</span>, a specific driver of α′/β′ lobes and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span>, and magnified by anti-GFP immunostaining. Representative images of anti-GFP (<b>b1</b>) and anti-MND (<b>b1′</b>) in Kenyon cells (<span class="html-italic">c305a-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>), merge (<b>b1″</b>). Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">c305a-Gal4</span> (<span class="html-italic">c305a-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>) in whole brain (<b>b2</b>). A total of 16 brains were examined. Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-mCD8::GFP</span>) (<b>e</b>). A total of 25 brains were examined. Representative z-projection of mCD8::GFP expression resulting from the intersection of <span class="html-italic">c305a-Gal4</span> and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-FLP/c305a-Gal4</span> &gt; <span class="html-italic">UAS</span> &gt; <span class="html-italic">stop</span> &gt; <span class="html-italic">mCD8::GFP</span>) (<b>b2e</b>). A total of 11 brains were examined. <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> is not expressed in the α′/β′ lobes of the MBs. (<b>c</b>) Genetic intersectional GFP expression between <span class="html-italic">H24-Gal4</span>, a specific driver of ɣ lobes and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span>, and magnified by anti-GFP immunostaining. Representative images of anti-GFP (<b>c1</b>) and anti-MND (<b>c1</b>′) in Kenyon cells (<span class="html-italic">H24-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>), merge (<b>c1″</b>). Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">H24-Gal4</span> (<span class="html-italic">H24-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>) in whole brain (<b>c2</b>). A total of 17 brains were examined. Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-mCD8::GFP</span>) (<b>e</b>). A total of 25 brains were examined. Representative z-projection of mCD8::GFP expression resulting from the intersection of <span class="html-italic">H24-Gal4</span> and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-FLP/H24-Gal4</span> &gt; <span class="html-italic">UAS &gt; stop</span> &gt; <span class="html-italic">mCD8::GFP)</span> (<b>c2e</b>). A total of 12 brains were examined. <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> is weakly expressed in the ɣ lobes. (<b>d</b>) Genetic intersectional GFP expression between <span class="html-italic">OK107-Gal4</span>, a specific driver of all MB lobes, and <span class="html-italic">mnd<sup>24−1</sup>-LexA</span>, and magnified by anti-GFP immunostaining. Representative images of anti-GFP (<b>d1</b>) and anti-MND (<b>d1′</b>) in Kenyon cells (<span class="html-italic">OK107-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>), merge (<b>d1″</b>) Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">OK107-Gal4</span> (<span class="html-italic">OK107-Ga4</span> &gt; <span class="html-italic">UAS-mCD8::GFP</span>) in whole brain (<b>d2</b>). A total of 10 brains were examined. Representative z-projection of mCD8::GFP expression driven by <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> (<span class="html-italic">mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-mCD8::GFP</span>) (<b>e</b>). A total of 25 brains were examined. Representative z-projection of mCD8::GFP expression resulting from the intersection of <span class="html-italic">OK107-Gal4</span> and <span class="html-italic">mnd<sup>24−1</sup>-LexA (mnd<sup>24−1</sup>-LexA</span> &gt; <span class="html-italic">LexAop-FLP/OK107-Gal4</span> &gt; <span class="html-italic">UAS</span> &gt; <span class="html-italic">stop</span> &gt; <span class="html-italic">mCD8::GFP)</span> (<b>d2e</b>). A total of 11 brains were examined. <span class="html-italic">mnd<sup>24−1</sup>-LexA</span> is expressed in the α/β and ɣ lobes of the MBs. For all images the scale bar is 50 µm.</p>
Full article ">Figure 4
<p>MND is required for AA sensing in the MBs. Real-time calcium imaging of ex vivo brains expressing a calcium sensor in the α/β lobes of the MBs in control brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>) or in <span class="html-italic">mnd</span> downregulated brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>;<span class="html-italic">UAS-mnd<sup>dsRNAkk</sup></span>) of 3-day-old flies exposed to each L-AA at a concentration of 20 mM. The different intensities of basal GCaMP6s levels in the vicinity of the two lobes are sometimes difficult to convert to similar rainbow colors, and thus in some brains only one lobe appears in false color. (<b>a</b>) Representative images, in false colors, showing the fluorescence level before (basal activity) and after the addition of either control Ringer’s solution or Glu (20 mM) in control brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>). Scale bar, 50 µm. Line plots of fluorescence changes (∆F/F<sub>0</sub>) in α/β lobe neurons stimulated with Ringer’s solution or 20 mM of Glu for one representative brain. Stimulus application is indicated by a red arrow. (<b>b</b>) Representative images, in false colors, showing the fluorescence level before (basal activity) and after the addition of either control Ringer’s solution or Glu (20 mM) in <span class="html-italic">mnd</span> downregulated brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>;<span class="html-italic">UAS-mnd<sup>dsRNAkk</sup></span>). Scale bar, 50 µm. Line plots of fluorescence changes (∆F/F<sub>0</sub>) in α/β lobe neurons stimulated with Ringer’s solution or 20 mM of Glu for one representative brain. Stimulus application is indicated by a red arrow. (<b>c</b>) Averaged fluorescence intensity of positive or negative peaks ± SEM for control brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>, blue histograms) and for <span class="html-italic">mnd</span> downregulated brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>;<span class="html-italic">UAS-mnd<sup>dsRNAkk</sup></span>, orange histograms) in response to either Ringer’s solution (Ctl) or a specific L-AA at 20 mM. All individual data are shown by dots (<span class="html-italic">n</span> = 10 to 23). For each L-AA, data obtained from <span class="html-italic">mnd</span> downregulated brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>; <span class="html-italic">UAS-mnd<sup>dsRNAkk</sup></span>) were compared to the corresponding control (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>) using a Mann–Whitney test. The absence of * for a given AA indicates that the data are not statistically different between the two conditions. *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001; ****: <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Leucine-mediated activity is TOR-dependent in the MBs. (<b>a</b>) Schematic representation of a Kenyon cell and the putative pathways downstream of leucine activity. (<b>b</b>) Real-time calcium imaging of ex vivo brains expressing a calcium sensor in the α/β lobes of the MBs in control brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>), in GDH-downregulated brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>; <span class="html-italic">UAS-GDH<sup>dsRNAkk</sup></span>), or in TOR downregulated brains (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>; <span class="html-italic">UAS-TOR<sup>TED</sup></span> or <span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>; <span class="html-italic">UAS-TOR<sup>dsRNA</sup></span>) of 3-day-old flies exposed to L-Leu at a concentration of 20 mM. Each histogram represents the averaged fluorescence intensity of peaks ± SEM in α/β lobes of the MBs. All individual data are shown by dots (<span class="html-italic">n</span> = 9 to 13). All data were compared with the control (<span class="html-italic">mnd<sup>24−1</sup>-Gal4</span> &gt; <span class="html-italic">GCaMP6s</span>) by a Mann–Whitney test. ns: not significant, ***: <span class="html-italic">p</span> &lt; 0.001; ****: <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p><b>A</b> model for AA sensing through MND in MBs.</p>
Full article ">
14 pages, 5079 KiB  
Review
Autophagy: Are Amino Acid Signals Dependent on the mTORC1 Pathway or Independent?
by Chenglong Jin, Min Zhu, Jinling Ye, Zhiwen Song, Chuntian Zheng and Wei Chen
Curr. Issues Mol. Biol. 2024, 46(8), 8780-8793; https://doi.org/10.3390/cimb46080519 - 13 Aug 2024
Viewed by 561
Abstract
Autophagy is a kind of “self-eating” phenomenon that is ubiquitous in eukaryotic cells. It mainly manifests in the damaged proteins or organelles in the cell being wrapped and transported by the autophagosome to the lysosome for degradation. Many factors cause autophagy in cells, [...] Read more.
Autophagy is a kind of “self-eating” phenomenon that is ubiquitous in eukaryotic cells. It mainly manifests in the damaged proteins or organelles in the cell being wrapped and transported by the autophagosome to the lysosome for degradation. Many factors cause autophagy in cells, and the mechanism of nutrient-deficiency-induced autophagy has been a research focus. It has been reported that amino-acid-deficiency-induced cellular autophagy is mainly mediated through the mammalian rapamycin target protein complex 1 (mTORC1) signaling pathway. In addition, some researchers also found that non-mTORC1 signaling pathways also regulate autophagy, and the mechanism of autophagy occurrence induced by the deficiency of different amino acids is not precisely the same. Therefore, this review aims to summarize the process of various amino acids regulating cell autophagy and provide a narrative review on the molecular mechanism of amino acids regulating autophagy. Full article
(This article belongs to the Special Issue Latest Review Papers in Molecular Biology 2024)
Show Figures

Figure 1

Figure 1
<p>Different categories of autophagy. Macro-autophagy: proteins, lipids, and organelles such as mitochondria, etc., were surrounded by phagophores (also known as isolation membranes) to build an autophagosome, subsequently fused with lysosomes to form an autolysosome, the internal nutrients, and organelles were accordingly degraded. Micro-autophagy: proteins, lipids, and organelles were swallowed by invagination of the lysosome and late endosome. Chaperone-mediated autophagy (CMA): Cytosolic Hsc70 and cochaperones identified the substrate protein (including a KFERQ-like pentapeptide sequence), then translocated into the lysosomal lumen with lysosomal Lamp-2A. All degradation products were used for protein synthesis, energy production, etc.</p>
Full article ">Figure 2
<p>Regulation of autophagy in response to amino acids. High AMP and stress activate AMPK and depresses mTORC1, and AMPK can directly promote autophagy by phosphorylation of ULK1. Whereas rich amino acids could stimulate mTORC1 activation and suppress autophagy, mTORC1 blocks the occurrence of autophagy by phosphorylation of ULK1 and Atg13. Bcl2 is an inhibitor of autophagy. The interaction between Bcl2 and Beclin1 is broken by Bcl2 phosphorylation. Beclin1 is released and moved to the Vps34-Vps15 complex. Phosphorylation Be-clin1 activates Vps34 kinase activity and production of PI3P and contributes to nascent autophagy. Additionally, phagophore formation needs Atg9 and Atg12-Atg5-Atg1 complex, and LC3-PE also participates in pre-autophagosomal structure (PAS) construction and further forms autophagosomes.</p>
Full article ">
18 pages, 1606 KiB  
Review
mTOR Dysregulation, Insulin Resistance, and Hypertension
by Silviu Marcel Stanciu, Mariana Jinga, Daniela Miricescu, Constantin Stefani, Remus Iulian Nica, Iulia-Ioana Stanescu-Spinu, Ileana Adela Vacaroiu, Maria Greabu and Silvia Nica
Biomedicines 2024, 12(8), 1802; https://doi.org/10.3390/biomedicines12081802 - 8 Aug 2024
Viewed by 679
Abstract
Worldwide, diabetes mellitus (DM) and cardiovascular diseases (CVDs) represent serious health problems associated with unhealthy diet and sedentarism. Metabolic syndrome (MetS) is characterized by obesity, dyslipidemia, hyperglycemia, insulin resistance (IR) and hypertension. The mammalian target of rapamycin (mTOR) is a serine/threonine kinase with [...] Read more.
Worldwide, diabetes mellitus (DM) and cardiovascular diseases (CVDs) represent serious health problems associated with unhealthy diet and sedentarism. Metabolic syndrome (MetS) is characterized by obesity, dyslipidemia, hyperglycemia, insulin resistance (IR) and hypertension. The mammalian target of rapamycin (mTOR) is a serine/threonine kinase with key roles in glucose and lipid metabolism, cell growth, survival and proliferation. mTOR hyperactivation disturbs glucose metabolism, leading to hyperglycemia and further to IR, with a higher incidence in the Western population. Metformin is one of the most used hypoglycemic drugs, with anti-inflammatory, antioxidant and antitumoral properties, having also the capacity to inhibit mTOR. mTOR inhibitors such as rapamycin and its analogs everolimus and temsirolimus block mTOR activity, decrease the levels of glucose and triglycerides, and reduce body weight. The link between mTOR dysregulation, IR, hypertension and mTOR inhibitors has not been fully described. Therefore, the main aim of this narrative review is to present the mechanism by which nutrients, proinflammatory cytokines, increased salt intake and renin–angiotensin–aldosterone system (RAAS) dysregulation induce mTOR overactivation, associated further with IR and hypertension development, and also mTOR inhibitors with higher potential to block the activity of this protein kinase. Full article
Show Figures

Figure 1

Figure 1
<p>Phosphatidylinositol 3-kinase (PI3K) protein kinases B (AKT)/mammalian target of rapamycin (mTOR) pathway in healthy conditions: Nutrients, growth factors, cytokines and insulin bind to tyrosine kinases receptors (RTKs), leading to insulin receptors substrate 1 or 2 (IRS1/IRS2) activation and further AKT activation by phosphorylation. Once activated, AKT will phosphorylate other protein kinases such as mTOR, composed of the two complexes mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2). mTORC1 activates sterol response element binding protein (SREBP) and eukaryotic translation initiator factor 4E binding protein (4EBP) and 70Ka ribosomal protein S56 kinase 1 (p70S6K1), leading to lipid and protein synthesis, respectively. mTORC1 inhibits the activity of unc-51-like kinase 1 (ULK1) and autophagy-related gene 13 (ATG13) blocking autophagy. Inactivation of AKT substrate 160 (AS160) and glycogen synthase 3 (GSK3) induces plasma membrane GLUT translocation. mTORC2 activates other protein kinases such as A, G and C, which positively regulate cellular metabolism. Activation of PI3K/AKT/mTOR will be correlated with cell growth, survival and proliferation. “+” activation.</p>
Full article ">Figure 2
<p>Phosphatidylinositol 3-kinase (PI3K) protein kinase B (AKT)/mammalian target of rapamycin (mTOR) pathway, insulin resistance and hypertension. Hypercaloric diet, branched-chain amino acids (BCAAs), proinflammatory cytokines, free fatty acids (FFAs), increased salt intake, the renin–angiotensin–aldosterone system (RAAS) and salt-inducible kinase (SIK) induce mTOR hyperactivation via RTKs or insulin receptor substrates (IRS1/2). IRS phosphorylation produced by angiotensin II (Ang II) and aldosterone decreases nitric oxide (NO) synthesis. Activation of 70Ka ribosomal protein S6 kinase 1 (p70S6K1) and glycogen synthase 3 (GSK3) inhibits IRS conducing to an increased blood glucose level because GLUT will be blocked inside the cell. mTOR complex 2 (mTORC2) activates serum/glucocorticoid-regulated kinase 1 (SGK1) stimulating Na transport. mTOR over-activation is associated with synthesis of advanced end products (AGEs), reactive oxygen species (ROS), reactive nitrogen species (RNS) and lipids. Metformin has the capacity to inhibit mTOR, while rapamycin, everolimus, temsirolimus and sodium-glucose transporter protein 2 (SGLT2) block mTORC1. All these events will lead to insulin resistance (IR) and further to hypertension. “+” activation; “↓” decrease; “↑” increase.</p>
Full article ">Figure 3
<p>Hyperglycemia is correlated with the activation of the polyol pathway, where glucose is reduced to sorbitol by aldolase reductase (AR) and NADPH. Further, sorbitol will be oxidized to fructose by the enzyme sorbitol dehydrogenase (SDH), and NADH is generated. Fructose will be metabolized into ketone bodies, triose phosphate or carbonylic compounds such as glyoxal, methylglyoxal and 3-deoxyglucose. The last three compounds will contribute to irreversible advanced glycation end product (AGE) formation. NADPH and NADH represent sources for reactive species generation. In hyperglycemic conditions, the hexosamine biosynthesis pathway (HBP) is also activated, leading to the formation of uridine diphosphate N-acetylglucosamine (UDP-GlcNAc), which can induce protein damage. Protein kinase C (PKC) is activated by hyperglycemia, which has the capacity to activate polyol and HBP pathways. PKC activation is associated with decreased levels of nitric oxide (NO) biosynthesis and elevated levels of vascular endothelial growth factor (VEGF). All these molecular events will induce, in the end, mTOR dysregulation. “+” activation; “↑” increase; “↓” decrease.</p>
Full article ">
16 pages, 31397 KiB  
Article
Fucoxanthin Inhibits the Proliferation and Metastasis of Human Pharyngeal Squamous Cell Carcinoma by Regulating the PI3K/Akt/mTOR Signaling Pathway
by Hao-Fei Du, Jia-Min Jiang, Si-Han Wu, Yan-Fang Shi, Hai-Tian Liu, Zheng-Hao Hua, Cai-Sheng Wang, Guo-Ying Qian and Hao-Miao Ding
Molecules 2024, 29(15), 3603; https://doi.org/10.3390/molecules29153603 - 30 Jul 2024
Viewed by 747
Abstract
Human pharyngeal squamous cell carcinoma (HPSCC) is the most common malignancy in the head and neck region, characterized by high mortality and a propensity for metastasis. Fucoxanthin, a carotenoid isolated from brown algae, exhibits pharmacological properties associated with the suppression of tumor proliferation [...] Read more.
Human pharyngeal squamous cell carcinoma (HPSCC) is the most common malignancy in the head and neck region, characterized by high mortality and a propensity for metastasis. Fucoxanthin, a carotenoid isolated from brown algae, exhibits pharmacological properties associated with the suppression of tumor proliferation and metastasis. Nevertheless, its potential to inhibit HPSCC proliferation and metastasis has not been fully elucidated. This study represents the first exploration of the inhibitory effects of fucoxanthin on two human pharyngeal squamous carcinoma cell lines (FaDu and Detroit 562), as well as the mechanisms underlying those effects. The results showed dose-dependent decreases in the proliferation, migration, and invasion of HPSCC cells after fucoxanthin treatment. Further studies indicated that fucoxanthin caused a significant reduction in the expression levels of proteins in the phosphoinositide 3−kinase (PI3K)/protein kinase B (AKT)/mechanistic target of rapamycin (mTOR) pathway, as well as the downstream proteins matrix metalloproteinase (MMP)−2 and MMP−9. Specific activators of PI3K/AKT reversed the effects of fucoxanthin on these proteins, as well as on cell proliferation and metastasis, in FaDu and Detroit 562 cells. Molecular docking assays confirmed that fucoxanthin strongly interacted with PI3K, AKT, mTOR, MMP−2, and MMP−9. Overall, fucoxanthin, a functional food component, is a potential therapeutic agent for HPSCC. Full article
(This article belongs to the Section Natural Products Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of fucoxanthin.</p>
Full article ">Figure 2
<p>Fucoxanthin exhibits significant inhibitory effects on FaDu and Detroit 562 cells. The MTT assay was used to measure the viability of (<b>A</b>) FaDu and (<b>B</b>) Detroit 562 cells after 24 and 48 h of treatment with various concentrations of fucoxanthin. (<b>C</b>) Morphological changes in FaDu and Detroit 562 cells after treatment with various concentrations of fucoxanthin. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Effects of fucoxanthin on the proliferation abilities of FaDu and Detroit 562 cells. (<b>A</b>) EdU fluorescence microscopy was performed to qualitatively and quantitatively assess proliferation in FaDu and Detroit 562 cells after 48 h of treatment with fucoxanthin. (<b>B</b>) Colony formation assays were conducted to evaluate the effects of various concentrations of fucoxanthin on the colony-forming abilities of FaDu and Detroit 562 cells. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Fucoxanthin inhibits the migration and invasion of FaDu and Detroit 562 cells. (<b>A</b>,<b>B</b>) Treatment of FaDu and Detroit 562 cells with various concentrations of fucoxanthin for 24 and 48 h resulted in varying degrees of wound healing inhibition. (<b>C</b>,<b>D</b>) Transwell assays were conducted to assess the effects of fucoxanthin treatment for 24 and 48 h on the migration and invasion abilities of FaDu and Detroit 562 cells. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Fucoxanthin induces cell cycle arrest in FaDu and Detroit 562 cells. (<b>A</b>) Flow cytometry was performed to analyze the cell cycle distribution in fucoxanthin-treated FaDu and Detroit 562 cells after 48 h. (<b>B</b>) The protein expression levels of CDK1, CDK2, CDK4, and cyclin E1 were assessed in fucoxanthin-treated FaDu and Detroit 562 cells after 48 h. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Changes in the PI3K/AKT/mTOR signaling pathway and the protein expression levels of MMP−2 and MMP−9 after FaDu and Detroit 562 cells had been treated with various concentrations of fucoxanthin for 48 h. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Rescue effects of the PI3K/AKT pathway activator, 740 Y−P. (<b>A</b>) Changes in the expression levels of PI3K/AKT/mTOR pathway-associated proteins, as well as MMP−2 and MMP−9, were assessed in FaDu and Detroit 562 cells after treatment with fucoxanthin for 48 h (Fx; 6 μg/mL) and 740 Y−P (50 μg/mL). The proliferation and metastasis of (<b>B</b>) FaDu and (<b>C</b>) Detroit 562 cells, after 48 h of treatment with Fx (6 μg/mL) and 740 Y−P (50 μg/mL), were evaluated with EdU and Transwell assays. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Molecular docking results of fucoxanthin with protein targets. (<b>A</b>) Binding energies of fucoxanthin and targets in molecular docking analyses. (<b>B</b>–<b>F</b>) Molecular docking between fucoxanthin and target proteins (PI3K, AKT, mTOR, MMP−2, and MMP−9). For clarity, only hydrogen bond-interacting residues are labeled. Hydrogen bonding interactions are represented by red dashed lines.</p>
Full article ">
18 pages, 5702 KiB  
Article
KIF1A, R1457Q, and P1688L Mutations Induce Protein Abnormal Aggregation and Autophagy Impairment in iPSC-Derived Motor Neurons
by Mingri Zhao, Junling Wang, Miao Liu, Yaoyao Xu, Jiali Huang, Yiti Zhang, Jianfeng He, Ao Gu, Mujun Liu and Xionghao Liu
Biomedicines 2024, 12(8), 1693; https://doi.org/10.3390/biomedicines12081693 - 30 Jul 2024
Viewed by 608
Abstract
Mutations in the C-terminal of KIF1A (Kinesin family member 1A) may lead to amyotrophic lateral sclerosis (ALS) through unknown mechanisms that are not yet understood. Using iPSC reprogramming technology and motor neuron differentiation techniques, we generated iPSCs from a healthy donor and two [...] Read more.
Mutations in the C-terminal of KIF1A (Kinesin family member 1A) may lead to amyotrophic lateral sclerosis (ALS) through unknown mechanisms that are not yet understood. Using iPSC reprogramming technology and motor neuron differentiation techniques, we generated iPSCs from a healthy donor and two ALS patients with KIF1A mutations (R1457Q and P1688L) and differentiated them into spinal motor neurons (iPSC-MN) to investigate KIF1A-related ALS pathology. Our in vitro iPSC-iMN model faithfully recapitulated specific aspects of the disease, such as neurite fragmentation. Through this model, we observed that these mutations led to KIF1A aggregation at the proximal axon of motor neurons and abnormal accumulation of its transport cargo, LAMP1, resulting in autophagy dysfunction and cell death. RNAseq analysis also indicated that the functions of the extracellular matrix, structure, and cell adhesion were significantly disturbed. Notably, using rapamycin during motor neuron differentiation can effectively prevent motor neuron death. Full article
(This article belongs to the Section Neurobiology and Clinical Neuroscience)
Show Figures

Figure 1

Figure 1
<p>Generation of iPSCs from ALS patient with a pathogenic <span class="html-italic">KIF1A</span> mutation. (<b>A</b>) The iPSC induction protocol. (<b>B</b>) Morphological changes at different stages of iPSC induction from T cell. Scale bar: 50 μm. (<b>C</b>) Detecting foreign plasmids using PCR and agarose gel electrophoresis. (<b>D</b>) Immunofluorescence for early motor neuron markers MNX1 and SMI32 of iPSC expressing markers for OCT4, SOX2, NANOG, SSEA4, and TRA-1-60, the DAPI was used to visualize the nucleus. Scale bar: 75 μm. (<b>E</b>) In vivo teratoma assays with HE staining demonstrated that the iPSCs are capable of differentiating into all three germ layers: endoderm, mesoderm, and ectoderm. Scale bar: 100 μm. (<b>F</b>) A schematic showing the KIF1A protein and its mutations in ALS-related disorders. (<b>G</b>). Analysis of the evolutionary conservation of amino acids in the KIF1A protein that were altered in association with ALS. AA number, amino acid number of human KIF1A protein. (<b>H</b>) G-band karyotyping confirmed the diploid chromosome content in the iPSCs. (<b>I</b>) Sanger DNA sequencing of the iPSC line from the ALS patient revealed a mutation in the <span class="html-italic">KIF1A</span> gene. AA number, amino acid number of human KIF1A protein.</p>
Full article ">Figure 2
<p>iPSC-derived motor neurons with the <span class="html-italic">KIF1A</span> mutation exhibit axonal morphology. (<b>A</b>) Schematic of the protocol for motor neuron differentiation. Abbreviations: NEP, neuroepithelial progenitor; MNP, motor neuron progenitor; iMN, induced motor neuron. (<b>B</b>) Bright-field images of motor neurons from a healthy donor and ALS patient iPSC differentiation. Scale bar: 50 μm. (<b>C</b>) Immunofluorescence for mature motor neuron markers CHAT and SMI32 on day 24. Scale bar: 50 μm. (<b>D</b>) <span class="html-italic">CHAT</span> gene expression in mature motor neurons from a healthy donor and ALS patient was quantified using qRT-PCR. (<b>E</b>) Bright-field images of motor neurons from a healthy donor and ALS patient on day 28 and day 33. Scale bar: 50 μm. (<b>F</b>) Immunofluorescence for mature motor neuron neurofilament by SMI32 on day 33. Normal: yellow arrow heads; Abnormal: white arrow heads. Scale bar: 100 μm. Data are shown as mean ± SD (ns, not significant; Student’s <span class="html-italic">t</span>-test was used for comparison).</p>
Full article ">Figure 3
<p>RNAseq analysis of the motor neurons from <span class="html-italic">KIF1A</span> mutant and healthy donor control. (<b>A</b>) PCA-plot between three sample groups. (<b>B</b>) VENN diagram of common differentially expressed genes between the R1457Q/healthy donor and P1688L/healthy donor groups. (<b>C</b>) a GO analysis of quantitative RNAseq data of the mature motor neurons from <span class="html-italic">KIF1A</span> mutant and healthy donor control. (<b>D</b>) KEGG pathway analysis of common differentially expressed genes. (<b>E</b>) Building the Gene-Act-Network of the KEGG database. Blue color indicates decreased expression levels in <span class="html-italic">KIF1A</span> mutant compared with healthy donor control, while red color indicates increased levels. The depth of the color represents the extent of changes, with log2FC &gt; 0.6 or &lt;−0.6 and FDR &lt; 0.05. (<b>F</b>) Validation of RNA-seq data by qRT-PCR of cell adhesion-associated genes. **** <span class="html-italic">p</span> &lt; 0.0001. (Student’s <span class="html-italic">t</span>-test was used for comparison).</p>
Full article ">Figure 4
<p>Differentiation of iPSCs to single spinal motor neurons. (<b>A</b>) Schematic of the protocol for single motor neuron differentiation. Abbreviations: MNP, motor neuron progenitor; iMN, induced motor neuron; PLO: Poly-L-ornithine. (<b>B</b>) Bright-field images showing the differentiation of MNPs into motor neurons from a healthy donor and an ALS patient. Scale bar: 50 μm. (<b>C</b>) Immunofluorescence for early motor neuron markers MNX1 and SMI32. Scale bar: 100 μm. (<b>D</b>) Quantification of MNX1, n = 5. (<b>E</b>) Bright-field images show the differentiation of motor neuron progenitors into single motor neurons. Scale bar: 50 μm. (<b>F</b>) Cell viability was detected by CCK-8 assay in motor neuron progenitors differentiation after 1 day. (<b>G</b>) Cell viability was detected by CCK-8 assay in motor neuron progenitors differentiation after 6 day. The experiments were performed in triplicate, with values presented as mean ± SD. Normal neurofilament morphology: yellow arrow heads; Abnormal neurofilament morphology: white arrow heads. ** <span class="html-italic">p</span> &lt; 0.01, ns, not significant. (Student’s <span class="html-italic">t</span>-test was used for comparison).</p>
Full article ">Figure 5
<p>Accumulation of mutant KIF1A in the proximal axon. (<b>A</b>) Representative confocal immunofluorescence images of KIF1A (red) in motor neurons, white arrows indicate KIF1A accumulation in the proximal axon. Scale bar: 25 μm. (<b>B</b>) Quantify the accumulation of mutant KIF1A in the proximal axon, n = 5. (<b>C</b>) Colocalization of synaptophysin with KIF1A in motor neurons and white arrows indicates synaptophysin accumulation in the proximal axon. (<b>D</b>) Quantify the accumulation of mutant synaptophysin in the proximal axon, n = 5. Scale bar: 10 μm. Normal distribution of KIF1A or synaptophysin in motor neurons: yellow arrow heads. **** <span class="html-italic">p</span> &lt; 0.0001, ns, not significant. (Student’s <span class="html-italic">t</span>-test was used for comparison).</p>
Full article ">Figure 6
<p>KIF1A mutations induce accumulation of LAMP1 and ATG9A in the proximal region of axons and autophagy dysfunction. (<b>A</b>) Colocalization of LAMP1 with KIF1A in motor neurons and white arrows indicates synaptophysin accumulation in the proximal axon. Scale bar: 50 μm. (<b>B</b>) Quantifies the accumulation of mutant LAMP1 in the proximal axon, n = 5. (<b>C</b>) Colocalization of ATG9A with KIF1A in motor neurons and white arrows indicates synaptophysin accumulation in the proximal axon. Scale bar: 10 μm. (<b>D</b>) Quantifies the accumulation of mutant LAMP1 in the proximal axon, n = 5. (<b>E</b>) Western blot analysis of autophagy-associated proteins. (<b>F</b>) Quantification of (<b>D</b>). (<b>G</b>) The autophagy activator rapamycin can prevent neuronal death. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, ns, not significant. (Student’s <span class="html-italic">t</span>-test was used for comparison).</p>
Full article ">
16 pages, 4106 KiB  
Article
Hypomyelination Leukodystrophy 16 (HLD16)-Associated Mutation p.Asp252Asn of TMEM106B Blunts Cell Morphological Differentiation
by Sui Sawaguchi, Miki Ishida, Yuki Miyamoto and Junji Yamauchi
Curr. Issues Mol. Biol. 2024, 46(8), 8088-8103; https://doi.org/10.3390/cimb46080478 - 27 Jul 2024
Viewed by 454
Abstract
Transmembrane protein 106B (TMEM106B), which is a type II transmembrane protein, is believed to be involved in intracellular dynamics and morphogenesis in the lysosome. TMEM106B is known to be a risk factor for frontotemporal lobar degeneration and has been recently identified as the [...] Read more.
Transmembrane protein 106B (TMEM106B), which is a type II transmembrane protein, is believed to be involved in intracellular dynamics and morphogenesis in the lysosome. TMEM106B is known to be a risk factor for frontotemporal lobar degeneration and has been recently identified as the receptor needed for the entry of SARS-CoV-2, independently of angiotensin-converting enzyme 2 (ACE2). A missense mutation, p.Asp252Asn, of TMEM106B is associated with hypomyelinating leukodystrophy 16 (HLD16), which is an oligodendroglial cell-related white matter disorder causing thin myelin sheaths or myelin deficiency in the central nervous system (CNS). However, it remains to be elucidated how the mutated TMEM106B affects oligodendroglial cells. Here, we show that the TMEM106B mutant protein fails to exhibit lysosome distribution in the FBD-102b cell line, an oligodendroglial precursor cell line undergoing differentiation. In contrast, wild-type TMEM106B was indeed localized in the lysosome. Cells harboring wild-type TMEM106B differentiated into ones with widespread membranes, whereas cells harboring mutated TMEM106B failed to differentiate. It is of note that the output of signaling through the lysosome-resident mechanistic target of rapamycin (mTOR) was greatly decreased in cells harboring mutated TMEM106B. Furthermore, treatment with hesperetin, a citrus flavonoid known as an activator of mTOR signaling, restored the molecular and cellular phenotypes induced by the TMEM106B mutant protein. These findings suggest the potential pathological mechanisms underlying HLD16 and their amelioration. Full article
(This article belongs to the Special Issue Molecules at Play in Neurological Diseases 2024)
Show Figures

Figure 1

Figure 1
<p>TMEM106B protein with the HLD16-associated D252N mutation is widely distributed throughout the cytoplasmic regions. (<b>A</b>) FBD-102b cells (surrounded by white dotted lines) were transfected with the plasmid encoding wild-type (WT) TMEM106B tagged with EGFP at its C-terminus or EGFP-tagged TMEM106B with the D252N mutation. Transfected cells were stained with DAPI for nuclear staining. Scan plots were created along the white dotted lines in the direction of the arrows in the images. (<b>B</b>) Graphs showing fluorescence intensities (F.I., arbitrary units) along the white dotted lines in the direction of the arrows are presented at the bottom of the representative fluorescence images. (<b>C</b>) Cells with abnormal, widely distributed structures were counted and statistically depicted (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 10 fields).</p>
Full article ">Figure 2
<p>Mutated TMEM106B is present in the lysosome. (<b>A</b>) Cells (surrounded by white dotted lines) were transfected with the plasmid encoding mutated TMEM106B (D252N). Transfected cells were stained with the respective antibodies against ER-specific KDEL, Golgi body-specific GM130, and lysosome-resident LAMP1. Scan plots were created along the white dotted lines in the direction of the arrows in the images. (<b>B</b>) Graphs showing fluorescence intensities (F.I., arbitrary units) along the white dotted lines in the direction of the arrows are presented at the bottom of the representative fluorescence images. (<b>C</b>) The respective merged percentages are depicted in bar graphs (<span class="html-italic">n</span> = 3 fields).</p>
Full article ">Figure 3
<p>Cells harboring mutated TMEM106B show decreased cell differentiation abilities. (<b>A</b>) Cells harboring wild-type (WT) or mutated (D252N) TMEM106B were allowed to differentiate for 0 or 5 days. Cells surrounded with dotted red lines in the middle panels are magnified in the right panels. The cell surrounded by a white dotted line is a typically differentiated one with widespread membranes. (<b>B</b>) Differentiated cells are statistically depicted (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 10 fields). (<b>C</b>) The lysates of cells at 5 days following the induction of differentiation were immunoblotted with the respective antibodies against differentiation markers PLP1 and MBP, cell lineage marker Sox10, and internal control actin. (<b>D</b>) Quantification of immunoreactive bands, using control immunoreactive bands as 100%, is depicted in the respective graphs of PLP1, MBP, Sox10, and actin (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 3 blots).</p>
Full article ">Figure 4
<p>Cells harboring mutated TMEM106B show decreased phosphorylation levels of ribosomal S6 and translational 4E-BP1 proteins. (<b>A</b>) The lysates of cells at 5 days following the induction of differentiation were immunoblotted with the respective antibodies against phosphorylated ribosomal S6 and translational 4E-BP1 proteins (pS6 and p4E-BP1). Total ribosomal S6 and translational 4E-BP1 protein (S6 and 4E-BP1) bands are also presented. (<b>B</b>) Quantification of immunoreactive bands, using control immunoreactive bands as 100%, is depicted in the respective graphs of pS6, S6, p4E-BP1, and 4E-BP1 (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 3 blots).</p>
Full article ">Figure 5
<p>Hesperetin recovers phenotypes of cells harboring mutated TMEM106B. (<b>A</b>) Cells harboring mutated TMEM106B were allowed to differentiate for 0 or 5 days in the presence or absence of 10 μm hesperetin (DMSO as the vehicle). Cells surrounded with dotted red lines in the middle panels are magnified in the right panels. The cell surrounded by a white dotted line is a typically differentiated one with widespread membranes. (<b>B</b>) Differentiated cells are statistically depicted (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 10 fields). (<b>C</b>) The lysates of cells at 5 days following the induction of differentiation were immunoblotted with the respective antibodies against differentiation markers PLP1 and MBP, cell lineage marker Sox10, and internal control actin. (<b>D</b>) Quantification of immunoreactive bands, using hesperetin plus immunoreactive bands as 100%, is depicted in the respective graphs of PLP1, MBP, Sox10, and actin (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 3 blots).</p>
Full article ">Figure 6
<p>Hesperetin recovers decreased phosphorylation levels of ribosomal S6 and translational 4E-BP1 proteins in cells harboring mutated TMEM106B. (<b>A</b>) The lysates of cells at 5 days following the induction of differentiation in the presence or absence of hesperetin were immunoblotted with the respective antibodies against phosphorylated ribosomal S6 and translational 4E-BP1 proteins (pS6 and p4E-BP1). Total ribosomal S6 and translational 4E-BP1 protein (S6 and 4E-BP1) bands are also presented. (<b>B</b>) Quantification of immunoreactive bands, using hesperetin plus immunoreactive bands as 100%, is depicted in the respective graphs of pS6, S6, p4E-BP1, and 4E-BP1 (** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 3 blots).</p>
Full article ">
20 pages, 6068 KiB  
Article
COL6A3 Exosomes Promote Tumor Dissemination and Metastasis in Epithelial Ovarian Cancer
by Chih-Ming Ho, Ting-Lin Yen, Tzu-Hao Chang and Shih-Hung Huang
Int. J. Mol. Sci. 2024, 25(15), 8121; https://doi.org/10.3390/ijms25158121 - 25 Jul 2024
Viewed by 644
Abstract
Our study explores the role of cancer-derived extracellular exosomes (EXs), particularly focusing on collagen alpha-3 (VI; COL6A3), in facilitating tumor dissemination and metastasis in epithelial ovarian cancer (EOC). We found that COL6A3 is expressed in aggressive ES2 derivatives, SKOV3 overexpressing COL6A3 (SKOV3/COL6A3), and [...] Read more.
Our study explores the role of cancer-derived extracellular exosomes (EXs), particularly focusing on collagen alpha-3 (VI; COL6A3), in facilitating tumor dissemination and metastasis in epithelial ovarian cancer (EOC). We found that COL6A3 is expressed in aggressive ES2 derivatives, SKOV3 overexpressing COL6A3 (SKOV3/COL6A3), and mesenchymal-type ovarian carcinoma stromal progenitor cells (MSC-OCSPCs), as well as their EXs, but not in less aggressive SKOV3 cells or ES2 cells with COL6A3 knockdown (ES2/shCOL6A3). High COL6A3 expression correlates with worse overall survival among EOC patients, as evidenced by TCGA and GEO data analysis. In vitro experiments showed that EXs from MSC-OCSPCs or SKOV3/COL6A3 cells significantly enhance invasion ability in ES2 or SKOV3/COL6A3 cells, respectively (both, p <0.001). In contrast, ES2 cells with ES2/shCOL6A3 EXs exhibited reduced invasion ability (p < 0.001). In vivo, the average disseminated tumor numbers in the peritoneal cavity were significantly greater in mice receiving intraperitoneally injected SKOV3/COL6A3 cells than in SKOV3 cells (p < 0.001). Furthermore, mice intravenously (IV) injected with SKOV3/COL6A3 cells and SKOV3/COL6A3-EXs showed increased lung colonization compared to mice injected with SKOV3 cells and PBS (p = 0.007) or SKOV3/COL6A3 cells and PBS (p = 0.039). Knockdown of COL6A3 or treatment with EX inhibitor GW4869 or rapamycin-abolished COL6A3-EXs may suppress the aggressiveness of EOC. Full article
(This article belongs to the Special Issue The Molecular Basis of Extracellular Vesicles in Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>Characteristics of EV nanoparticle-tracking analyses of the particle sizes of ES2 EVs, ES2TR EVs, ES2 tumorsphere EVs, and ES2TR tumorsphere EVs. The vertical axes in the graphs show the number of EV particles (×10<sup>6</sup>)/mL, and the horizontal axes indicate the particle size (nm).</p>
Full article ">Figure 2
<p>Exosome characterization of cell lines. (<b>right</b>) These are phase-contrast images of #006 and #007 human ovarian carcinoma ascites (<b>upper</b> and <b>middle</b>) and #007 human ovarian carcinoma tissue (<b>lower</b>)-derived cells (P2). The adherent culture conditions were M199 + 10% FBS + 20 ng/mL of EGF + 0.4 μg/mL of hydrocortisone. (<b>left</b>) These are surface expression markers of human ovarian carcinoma ascites and tissue-derived cells with spindle-like mesenchymal-like (MSC-) (<b>right upper</b> and <b>middle</b>) ovarian carcinoma stromal progenitor cells (OCSPCs) and roundish epithelial-like (epi-) (<b>right lower</b>) ovarian-carcinoma-tissue-derived cells from 2 advanced ovarian cancer patients. (<b>left</b>) High expressions of vimentin in MSC-OCSPCs and CK18 and E-cadherin in epi-OCSPCs were noted. High expression of CD44, CD73, CD90, FLT4, CA125, and SSEA4 was noted in both cells.</p>
Full article ">Figure 3
<p>The percentage of positive stemness markers such as CD133, CD117, CD34, and CD105 was consistent in ES2, ES2-COL6A3 shRNA, MSC-OCSPCs, ES2TR160, SKOV3, and SKOV3-COL6A3 cells (<b>upper</b>) and exosomes (EXs) (<b>lower</b>). ES2TR160 and SKOV3-COL6A3 processed the highest percentage of CD133, CD117, and CD34 stemness phenotypes in cells and exosomes.</p>
Full article ">Figure 4
<p>Invasion ability of EOC-cell-line-derived exosomes. Their invasion ability was examined in ES2 (<b>A</b>) and SKOV3 (<b>B</b>) treated with ES2, ES2TR, ES2 tumor sphere, and ES2TR tumor sphere exosomes and not treated with said exosomes. The invasion ability of exosomes from ES2, ES2 TS, ES2TR, and ES2TR TS was more remarkably enhanced in ES2 than in SKOV3 (*** <span class="html-italic">p</span> &lt; 0.001 for ES2 and ** <span class="html-italic">p</span> &lt; 0.01 for SKOV3, respectively).</p>
Full article ">Figure 5
<p>Invasion ability of autocrine and paracrine effects in EOC-cell-line-derived exosomes. (<b>A</b>) Invasion ability was examined in epi-OCSPCs and MSC-OCSPCs treated with ES2, ES2TR, ES2 tumor spheres, and ES2TR tumor sphere exosomes and not treated with said substances. The invasion ability was only significantly enhanced in the MSC-OCSPCs (** <span class="html-italic">p</span> &lt; 0.01) treated with ES2 exosomes, not in epi-OCSPCs. (<b>B</b>) The invasion ability was substantially increased in the MSC-OCSPCs that were treated with ES2 exosomes than in those that were not (* <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001). Simultaneously, the invasion ability was greater in ES2 cells treated with MSC-OCSPC exosomes than in those without MSC-OCSPC exosomes (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Heat map of differential protein expression of EOC exosomes. The heat map shows group 1—ES2 cells and ES2 EXs versus ES2 cells and group 3—MSC-OCSPCs and ES2 EXs versus MSC-OCSPCs, which were examined using LC-MS/MS analyses.</p>
Full article ">Figure 7
<p>COL6A3 expression in EOC cell lines and derived exosomes. (<b>A</b>,<b>B</b>) COL6A3 was expressed in ES2 derivatives, SKOV3/COL6A3, and MSC-OCSPC-derived-exosomes and lysates, while there was no expression in SKOV3- and ES2/shRNA-derived exosomes and cell lysates. (<b>C</b>) The CD9 and CD63 representative exosome markers were seen in ES2-derivative-, SKOV3/COL6A3-, and MSC-OCSPCs-derived exosomes, but CD9 and CD63 were not detected in those cell lysates. (<b>D</b>) Immunostaining of COL6A3 was positive in ovarian serous carcinoma stromal cells, which surrounded cancer cells with negative staining.</p>
Full article ">Figure 8
<p>The invasion ability of overexpressed and knockdown EOC cells with the exosomes derived from them (<b>A</b>) was examined in SKOV3 and SKOV3-COL6A3 cells treated with and without these respective EXs. The invasion ability was significantly greater in SKOV3 and SKOV3-COL6A3 cells treated with the respective EXs than in those without EXs (** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001). (<b>B</b>) Invasion ability was examined in ES2 cells with ES2 EXs and ES2 knockdown COL6A3 cells (ES2-shCOL6A3) with ES2-shCOL6A3 EXs. Invasion ability was significantly greater in ES2 cells with ES2 EXs than in ES2/shCOL6A3 cells with ES2-shCOL6A3 EXs (both, * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>EOC-derived EXs accelerated cancerous peritoneal dissemination. (<b>a</b>) Representative pictures of 6/7 mice IP injected with 1 × 10<sup>6</sup> SKOV3 cells with ES2 exosomes showing disseminated tumors (red arrows) in the peritoneal cavity compared to the 1/3 mice injected with 1 × 10<sup>6</sup> SKOV3 cells with PBS (<span class="html-italic">p</span> = 0.097, as determined using Student’s <span class="html-italic">t</span>-test). The average disseminated tumor numbers in the peritoneal cavity were significantly greater in mice receiving SKOV3 cells with ES2-exosomes than in those administered SKOV3 cells with PBS (** <span class="html-italic">p</span> &lt; 0.01, as determined using Student’s <span class="html-italic">t</span>-test). (<b>b</b>) Representative histologic pictures of disseminated peritoneal tumors are shown at microscopic scales of 40× and 200×.</p>
Full article ">Figure 10
<p>Overexpressed COL6A3 in EOC-derived EXs accelerated cancerous peritoneal dissemination. (<b>a</b>) Average disseminated tumor numbers in the peritoneal cavity were significantly greater in mice receiving SKOV3-overexpressed COL6A3 (SKOV3/COL6A3) than in SKOV3 cells (*** <span class="html-italic">p</span> &lt; 0.001, as determined using Student’s <span class="html-italic">t</span>-test). Red arrows indicated disseminated tumors in the peritoneal cavity. A total of 1/8 of the mice IV injected with 1 × 10<sup>6</sup> SKOV3/COL6A3 cells had colonization in the lung, while this was only the case for 0/32 of the mice injected with 1 × 10<sup>6</sup> SKOV3 cells only (<span class="html-italic">p</span> = 0.043, as determined using Student’s <span class="html-italic">t</span>-test). (<b>b</b>) Representative histologic pictures of the peritoneal tumor and lung colonization are shown at microscopic scales of 40× and 200×. The right lower panel shows the differential body weights of mice among the IP and IV groups treated with SKOV3 cells, SKOV3 cells with ES2 exosomes, and SKOV3/COL6A3 cells.</p>
Full article ">Figure 11
<p>Overexpressed COL6A3 in EOC-derived EXs accelerated lung colonization. In total, 5/8 mice IV injected with 1 × 10<sup>6</sup> SKOV3/COL6A3 cells and 10 μg of SKOV3/COL6A3 exosomes had colonization in the lung, while this was the case for 0/8 mice injected with 1 × 10<sup>6</sup> SKOV3 cells and PBS (<span class="html-italic">p</span> = 0.007, as determined using Student’s <span class="html-italic">t</span>-test) and 1/8 mice IV injected with 1 × 10<sup>6</sup> SKOV3/COL6A3 cells (<span class="html-italic">p</span> = 0.039) (<a href="#ijms-25-08121-f010" class="html-fig">Figure 10</a>b). Histologic pictures of lung colonization tumors are shown at 40× and 200× microscope magnification.</p>
Full article ">Figure 12
<p>The overall survival regarding COL6A3 expression. The overall survival for high-expression COL6A3 in tissue was significantly higher than that of low expression in (<b>a</b>) all subtypes and (<b>b</b>) serous subtypes of EOC patients from TCGA and GEO data. The best cut-off determined by splitting patients into high- and low-expression groups was used as an auto-selection method, which was used to evaluate all possible cut-off values between the lower and upper quartiles of COL6A3 expression levels. The threshold that provided the best separation between the groups regarding survival outcomes was selected. This approach ensured that the cut-off point maximized the statistical power for detecting differences in survival between the high- and low-expression groups.</p>
Full article ">Figure 13
<p>GW4869 and rapamycin decreased the invasion ability of EOC EXs. The invasion ability was inhibited more in ES2 with ES2-treated (<b>A</b>,<b>C</b>) GW4869 (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01, as determined using Student’s <span class="html-italic">t</span>-test) or (<b>B</b>,<b>D</b>) rapamycin exosomes than in ES2 with ES2 exosomes (*** <span class="html-italic">p</span> &lt; 0.001, as determined using Student’s <span class="html-italic">t</span>-test).</p>
Full article ">
Back to TopTop