[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,818)

Search Parameters:
Keywords = lysosome

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 2698 KiB  
Article
ANT-Mediated Inhibition of the Permeability Transition Pore Alleviates Palmitate-Induced Mitochondrial Dysfunction and Lipotoxicity
by Natalia V. Belosludtseva, Anna I. Ilzorkina, Dmitriy A. Serov, Mikhail V. Dubinin, Eugeny Yu. Talanov, Maxim N. Karagyaur, Alexandra L. Primak, Jiankang Liu and Konstantin N. Belosludtsev
Biomolecules 2024, 14(9), 1159; https://doi.org/10.3390/biom14091159 - 15 Sep 2024
Viewed by 348
Abstract
Hyperlipidemia is a major risk factor for vascular lesions in diabetes mellitus and other metabolic disorders, although its basis remains poorly understood. One of the key pathogenetic events in this condition is mitochondrial dysfunction associated with the opening of the mitochondrial permeability transition [...] Read more.
Hyperlipidemia is a major risk factor for vascular lesions in diabetes mellitus and other metabolic disorders, although its basis remains poorly understood. One of the key pathogenetic events in this condition is mitochondrial dysfunction associated with the opening of the mitochondrial permeability transition (MPT) pore, a drop in the membrane potential, and ROS overproduction. Here, we investigated the effects of bongkrekic acid and carboxyatractyloside, a potent blocker and activator of the MPT pore opening, respectively, acting through direct interaction with the adenine nucleotide translocator, on the progression of mitochondrial dysfunction in mouse primary lung endothelial cells exposed to elevated levels of palmitic acid. Palmitate treatment (0.75 mM palmitate/BSA for 6 days) resulted in an 80% decrease in the viability index of endothelial cells, which was accompanied by mitochondrial depolarization, ROS hyperproduction, and increased colocalization of mitochondria with lysosomes. Bongkrekic acid (25 µM) attenuated palmitate-induced lipotoxicity and all the signs of mitochondrial damage, including increased spontaneous formation of the MPT pore. In contrast, carboxyatractyloside (10 μM) stimulated cell death and failed to prevent the progression of mitochondrial dysfunction under hyperlipidemic stress conditions. Silencing of gene expression of the predominate isoform ANT2, similar to the action of carboxyatractyloside, led to increased ROS generation and cell death under conditions of palmitate-induced lipotoxicity in a stably transfected HEK293T cell line. Altogether, these results suggest that targeted manipulation of the permeability transition pore through inhibition of ANT may represent an alternative approach to alleviate mitochondrial dysfunction and cell death in cell culture models of fatty acid overload. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of bongkrekic acid (BA) and carboxyatractyloside (CAT) on the viability of mouse lung endotheliocytes under conditions of normo- (<b>A</b>) and hyperlipidemia (<b>B</b>). (<b>A</b>) Cells were treated with BA and CAT at different concentrations for 48 h, and the cell viability index was quantified. (<b>B</b>) Effect of 25 µM BA and 10 µM CAT on palmitate (PA)-induced lipotoxicity (0.75 mM PA/fatty acid-free BSA complex solution for 6 days) in the mouse lung endothelial cells. Data represent the mean ± SD from 3–4 independent experiments, including at least 25 fields of view.</p>
Full article ">Figure 2
<p>Effect of bongkrekic acid (BA, 25 µM) and carboxyatractyloside (CAT, 10 µM) on production of reactive oxygen species in mouse lung endothelial cells under conditions of palmitate lipotoxicity (0.75 mM PA/fatty acid-free BSA complex solution for 48 h). (<b>A</b>) DCF fluorescence level reflecting ROS production in the cell cytoplasm; (<b>B</b>) MitoSOX red fluorescence, reflecting the production of superoxide anion in the mitochondria. The addition of 10 μM antimycin A (AA), an inhibitor of the respiratory chain complex III, demonstrated the highest level of superoxide anion generation by mitochondria of mouse lung endothelial cells. Data represent the mean ± SD from 3–4 independent experiments, including at least 25 fields of view.</p>
Full article ">Figure 3
<p>Effect of bongkrekic acid (BA, 25 µM) and carboxyatractyloside (CAT, 10 µM) on mitochondrial membrane potential (Δψ) in mouse lung endothelial cells under conditions of palmitate-induced lipotoxicity (0.75 mM PA/fatty acid-free BSA complex solution for 48 h). Data represent the mean ± SD from four independent experiments.</p>
Full article ">Figure 4
<p>MPT pore opening in mouse lung endothelial cells. (<b>A</b>) Typical images of calcein fluorescence in the presence of CoCl<sub>2</sub> in endothelial cells of the experimental groups. Scale bar—25 μm. (<b>B</b>) Intensity of calcein fluorescence in mitochondria of the mouse lung endothelial cells from six experimental groups. Conditions: CTR—BSA solution, PA—0.75 mM palmitate/BSA complex solution. Data represent the mean ± SD from four independent experiments, including at least 25 fields of view.</p>
Full article ">Figure 5
<p>Effect of bongkrekic acid (25 µM) and carboxyatractyloside (10 µM) on the level of colocalization of mitochondria and lysosomes in endothelial cells during palmitate-induced lipotoxicity. (<b>A</b>) Typical fluorescence images of MitoTracker DeepRed FM (red dots) and LysoTracker Green (green dots) and their colocalization are shown. Scale bar—10 μm. (<b>B</b>) Number of mitochondria (%) colocalized with lysosomes in the mouse lung endotheliocytes from four experimental groups. Abbreviations used: BA, bongkrekic acid; CAT, carboxyatractyloside; PA, palmitic acid. Conditions: CTR—BSA solution, PA—0.75 mM palmitate/BSA complex solution. Data represent the mean ± SD from four independent experiments, including at least 10 fields of view.</p>
Full article ">Figure 6
<p>The amount of ANT2 protein in HEK293T cells with normal (WT) and decreased (ANT2-) expression of ANT2 (<b>A</b>). Survival of HEK293T cells with normal and reduced expression of ANT2 under conditions of palmitate PA-induced lipotoxicity (0.5 mM PA/fatty acid-free BSA complex solution for 6 days) (<b>B</b>). Conditions: 0—BSA solution, 0.5 PA—0.5 mM PA/BSA complex solution. Data represent the mean ± SD from four independent experiments, including at least 25 fields of view. Original images of (<b>A</b>) can be found in <a href="#app1-biomolecules-14-01159" class="html-app">supplementary materials (Figure S3)</a>.</p>
Full article ">Figure 7
<p>Changes in production of reactive oxygen species (<b>A</b>) and mitochondrial membrane potential (Δψ) (<b>B</b>) in HEK293T cells with normal (WT) and reduced (ANT2-) expression of ANT2 under conditions of PA-induced lipotoxicity (0.5 mM PA/fatty acid-free BSA complex solution for 48 h). Conditions: 0—BSA solution, 0.5 PA—0.5 mM palmitate/BSA complex solution. Data represent the mean ± SD from four independent experiments, including at least 25 fields of view.</p>
Full article ">Figure 8
<p>MPT pore opening in HEK293T cells with normal (WT) and reduced expression of ANT2 (ANT2-). Intensity of calcein fluorescence in HEK293T cell mitochondria from four experimental groups. Conditions: 0—BSA solution, PA—0.5 mM palmitate/BSA complex solution. Data represent the mean ± SD from 3–4 independent experiments, including at least 25 fields of view.</p>
Full article ">
16 pages, 5499 KiB  
Article
Transcriptomic Differences by RNA Sequencing for Evaluation of New Method for Long-Time In Vitro Culture of Cryopreserved Testicular Tissue for Oncologic Patients
by Cheng Pei, Plamen Todorov, Qingduo Kong, Mengyang Cao, Evgenia Isachenko, Gohar Rahimi, Frank Nawroth, Nina Mallmann-Gottschalk, Wensheng Liu and Volodimir Isachenko
Cells 2024, 13(18), 1539; https://doi.org/10.3390/cells13181539 - 13 Sep 2024
Viewed by 358
Abstract
Background: Earlier studies have established that culturing human ovarian tissue in a 3D system with a small amount of soluble Matrigel (a basement membrane protein) for 7 days in vitro increased gene fusion and alternative splicing events, cellular functions, and potentially impacted gene [...] Read more.
Background: Earlier studies have established that culturing human ovarian tissue in a 3D system with a small amount of soluble Matrigel (a basement membrane protein) for 7 days in vitro increased gene fusion and alternative splicing events, cellular functions, and potentially impacted gene expression. However, this method was not suitable for in vitro culture of human testicular tissue. Objective: To test a new method for long-time in vitro culture of testicular fragments, thawed with two different regimes, with evaluation of transcriptomic differences by RNA sequencing. Methods: Testicular tissue samples were collected, cryopreserved (frozen and thawed), and evaluated immediately after thawing and following one week of in vitro culture. Before in vitro culture, tissue fragments were encapsulated in fibrin. Four experimental groups were formed. Group 1: tissue quickly thawed (in boiling water at 100 °C) and immediately evaluated. Group 2: tissue quickly thawed (in boiling water at 100 °C) and evaluated after one week of in vitro culture. Group 3: tissue slowly thawed (by a physiological temperature 37 °C) and immediately evaluated. Group 4: tissue slowly thawed (by a physiological temperature 37 °C) and evaluated after one week of in vitro culture. Results: There are the fewest differentially expressed genes in the comparison between Group 2 and Group 4. In this comparison, significantly up-regulated genes included C4B_2, LOC107987373, and GJA4, while significantly down-regulated genes included SULT1A4, FBLN2, and CCN2. Differential genes in cells of Group 2 were mainly enriched in KEGG: regulation of actin cytoskeleton, lysosome, proteoglycans in cancer, TGF-beta signaling pathway, focal adhesion, and endocytosis. These Group 2- genes were mainly enriched in GO: spermatogenesis, cilium movement, collagen fibril organization, cell differentiation, meiotic cell cycle, and flagellated spermatozoa motility. Conclusions: Encapsulation of testicular tissue in fibrin and long-time in vitro culture with constant stirring in a large volume of culture medium can reduce the impact of thawing methods on cryopreserved testicular tissue. Full article
Show Figures

Figure 1

Figure 1
<p>Design of experiments.</p>
Full article ">Figure 2
<p>Cryopreserved fragments of testicular tissue from three patients. (<b>A1</b>–<b>C1</b>) Cryopreserved tissue fragments from three patients immediately after thawing (at 100 °C) in a freezing solution (6% dimethyl sulfoxide + 6% ethylene glycol + 0.15 M sucrose). (<b>A2</b>–<b>C2</b>) The same fragments 3 min after the beginning of removal of cryoprotectants in 0.5 M sucrose. (<b>A3</b>–<b>C3</b>) The same fragments in an isotonic solution after the end of removal of cryoprotectants (rehydration) (<b>A4</b>–<b>C4</b>) The same fragments 1 min after the beginning of formation of fibrin granules. Bar = 1.0mm.</p>
Full article ">Figure 3
<p>Cryopreserved fragments of testicular tissue sealed in fibrin granules after long-time in vitro culture. (<b>A5</b>–<b>C5</b>) Tissue fragments from three patients A, B, and C, shown in <a href="#cells-13-01539-f002" class="html-fig">Figure 2</a> (<b>1</b>–<b>8</b>). Photos show the “behavior” of fibrin granules with fragments embedded using various embedding methods after long-time in vitro culture, (unpublished data). (<b>9</b>) The process of embedding a testicular tissue fragment in a fibrin gel: photo demonstrating the friability of the fragment and, consequently, the inevitability of its disintegration during in vitro culture. Bar = 2.0 mm.</p>
Full article ">Figure 4
<p>Hematoxylin-Eosin (HE)-staining of cryopreserved and in vitro cultured testicular tissue. (<b>A1</b>,<b>A2</b>) HE-staining of cells from Group 1 (quick thawing). (<b>B1</b>,<b>B2</b>) HE-staining of Group 2 (quick thawing and in vitro culture). (<b>C1</b>,<b>C2</b>) HE-staining of Group 3 (slow thawing). (<b>D1</b>,<b>D2</b>) HE-staining of Group 4 (slow thawing and in vitro culture). Bar for (<b>A1</b>–<b>D1</b>) = 500 μm, Bar for (<b>A2</b>–<b>D2</b>) = 50 μm. Black arrow indicates the space between basal membrane and cells in seminiferous tubules.</p>
Full article ">Figure 5
<p>Volcano map showing differentially expressed genes (DEGs) between cryopreserved and in vitro cultured testicular tissue. (<b>A</b>) DEG volcano map: Group 2 cells (quick thawing and in vitro culture) vs. Group 1 (quick thawing). (<b>B</b>) DEG volcano map: Group 4 (slow thawing and in vitro culture) vs. group 3 (slow thawing). (<b>C</b>) DEG volcano map: Group 2 (quick thawing and in vitro culture) vs. Group 4 (slow thawing and in vitro culture). (<b>D</b>) DEG volcano map: Group 2 (quick thawing and in vitro culture) and Group 4 (slow thawing and in vitro culture) vs. Group 1 (quick thawing) and Group 3 (slow thawing).</p>
Full article ">Figure 6
<p>Bubble chart of differentially expressed genes (DEGs) displaying KEGG pathways and GO enrichment. (<b>A</b>) KEGG pathway chart from Group 2 (quick thawing and in vitro culture) and from Group 1 (quick thawing). (<b>B</b>) KEGG pathway chart of Group 4 cells (slow thawing and in vitro culture) and Group 3 cells (slow thawing). (<b>C</b>) KEGG pathway chart from Group 2 cells (quick thawing and in vitro culture) and from Group 4 (slow thawing and in vitro culture). (<b>D</b>) KEGG pathway chart of DEG in cells from Group 2 (quick thawing and in vitro culture) and Group 4 (slow thawing and in vitro culture) vs. cells of Group 1 (quick thawing) and Group 3 (slow thawing). (<b>E</b>) GO enrichment bubble chart for cells from Group 2 (quick thawing and in vitro culture) and Group 1 (quick thawing). (<b>F</b>) GO enrichment bubble chart for cells from Group 4 (slow thawing and in vitro culture) and Group 3 (slow thawing). (<b>G</b>) GO enrichment bubble chart for Group 2 (quick thawing and in vitro culture) and Group 4 (slow thawing and in vitro culture). (<b>H</b>) GO enrichment bubble chart for Group 2 (quick thawing and in vitro culture) and Group 4 (slow thawing and in vitro culture) vs. Group 1 (quick thawing) and Group 3 (slow thawing).</p>
Full article ">
18 pages, 4607 KiB  
Article
Inhibition of Autophagy by Berbamine Hydrochloride Mitigates Tumor Immune Escape by Elevating MHC-I in Melanoma Cells
by Jinhuan Xian, Leilei Gao, Zhenyang Ren, Yanjun Jiang, Junjun Pan, Zheng Ying, Zhenyuan Guo, Qingsong Du, Xu Zhao, He Jin, Hua Yi, Jieying Guan and Shan Hu
Cells 2024, 13(18), 1537; https://doi.org/10.3390/cells13181537 - 13 Sep 2024
Viewed by 269
Abstract
Impaired tumor cell antigen presentation contributes significantly to immune evasion. This study identifies Berbamine hydrochloride (Ber), a compound derived from traditional Chinese medicine, as an effective inhibitor of autophagy that enhances antigen presentation in tumor cells. Ber increases MHC-I-mediated antigen presentation in melanoma [...] Read more.
Impaired tumor cell antigen presentation contributes significantly to immune evasion. This study identifies Berbamine hydrochloride (Ber), a compound derived from traditional Chinese medicine, as an effective inhibitor of autophagy that enhances antigen presentation in tumor cells. Ber increases MHC-I-mediated antigen presentation in melanoma cells, improving recognition and elimination by CD8+ T cells. Mutation of Atg4b, which blocks autophagy, also raises MHC-I levels on the cell surface, and further treatment with Ber under these conditions does not increase MHC-I, indicating Ber’s role in blocking autophagy to enhance MHC-I expression. Additionally, Ber treatment leads to the accumulation of autophagosomes, with elevated levels of LC3-II and p62, suggesting a disrupted autophagic flux. Fluorescence staining and co-localization analyses reveal that Ber likely inhibits lysosomal acidification without hindering autophagosome–lysosome fusion. Importantly, Ber treatment suppresses melanoma growth in mice and enhances CD8+ T cell infiltration, supporting its therapeutic potential. Our findings demonstrate that Ber disturbs late-stage autophagic flux through abnormal lysosomal acidification, enhancing MHC-I-mediated antigen presentation and curtailing tumor immune escape. Full article
(This article belongs to the Special Issue Crosstalk of Autophagy and Apoptosis: Recent Advances)
Show Figures

Figure 1

Figure 1
<p>Ber enhances MHC-I-mediated antigen presentation in melanoma cells. (<b>A</b>) Chemical structure of Ber. (<b>B</b>) Flow cytometry analysis showing the mean fluorescence intensity (normalized) of surface MHC-I (H-2K<sup>b</sup>) on B16 melanoma cells after treatment with indicated concentrations of Ber. Histograms represent a representative experiment. (<b>C</b>) Immunofluorescence images of A375 cells treated with Ber, showing an increase in the total number of HLA-positive signals (red fluorescence). Scale bars represent 20 μm. (<b>D</b>) Western blot analysis of MHC-I protein levels (<span class="html-italic">human</span>, HLA-A/B; <span class="html-italic">mouse</span>, H-2K<sup>b</sup>) in melanoma cells under various conditions. Treatments included control (no Ber) and Ber at concentrations of 0.2 μM, 1 μM, and 5 μM. (<b>E</b>) Quantification of cell-surface expression of H-2K<sup>b</sup>-SIINFEKL complex on B16 cells by flow cytometry. The graph on the right displays the percentage of H-2K<sup>b</sup>-SIINFEKL-positive cells. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 indicate levels of statistical significance.</p>
Full article ">Figure 2
<p>Ber potentiates the cytotoxicity of CD8<sup>+</sup> T cells against melanoma cells. (<b>A</b>) Schematic representation of the co-culture setup between CD8<sup>+</sup> T cells and B16-OVA melanoma cells. (<b>B</b>) Analysis of cell proliferation in B16-OVA cells treated with specified concentrations of Ber and co-cultured with CD8<sup>+</sup> T cells, assessed using a crystal violet staining assay. (<b>C</b>) Quantification of IFN-γ production in the co-cultures of CD8<sup>+</sup> T cells and B16-OVA cells, measured by ELISA. (<b>D</b>) Following treatment with Ber, B16-OVA cells were co-cultured with CD8<sup>+</sup> T cells for 48 h; apoptotic cells were then quantified by flow cytometry. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 indicate levels of statistical significance.</p>
Full article ">Figure 3
<p>Ber enhances MHC-I levels in melanoma cells by inhibiting autophagy. (<b>A</b>) Western blot analysis showing concentration-dependent effects of doxycycline (Dox) on H-2K<sup>b</sup> protein expression in B16 melanoma cells harboring a Dox-inducible mTurquoise2-Atg4b(C74A) construct. Dox induces autophagy inhibition in this cell line, enabling the study of autophagy’s role in MHC-I expression. (<b>B</b>) Flow cytometric analysis of cell surface H-2K<sup>b</sup> levels in B16 cells expressing Dox-inducible mTurquoise2-Atg4b(C74A) following treatment with a gradient concentration of Dox. (<b>C</b>) Western blot assay demonstrating protein expression levels of H-2K<sup>b</sup> in B16-OVA cells carrying Dox-inducible mTurquoise2-Atg4b(C74A) treated with Ber (5 µM), Dox (8 µg/mL), or their combination for 24 h. (<b>D</b>) Quantification of surface MHC-I (H-2K<sup>b</sup>) levels in the same cell lines under the same treatment conditions as in (<b>C</b>). (<b>E</b>) Quantification of cell-surface expression of the H-2K<sup>b</sup>-SIINFEKL complex on B16-OVA cells with Dox-inducible mTurquoise2-Atg4b(C74A) treated with a gradient concentration of Dox, assessed by flow cytometry. (<b>F</b>) Flow cytometric analysis of cell-surface expression of the H-2K<sup>b</sup>-SIINFEKL complex in B16-OVA cells with Dox-inducible mTurquoise2-Atg4b(C74A) following treatment with Ber (5 µM), Dox (8 µg/mL), or their combination for 24 h. *** <span class="html-italic">p</span> &lt; 0.001 indicate levels of statistical significance, ns, not significant.</p>
Full article ">Figure 4
<p>Ber inhibits autophagic flux in melanoma cells. (<b>A</b>) Fluorescence microscopy images and quantitative analysis of GFP-LC3 puncta in A375 cells transfected with a GFP-LC3 plasmid, followed by treatment with Ber (5 µM), bafilomycin A1 (Baf, 20 nM) for 24 h, or starvation using Hank’s Balanced Salt Solution (HBSS) for 6 h. Scale bar: 20 µm. (<b>B</b>) Western blot analysis showing the levels of LC3-I/II and p62 proteins in A375 and B16 cells treated with Ber at indicated concentrations for 24 h. (<b>C</b>) Time-course analysis of LC3-I/II and p62 degradation in A375 cells treated with Ber (5 µM) over different durations (0, 2, 6, 12, 24 h). (<b>D</b>) Evaluation of autophagosome formation in A375 cells transfected with GFP-LC3 and treated with or without Ber (5 µM) under starvation conditions (HBSS for 6 h). GFP-LC3 puncta, indicative of autophagosome accumulation, were assessed via fluorescence microscopy. Scale bar: 20 µm. Data are presented as mean ± SD. Statistical significance is indicated by asterisks (* <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001). Chloroquine, CQ.</p>
Full article ">Figure 5
<p>Ber suppresses late-stage autophagy in melanoma cells by inhibiting lysosomal acidification. (<b>A</b>) Fluorescence microscopy analysis of A375 cells transiently transfected with mCherry-GFP-LC3, treated with HBSS (starvation) for 6 h, chloroquine (CQ, 20 μM), or Ber (5 μM). Scale bar: 20 μm. Cells treated with HBSS and CQ served as positive controls for starvation and autophagy inhibition, respectively. (<b>B</b>) Confocal microscopy images of A375 cells transfected with GFP-LC3 plasmids (green) to label autophagosomes, treated with Ber (10 μM) or CQ (20 μM) for 24 h. Lysosomes were stained with LysoBrite™ red. Yellow fluorescence indicates colocalization of lysosomes and autophagosomes. Scale bar: 20 μm. (<b>C</b>) A375 cells transfected with the hLAMP1-mCherry plasmid were treated with Ber (5 μM) or chloroquine (CQ, 20 μM) for 24 h. After immunostaining for LC3-I/II (green), images were captured using confocal microscopy, and the degree of colocalization between hLAMP1-mCherry (red) and LC3 (green) was quantified using ImageJ software. Yellow fluorescence indicates the colocalization of lysosomes and autophagosomes. The red and green traces in the figure represent the arbitrary units (a.u.) of red and green intensities, respectively, within the rectangular region highlighted in the magnified image. Scale bar: 20 μm. (<b>D</b>) Representative images of LysoTracker Red staining in A375 cells treated with Ber (5 μM) or bafilomycin A1 (Baf, 50 nM). Quantification of LysoTracker Red fluorescence intensity indicated a decrease in lysosomal acidity following treatment with Ber or Baf. Scale bar: 20 μm. (<b>E</b>) Western blot analysis of cathepsin maturation in A375 and B16 cells treated with varying concentrations of Ber and different durations of Ber (5 μM) or Baf (50 nM). (<b>F</b>) Electron microscopy of A375 cells treated with Ber (5 μM) for 24 h. Autolysosomes are indicated by red arrowheads. *** <span class="html-italic">p</span> &lt; 0.001 indicate levels of statistical significance.</p>
Full article ">Figure 6
<p>Ber suppresses melanoma tumor growth and enhances CD8<sup>+</sup> T cell infiltration in mice. (<b>A</b>) Subcutaneous tumors were established in C57BL/6 mice by injecting B16 melanoma cells into the right axillary region. Mice were randomly assigned to three groups and received intraperitoneal injections of Ber at specified doses for 14 days. Post-treatment, the tumors were surgically excised for analysis (<span class="html-italic">n</span> = 5 per group). Scale bar: 1 cm. (<b>B</b>) Graph showing the progression of tumor growth in each group over the treatment period, with values expressed as mean ± SEM (<span class="html-italic">n</span> = 5 per group). (<b>C</b>) Tumor weights were measured to assess the efficacy of Ber in inhibiting tumor growth across the treatment groups. (<b>D</b>) Representative immunohistochemistry images displaying CD8<sup>+</sup> T cell staining in tumor tissues from each experimental group, illustrating the degree of immune cell infiltration. Scale bar: 50 μm. (<b>E</b>) Confocal immunofluorescent images and corresponding quantification of CD8<sup>+</sup> T cells in B16 melanoma tissues from mice. Scale bar, 40 μm. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 indicates levels of statistical significance.</p>
Full article ">
16 pages, 2267 KiB  
Article
New Advances in the Study of CMTM6, a Focus on Its Novel Non-Canonical Cellular Locations, and Functions beyond Its Role as a PD-L1 Stabilizer
by Pedro Ivan Urciaga-Gutierrez, Ramon Antonio Franco-Topete, Blanca Estela Bastidas-Ramirez, Fabiola Solorzano-Ibarra, Jose Manuel Rojas-Diaz, Nadia Tatiana Garcia-Barrientos, Ksenia Klimov-Kravtchenko, Martha Cecilia Tellez-Bañuelos, Pablo Cesar Ortiz-Lazareno, Oscar Peralta-Zaragoza, Angelica Meneses-Acosta, Alan Guillermo Alejandre-Gonzalez, Miriam Ruth Bueno-Topete, Jesse Haramati and Susana del Toro-Arreola
Cancers 2024, 16(18), 3126; https://doi.org/10.3390/cancers16183126 - 11 Sep 2024
Viewed by 355
Abstract
CMTM6 is a membrane protein that acts as a regulator of PD-L1, maintaining its expression on the cell surface, and can prevent its lysosome-mediated degradation. It is unknown if CMTM6 is present in the plasma of patients with cervical cancer, and if it [...] Read more.
CMTM6 is a membrane protein that acts as a regulator of PD-L1, maintaining its expression on the cell surface, and can prevent its lysosome-mediated degradation. It is unknown if CMTM6 is present in the plasma of patients with cervical cancer, and if it has non-canonical subcellular localizations in cell lines derived from cervical cancer. Our objective was to determine whether CMTM6 is found in plasma derived from cervical cancer patients and its subcellular localization in cell lines. Patient plasma was separated into exosome-enriched, exosome-free, and total plasma fractions. The levels of CMTM6 in each fraction were determined using ELISA and Western blot. Finally, for the cellular model, HeLa, SiHa, CaSki, and HaCaT were used; the subcellular locations of CMTM6 were determined using immunofluorescence and flow cytometry. Soluble CMTM6 was found to be elevated in plasma from patients with cervical cancer, with a nearly three-fold increase in patients (966.27 pg/mL in patients vs. 363.54 pg/mL in controls). CMTM6 was preferentially, but not exclusively, found in the exosome-enriched plasma fraction, and was positively correlated with exosomal PD-L1; CMTM6 was identified in the membrane, intracellular compartments, and culture supernatant of the cell lines. These results highlight that CMTM6, in its various presentations, may play an important role in the biology of tumor cells and in immune system evasion. Full article
(This article belongs to the Special Issue Extracellular Vesicles (EVs) in Cancer Diagnostics and Therapy)
Show Figures

Figure 1

Figure 1
<p>Total CMTM6 and total PD-L1 in plasma from the healthy donors (HDs) group and the cervical cancer patients (CC) group. (<b>a</b>) Concentrations of CMTM6 in the plasma of HDs (<span class="html-italic">n</span> = 23) and CC (<span class="html-italic">n</span> = 23) were measured by enzyme-linked immunosorbent assay (ELISA). Data are shown as pg/mL of CMTM6. (<b>b</b>) PD-L1 levels in the plasma of HDs (<span class="html-italic">n</span> = 11) and CC (<span class="html-italic">n</span> = 21) are shown as pg/mL of PD-L1. Student’s <span class="html-italic">t</span>-test was used. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 2
<p>CMTM6 and PD-L1 were found in exosomes from the plasma of the healthy donor group (HD) and the group of patients with cervical cancer (CC). (<b>a</b>) CMTM6 is preferentially released in exosomes in both the HD (<span class="html-italic">n</span> = 10) and CC (<span class="html-italic">n</span> = 20) groups; data shown as pg/mL of CMTM6. (<b>b</b>) PD-L1 was elevated in the exosome-enriched plasma fractions from both the HDs (<span class="html-italic">n</span> = 6) and CC (<span class="html-italic">n</span> = 15); data shown as pg/mL. (<b>c</b>) Correlation between exosomal CMTM6 and PD-L1 in HD samples (<span class="html-italic">r</span> = 0.996). (<b>d</b>) Correlation between exosomal CMTM6 and PD-L1 in CC samples (<span class="html-italic">r</span> = 0.8346). (<b>e</b>) CMTM6 was found to increase in the exosomal fractions from HDs and CC patients by Western blot. CD63 was used as an exosomal marker. The uncropped blots are shown in <a href="#app1-cancers-16-03126" class="html-app">Figure S1</a>. Wilcoxon ranked test was used. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> &lt; 0.005.</p>
Full article ">Figure 2 Cont.
<p>CMTM6 and PD-L1 were found in exosomes from the plasma of the healthy donor group (HD) and the group of patients with cervical cancer (CC). (<b>a</b>) CMTM6 is preferentially released in exosomes in both the HD (<span class="html-italic">n</span> = 10) and CC (<span class="html-italic">n</span> = 20) groups; data shown as pg/mL of CMTM6. (<b>b</b>) PD-L1 was elevated in the exosome-enriched plasma fractions from both the HDs (<span class="html-italic">n</span> = 6) and CC (<span class="html-italic">n</span> = 15); data shown as pg/mL. (<b>c</b>) Correlation between exosomal CMTM6 and PD-L1 in HD samples (<span class="html-italic">r</span> = 0.996). (<b>d</b>) Correlation between exosomal CMTM6 and PD-L1 in CC samples (<span class="html-italic">r</span> = 0.8346). (<b>e</b>) CMTM6 was found to increase in the exosomal fractions from HDs and CC patients by Western blot. CD63 was used as an exosomal marker. The uncropped blots are shown in <a href="#app1-cancers-16-03126" class="html-app">Figure S1</a>. Wilcoxon ranked test was used. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> &lt; 0.005.</p>
Full article ">Figure 3
<p>CMTM6 and PD-L1 were found in the lysates of CC-derived cells. (<b>a</b>) CMTM6 expression was found to be higher in the total lysate of CaSki cells compared to those of HeLa and SiHa cells, shown as bands corresponding to the approximate molecular weight reported for CMTM6 and its densitometric analysis. (<b>b</b>) The bands corresponding to the molecular weight reported for PD-L1 were also found in the three cell lines, coinciding again to show that CaSki cells express the highest levels of this protein; the band pattern and its densitometric analysis are shown. In both cases, β-actin was used as a constitutive protein and loading control. The uncropped blots are shown in <a href="#app1-cancers-16-03126" class="html-app">Figure S2</a>. Data are shown as mean ± SD; three independent experiments were performed for each condition. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001, ns: no significance.</p>
Full article ">Figure 4
<p>CMTM6 was found in the cell membrane and intracellularly in cell lines derived from cervical cancer. (<b>a</b>) The percentage of CMTM6-positive cells was determined by flow cytometry. CMTM6 was found both intracellularly and associated with the plasma membranes of all cell lines. Interestingly, the highest percentages of CMTM6, both intracellular and on the membrane, were found in CaSki cells. Data are shown as mean ± SD; three independent experiments were performed for each condition. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001, ns: no significance. (<b>b</b>) Immunofluorescence staining verified that CMTM6 (AF-594, red stain) is found in different subcellular locations such as intracellular (cytoplasm shown with diamond-tipped arrow) and the plasma membrane (closed arrow). The nuclei (open arrow) are stained with DAPI (blue). Note the intracellular staining in SiHa cells, which obscures the nucleus, and the visible nucleus and membrane staining in CaSki cells. Images taken using the 10× objective (left side) and 30× (right side).</p>
Full article ">Figure 5
<p>PD-L1 is present in the cell membranes and cytoplasm of cell lines derived from cervical cancer. (<b>a</b>) The percentages of PD-L1-positive cells were determined by traditional flow cytometry. Interestingly, the highest percentages of membrane PD-L1 were in CaSki cells. Data are shown as mean ± SD; two independent experiments were performed for each condition. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, **** <span class="html-italic">p</span> ≤ 0.0001, ns: no significance. (<b>b</b>) Immunofluorescence was used to verify if PD-L1 (AF-488, green stain) was found in different subcellular locations, such as intracellular (cytoplasm marked with diamond-tipped arrow) and plasma membrane (closed arrow). Nuclei were stained with DAPI (blue). Images taken using the 10× objective (left side) and 30× (right side).</p>
Full article ">Figure 6
<p>CMTM6 released by cell lines derived from cervical cancer. ELISA of culture supernatants to quantify the total levels of supernatant CMTM6. Detectable levels of supernatant CMTM6 were seen in all cell lines; however, HeLa cells were found to release the highest concentration (11,015 pg/mL). Data are shown as mean ± SD; two independent experiments were performed for each condition. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
19 pages, 8400 KiB  
Article
Insights into the Pathobiology of GM1 Gangliosidosis from Single-Nucleus Transcriptomic Analysis of CNS Cells in a Mouse Model
by Sichi Liu, Ting Xie and Yonglan Huang
Int. J. Mol. Sci. 2024, 25(17), 9712; https://doi.org/10.3390/ijms25179712 - 8 Sep 2024
Viewed by 323
Abstract
GM1 gangliosidosis is a lysosomal storage disorder characterized by the accumulation of GM1 ganglioside, leading to severe neurodegeneration and early mortality. The disease primarily affects the central nervous system, causing progressive neurodegeneration, including widespread neuronal loss and gliosis. To gain a deeper understanding [...] Read more.
GM1 gangliosidosis is a lysosomal storage disorder characterized by the accumulation of GM1 ganglioside, leading to severe neurodegeneration and early mortality. The disease primarily affects the central nervous system, causing progressive neurodegeneration, including widespread neuronal loss and gliosis. To gain a deeper understanding of the neuropathology associated with GM1 gangliosidosis, we employed single-nucleus RNA sequencing to analyze brain tissues from both GM1 gangliosidosis model mice and control mice. No significant changes in cell proportions were detected between the two groups of animals. Differential expression analysis revealed cell type-specific changes in gene expression in neuronal and glial cells. Functional analysis highlighted the neurodegenerative processes, oxidative phosphorylation, and neuroactive ligand–receptor interactions as the significantly affected pathways. The contribution of the impairment of neurotransmitter system disruption and neuronal circuitry disruption was more important than neuroinflammatory responses to GM1 pathology. In 16-week-old GM1 gangliosidosis mice, no microglial or astrocyte activation or increased expression of innate immunity genes was detected. This suggested that nerve degeneration did not induce the inflammatory response but rather promoted glial cell clearance. Our findings provide a crucial foundation for understanding the cellular and molecular mechanisms of GM1 gangliosidosis, potentially guiding future therapeutic strategies. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>Construction of a single-cell transcriptomic atlas of brain tissue in GM1 mice. (<b>A</b>) Single-nucleus RNA sequencing (snRNA-seq) profiling workflow. (<b>B</b>) A dot plot demonstrating the classical marker genes used for the identification of a cluster cell type. The dot size reflects the percentage and the color intensity is proportional to the average expression. (<b>C</b>) A Uniform Manifold Approximation and Projection (UMAP) plot showing the six major cell types across the brain tissues, based on data from snRNA-seq. (<b>D</b>) The proportions of the main cell types in both GM1 and WT brains.</p>
Full article ">Figure 2
<p>GM1-related changes in gene expression. (<b>A</b>) A volcano plot showing –log10 (false discovery rate [FDR]) and logFoldChange (FC) values for the differentially expressed genes (DEGs) of the six cell types. (<b>B</b>) A histogram of the Gene Set Enrichment Analysis (GSEA-based pathway enrichment) for the DEGs identified in the major cell types.</p>
Full article ">Figure 2 Cont.
<p>GM1-related changes in gene expression. (<b>A</b>) A volcano plot showing –log10 (false discovery rate [FDR]) and logFoldChange (FC) values for the differentially expressed genes (DEGs) of the six cell types. (<b>B</b>) A histogram of the Gene Set Enrichment Analysis (GSEA-based pathway enrichment) for the DEGs identified in the major cell types.</p>
Full article ">Figure 3
<p>Transcriptional changes in neurons of GM1 gangliosidosis model mice. (<b>A</b>) A dot plot showing Gene Ontology (GO) term enrichment based on the differentially expressed genes (DEGs) identified in neurons. (<b>B</b>) A UMAP plot showing the NEU subclusters. (<b>C</b>) A violin plot showing the expression of marker genes (excitatory neurons [ExNs]: <span class="html-italic">Slc17a7</span>, <span class="html-italic">Slc17a6</span>; inhibitory neurons [InNs]: <span class="html-italic">Gad1</span>, <span class="html-italic">Gad2</span>) in neuronal subclusters. (<b>D</b>) The proportions of the neuronal subclusters in the brains of GM1 and wild-type (WT) mice. (<b>E</b>) A dot plot showing GO term enrichment based on the DEGs identified in subcluster 9.</p>
Full article ">Figure 4
<p>Transcriptional changes in microglia (MG) in GM1 gangliosidosis model mice. (<b>A</b>) A dot plot showing GO term enrichment based on the differentially expressed genes (DEGs) identified in microglia. (<b>B</b>) UMAP plot showing the MG subclusters. (<b>C</b>) The proportions of cells of the MG subclusters in the brains of GM1 and wild-type mice. (<b>D</b>) A UMAP plots showing the expression of marker genes in the MG subclusters. (<b>E</b>) A dot plot showing GO term enrichment based on the DEGs identified in the MG subclusters.</p>
Full article ">Figure 5
<p>Transcriptional changes in astrocytes (ASCs) in GM1 gangliosidosis model mice. (<b>A</b>) A dot plot showing Gene Ontology (GO) term enrichment based on the differentially expressed genes (DEGs) identified in ASCs. (<b>B</b>) A UMAP plot showing the ASC subclusters. (<b>C</b>) The proportions of the ASC subclusters in the brains of GM1 and wild-type mice. (<b>D</b>) A UMAP plots showing the expression of marker genes in the ASC subclusters. (<b>E</b>) Pseudo-time showing the relative positioning of ASCs along the trajectory. (<b>F</b>) A dot plot showing GO term enrichment based on the DEGs identified in the ASC0 subcluster.</p>
Full article ">Figure 6
<p>Transcriptional changes in oligodendroglial lineage cells in GM1 gangliosidosis model mice. (<b>A</b>) A dot plot showing Gene Ontology (GO) term enrichment based on the differentially expressed genes (DEGs) identified in oligodendrocytes (OLGs). (<b>B</b>) A dot plot showing GO term enrichment based on the DEGs identified in oligodendrocyte progenitor cells (OPCs). (<b>C</b>) Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways associated with the DEGs in OLGs. (<b>D</b>) KEGG pathways associated with the DEGs in the OPCs.</p>
Full article ">Figure 6 Cont.
<p>Transcriptional changes in oligodendroglial lineage cells in GM1 gangliosidosis model mice. (<b>A</b>) A dot plot showing Gene Ontology (GO) term enrichment based on the differentially expressed genes (DEGs) identified in oligodendrocytes (OLGs). (<b>B</b>) A dot plot showing GO term enrichment based on the DEGs identified in oligodendrocyte progenitor cells (OPCs). (<b>C</b>) Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways associated with the DEGs in OLGs. (<b>D</b>) KEGG pathways associated with the DEGs in the OPCs.</p>
Full article ">Figure 7
<p>Changes in intercellular communication in GM1 gangliosidosis model mice. (<b>A</b>) The number and strength of ligand–receptor interactions in the GM1 and wild-type (WT) groups. (<b>B</b>) Heatmaps of the interactions change between GM1 and WT groups. (<b>C</b>) Heatmaps of the incoming signaling patterns in the GM1 and WT groups.</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 266
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 ">
17 pages, 4216 KiB  
Article
An Essential Role for Calnexin in ER-Phagy and the Unfolded Protein Response
by Daniel Wolf, Chiara Röder, Michael Sendtner and Patrick Lüningschrör
Cells 2024, 13(17), 1498; https://doi.org/10.3390/cells13171498 - 6 Sep 2024
Viewed by 379
Abstract
ER-phagy is a specialized form of autophagy, defined by the lysosomal degradation of ER subdomains. ER-phagy has been implicated in relieving the ER from misfolded proteins during ER stress upon activation of the unfolded protein response (UPR). Here, we identified an essential role [...] Read more.
ER-phagy is a specialized form of autophagy, defined by the lysosomal degradation of ER subdomains. ER-phagy has been implicated in relieving the ER from misfolded proteins during ER stress upon activation of the unfolded protein response (UPR). Here, we identified an essential role for the ER chaperone calnexin in regulating ER-phagy and the UPR in neurons. We showed that chemical induction of ER stress triggers ER-phagy in the somata and axons of primary cultured motoneurons. Under basal conditions, the depletion of calnexin leads to an enhanced ER-phagy in axons. However, upon ER stress induction, ER-phagy did not further increase in calnexin-deficient motoneurons. In addition to increased ER-phagy under basal conditions, we also detected an elevated proteasomal turnover of insoluble proteins, suggesting enhanced protein degradation by default. Surprisingly, we detected a diminished UPR in calnexin-deficient early cortical neurons under ER stress conditions. In summary, our data suggest a central role for calnexin in orchestrating both ER-phagy and the UPR to maintain protein homeostasis within the ER. Full article
(This article belongs to the Special Issue Endoplasmic Reticulum Stress in Neurodegenerative Diseases)
Show Figures

Figure 1

Figure 1
<p>ER stress triggers local ER-phagy in the axons of primary motoneurons. (<b>A</b>) Blocking protein degradation causes enhanced UPR activation. Western blot showing the levels of Chop after tunicamycin treatment for 8 h and simultaneous treatment with bafilomycin A1 or MG132 during the last 4 h. (<b>B</b>) Quantification of the Chop intensity. The Chop levels were normalized to Gapdh and the tunicamycin condition was set to 1 in each experiment. (<b>C</b>) Expression of mRFP-GFP-KDEL in primary motoneurons showing the distribution of mRFP-GFP-KDEL to the axon and soma. Scale bar: 20 µm. Inset; Scale bar: 5 µm (<b>D</b>) Treatment with tunicamycin causes an increase in Lamp1<sup>+</sup>mRFP<sup>+</sup>GFP<sup>+</sup> vesicles in the axons of primary motoneurons. Scale bar: 20 µm. Inset; Scale bar: 5 µm. (<b>E</b>) In the somata of cultured motoneurons, treatment with tunicamycin resulted in an increase in Lamp1<sup>+</sup>mRFP<sup>+</sup>GFP<sup>+</sup> and Lamp1<sup>+</sup>mRFP<sup>+</sup>GFP<sup>−</sup> vesicles. Arrowheads point to mRFP<sup>+</sup>GFP<sup>+</sup> punctae. Arrows point to mRFP<sup>+</sup>GFP<sup>−</sup> punctae. The dashed lines label the border between the cytosol and the nucleus. Scale bar: 2.5 µm (<b>F</b>,<b>G</b>) Quantification of mRFP<sup>+</sup>GFP<sup>−</sup> and mRFP<sup>+</sup>GFP<sup>+</sup> punctae in the axons (<b>F</b>) and somata (<b>G</b>) of primary motoneurons treated with tunicamycin for the indicated times or left untreated. Four independent experiments with at least 15 neurons analyzed. One-way ANOVA, Dunnett’s multiple comparisons test. All data in <a href="#cells-13-01498-f001" class="html-fig">Figure 1</a> are shown as the mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Absence of calnexin causes enhanced ER-phagy and protein degradation by default. (<b>A</b>) Calnexin depletion results in an increase in Lamp1<sup>+</sup>mRFP<sup>+</sup>GFP<sup>+</sup> vesicles in the axons of primary MNs under basal conditions. Scale bar: 5 µm. mRFP-GFP-KDEL expressing MNs were cultured for 5 days and treated with tunicamycin for 1 h or left untreated. TM, Tunicamycin. (<b>B</b>) Quantification of mRFP<sup>+</sup>GFP<sup>−</sup> and mRFP<sup>+</sup>GFP<sup>+</sup> punctae in axons of primary MNs treated with tunicamycin for 1 h or left untreated. At least 15 neurons were analyzed per experiment. Two-way ANOVA, Tukey’s multiple comparisons test. (<b>C</b>) In the somata of cultured MNs, the loss of calnexin resulted in an increase in Lamp1<sup>+</sup>mRFP<sup>+</sup>GFP<sup>+</sup> and Lamp1<sup>+</sup>mRFP<sup>+</sup>GFP<sup>−</sup> vesicles under basal conditions. Arrowheads point to mRFP<sup>+</sup>GFP<sup>+</sup> punctae. mRFP-GFP-KDEL expressing MNs were cultured for 5 days and treated with tunicamycin for 1 h or left untreated. TM, Tunicamycin. Scale bar: 2.5 µm. (<b>D</b>) Quantification of mRFP<sup>+</sup>GFP<sup>−</sup> and mRFP<sup>+</sup>GFP<sup>+</sup> punctae in the somata of primary MNs treated with tunicamycin for 1 h or left untreated. At least 15 neurons were analyzed per experiment. Two-way ANOVA, Tukey’s multiple comparisons test. (<b>E</b>,<b>F</b>) Western blot showing the insoluble (<b>E</b>) and soluble (<b>F</b>) protein levels of LC3, p62, and polyubiquitinated proteins. Wildtype and calnexin-deficient early cortical neurons were treated as indicated and subjected to the fractionation procedure. (<b>G</b>–<b>L</b>) Quantification of the LC3-II levels in the TX100-soluble (<b>G</b>) and SDS-soluble (<b>H</b>) fraction. Quantification of the p62 levels in the TX100-soluble (<b>I</b>) and SDS-soluble (<b>J</b>) fractions. Quantification of the polyubiquitinated protein levels in the TX100-soluble (<b>K</b>) and SDS-soluble (<b>L</b>) fractions. The protein levels were normalized to Gapdh in the soluble fraction. Three-independent experiments. <span class="html-italic">n</span> = 3. All data in <a href="#cells-13-01498-f002" class="html-fig">Figure 2</a> are shown as the mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Calnexin depletion leads to a diminished activation of the PERK branch of the UPR under ER stress conditions. (<b>A</b>,<b>B</b>) Expression levels of spliced (<b>A</b>) and total <span class="html-italic">Xbp1</span> (<b>B</b>) in wildtype and calnexin-deficient cortical neurons analyzed by qRT-PCR. Basal expression levels were normalized to Gapdh, and <span class="html-italic">canx<sup>+/+</sup></span> was set to 1 for each experiment. <span class="html-italic">n</span> = 3, One-sample <span class="html-italic">t</span>-test. ER stress-induced expression levels were normalized to Gapdh, and the untreated control condition was set to 1 in each experiment. Four independent experiments, <span class="html-italic">n</span> = 4. Two-way ANOVA, Šídák’s multiple comparisons test. (<b>C</b>) Reduced clustering of IRE1-3F6HGFP in calnexin-deficient early cortical neurons after treatment with tunicamycin. IRE1-3F6HGFP was lentivirally expressed in early cortical neurons and the cells were exposed to tunicamycin for the indicated times. Representative images are shown. Scale bar: 20 µm. (<b>D</b>) Quantification of IRE1-3F6HGFP clustering. <span class="html-italic">n</span> = 1, each data point represents one quantified cell. (<b>E</b>) Western blot showing an impaired activation of the PERK UPR branch in calnexin-deficient early cortical neurons upon thapsigargin treatment. The cells were exposed to thapsigargin for the indicated times, lysed, and processed for Western blotting. Western blots were probed with the indicated antibodies. (<b>F</b>–<b>H</b>) Western blot quantifications revealed a reduced phosphorylation of PERK (<b>F</b>) and eIF2α (<b>G</b>) and a reduced expression of Chop (<b>H</b>) in <span class="html-italic">canx<sup>−/−</sup></span> cells upon ER stress induction by thapsigargin. Three independent experiments, <span class="html-italic">n</span> = 3. Two-way ANOVA, Šídák’s multiple comparisons test. (<b>I</b>) Deficiency in calnexin did not cause alterations in the translocation of ATF6 to the Golgi and nucleus. ATF6-EGFP was lentivirally expressed in early cortical neurons and GM130 was visualized by immunocytochemical staining. The cells were exposed to DTT for the indicated times. Scale bar: 20 µm. (<b>J</b>,<b>K</b>) Quantification of the EGFP fluorescent intensity at the Golgi (<b>J</b>) and in the nucleus (<b>K</b>). The intensity of the GFP signal overlapping with GM130 (Golgi) and DAPI (nucleus) was analyzed. Three independent experiments, <span class="html-italic">n</span> = 3. Two-way ANOVA, Šídák’s multiple comparisons test. All data in <a href="#cells-13-01498-f003" class="html-fig">Figure 3</a> are shown as the mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
17 pages, 4362 KiB  
Article
Development of Dual-Targeted Mixed Micelles Loaded with Celastrol and Evaluation on Triple-Negative Breast Cancer Therapy
by Siying Huang, Simeng Xiao, Xuehao Li, Ranran Tao, Zhangwei Yang, Ziwei Gao, Junjie Hu, Yan Meng, Guohua Zheng and Xinyan Chen
Pharmaceutics 2024, 16(9), 1174; https://doi.org/10.3390/pharmaceutics16091174 - 6 Sep 2024
Viewed by 321
Abstract
Considering that the precise delivery of Celastrol (Cst) into mitochondria to induce mitochondrial dysfunction may be a potential approach to improve the therapeutic outcomes of Cst on TNBC, a novel tumor mitochondria dual-targeted mixed-micelle nano-system was fabricated via self-synthesized triphenylphosphonium-modified cholesterol (TPP-Chol) and [...] Read more.
Considering that the precise delivery of Celastrol (Cst) into mitochondria to induce mitochondrial dysfunction may be a potential approach to improve the therapeutic outcomes of Cst on TNBC, a novel tumor mitochondria dual-targeted mixed-micelle nano-system was fabricated via self-synthesized triphenylphosphonium-modified cholesterol (TPP-Chol) and hyaluronic acid (HA)-modified cholesterol (HA-Chol). The Cst-loaded mixed micelles (Cst@HA/TPP-M) exhibited the characteristics of a small particle size, negative surface potential, high drug loading of up to 22.8%, and sustained drug release behavior. Compared to Cst-loaded micelles assembled only by TPP-Chol (Cst@TPP-M), Cst@HA/TPP-M decreased the hemolysis rate and upgraded the in vivo stability and safety. In addition, a series of cell experiments using the triple-negative breast cancer cell line MDA-MB-231 as a cell model proved that Cst@HA/TPP-M effectively increased the cellular uptake of the drug through CD44-receptors-mediated endocytosis, and the uptake amount was three times that of the free Cst group. The confocal results demonstrated successful endo-lysosomal escape and effective mitochondrial transport triggered by the charge converse of Cst@HA/TPP-M after HA degradation in endo-lysosomes. Compared to the free Cst group, Cst@HA/TPP-M significantly elevated the ROS levels, reduced the mitochondrial membrane potential, and promoted tumor cell apoptosis, showing a better induction effect on mitochondrial dysfunction. In vivo imaging and antitumor experiments based on MDA-MB-231-tumor-bearing nude mice showed that Cst@HA/TPP-M facilitated drug enrichment at the tumor site, attenuated drug systemic distribution, and polished up the antitumor efficacy of Cst compared with free Cst. In general, as a target drug delivery system, mixed micelles co-constructed by TPP-Chol and HA-Chol might provide a promising strategy to ameliorate the therapeutic outcomes of Cst on TNBC. Full article
(This article belongs to the Section Drug Delivery and Controlled Release)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Appearance and Tyndall effect of Cst@HA/TPP-M. (<b>B</b>) TEM image of Cst@HA/TPP-M. (<b>C</b>,<b>D</b>) Changes in the particle size and encapsulation efficiency of Cst@TPP-M and Cst@HA/TPP-M stored at 4 °C for 20 days. (<b>E</b>,<b>F</b>) Changes in the particle size and Zeta potential of Cst@TPP-M and Cst@HA/TPP-M after incubation with 10% FBS.</p>
Full article ">Figure 2
<p>(<b>A</b>) Change in Zeta potential of Cst@HA/TPP-M after incubation with and without HAase (1 mg/mL) at different pHs (pH 7.4 and 5.0) over time. (<b>B</b>) In vitro cumulative release rate of Cst. (<b>C</b>) Hemolysis rate of Cst@TPP-M and Cst@HA/TPP-M at different Cst concentrations, ** <span class="html-italic">p</span> &lt; 0.01, vs. Cst@TPP-M. (<b>D</b>) Cell viability of LO2 cells after treatment with TPP-M and HA/TPP-M for 24 h.</p>
Full article ">Figure 3
<p>(<b>A</b>) Fluorescent images of MDA-MB-231 cells incubated with free C6 and C6@HA/TPP-M after 1 h of HA pretreatment and non-pretreatment. Scale bar: 50 μm. (<b>B</b>) Quantification of fluorescent intensity in MDA-MB-231 cells incubated with free C6 and C6@HA/TPP-M after 1 h of HA pretreatment and non-pretreatment. *** <span class="html-italic">p</span> &lt; 0.001, vs. free C6. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3). (<b>C</b>) Flow cytometry analysis of MDA-MB-231 cells incubated with free C6 and C6@HA/TPP-M after 1 h of HA pretreatment and non-pretreatment.</p>
Full article ">Figure 4
<p>(<b>A</b>) Endo-lysosome escape observation of C6@HA/TPP-M. Scale bar: 50 μm. (<b>B</b>) CLSM images of MDA-MB-231 cells after incubation with free C6, C6@HA/TPP-M for 4 h. The mitochondria were stained with MitoTracker Red (Biyuntian, Shanghai, China). Scale bar: 100 μm.</p>
Full article ">Figure 5
<p>(<b>A</b>) Fluorescent images of intracellular ROS in MDA-MB-231cells treated with free Cst and Cst@HA/TPP-M. Scale bar: 50 μm. (<b>B</b>) Quantification of the ROS level, *** <span class="html-italic">p</span> &lt; 0.001, vs. free Cst. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3). (<b>C</b>) Flow cytometry analysis of MDA-MB-231 cells treated with free Cst and Cst@HA/TPP-M. (<b>D</b>) The intensity ratios of red fluorescence to green fluorescence in MDA-MB-231 cells after treatment with free Cst and Cst@HA/TPP-M, *** <span class="html-italic">p</span> &lt; 0.001, vs. free Cst.</p>
Full article ">Figure 6
<p>(<b>A</b>) Representative flow cytometric graphs of MDA-MB-231cells after 24 h of incubation with different Cst formulations. (<b>B</b>) Apoptosis rate analysis of MDA-MB-231 cells after 24 h of incubation with different Cst formulations, *** <span class="html-italic">p</span> &lt; 0.001, vs. control. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3). (<b>C</b>) Viability of MDA-MB-231 cells after incubation with different Cst formulations for 24 h.</p>
Full article ">Figure 7
<p>(<b>A</b>) In vivo real-time imaging of MDA-MB-231-tumor-bearing nude mice after intravenous injection with free NR, NR@TPP-M, and NR@HA/TPP-M. (<b>B</b>) The total radiant efficiency in the tumors of MDA-MB-231-tumor-bearing nude mice after intravenous injection with free NR, NR@TPP-M, and NR@HA/TPP-M. (<b>C</b>) Ex vivo optical images of the major organs (heart, liver, spleen, lung, kidney, and tumor) collected at 24 h post-injection. (<b>D</b>) The total radiant efficiency in the tumors which were taken after the mice were sacrificed at 24 h after injection. *** <span class="html-italic">p</span> &lt; 0.001, vs. NR. All data are represented as mean ± SD.</p>
Full article ">Figure 8
<p>(<b>A</b>) Change in the body weight of MDA-MB-231-tumor-bearing nude mice after intravenous injection with saline, Cst, and Cst@HA/TPP-M, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (<b>B</b>) Change in the tumor volume of MDA-MB-231-tumor-bearing nude mice after intravenous injection with saline, Cst, and Cst@HA/TPP-M, ** <span class="html-italic">p</span> &lt; 0.01. (<b>C</b>) Tumor index analysis of tumor weight relative to body weight, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. All data are represented as mean ± SD. (<b>D</b>) Tumor inhibition rate analysis, *** <span class="html-italic">p</span> &lt; 0.001, vs. Cst. (<b>E</b>) Macroscopic appearance of tumors collected from MDA-MB-231-tumor-bearing nude mice after treatment.</p>
Full article ">Scheme 1
<p>Schematic illustration of the preparation, accumulation at the breast tumor site, CD44-receptor-mediated endocytosis, HA degradation, endo-lysosomal escape, and mitochondrial targeting of Cst@HA/TPP-M.</p>
Full article ">
20 pages, 3319 KiB  
Article
Combinatory Effect of Pequi Oil (Caryocar brasiliense)-Based Nanoemulsions Associated to Docetaxel and Anacardic Acid (Anacardium occidentale) in Triple-Negative Breast Cancer Cells In Vitro
by Alicia Simalie Ombredane, Natália Ornelas Martins, Gabriela Mara Vieira de Souza, Victor Hugo Sousa Araujo, Ísis O. Szlachetka, Sebastião William da Silva, Márcia Cristina Oliveira da Rocha, Andressa Souza de Oliveira, Cleonice Andrade Holanda, Luiz Antonio Soares Romeiro, Elysa Beatriz de Oliveira Damas, Ricardo Bentes Azevedo and Graziella Anselmo Joanitti
Pharmaceutics 2024, 16(9), 1170; https://doi.org/10.3390/pharmaceutics16091170 - 5 Sep 2024
Viewed by 401
Abstract
Combination therapy integrated with nanotechnology offers a promising alternative for breast cancer treatment. The inclusion of pequi oil, anacardic acid (AA), and docetaxel (DTX) in a nanoemulsion can amplify the antitumor effects of each molecule while reducing adverse effects. Therefore, the study aims [...] Read more.
Combination therapy integrated with nanotechnology offers a promising alternative for breast cancer treatment. The inclusion of pequi oil, anacardic acid (AA), and docetaxel (DTX) in a nanoemulsion can amplify the antitumor effects of each molecule while reducing adverse effects. Therefore, the study aims to develop pequi oil-based nanoemulsions (PeNE) containing DTX (PDTX) or AA (PAA) and to evaluate their cytotoxicity against triple-negative breast cancer cells (4T1) in vitro. The PeNE without and with AA (PAA) and DTX (PDTX) were prepared by sonication and characterized by ZetaSizer® and electronic transmission microscopy. Viability testing and combination index (CI) were determined by MTT and Chou-Talalay methods, respectively. Flow cytometry was employed to investigate the effects of the formulations on cell structures. PeNE, PDTX, and PAA showed hydrodynamic diameter < 200 nm and a polydispersity index (PdI) of 0.3. The association PDTX + PAA induced a greater decrease in cell viability (~70%, p < 0.0001) and additive effect (CI < 1). In parallel, an association of the DTX + AA molecules led to antagonism (CI > 1). Additionally, PDTX + PAA induced an expressive morphological change, a major change in lysosome membrane permeation and mitochondria membrane permeation, cell cycle blockage in G2/M, and phosphatidylserine exposure. The study highlights the successful use of pequi oil nanoemulsions as delivery systems for DTX and AA, which enhances their antitumor effects against breast cancer cells. This nanotechnological approach shows significant potential for the treatment of triple-negative breast cancer. Full article
Show Figures

Figure 1

Figure 1
<p>Shape of pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (PAA), and docetaxel (PDTX), evaluated by transmission electron microscopy. Scale bar = 200 nm.</p>
Full article ">Figure 2
<p>Stability of pequi oil of pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated to anacardic acid (PAA), and docetaxel (PDTX) at 4 °C storage according to hydrodynamic diameter (<b>A</b>), polydispersity index (<b>B</b>), and zeta potential (<b>C</b>). One-way ANOVA <span class="html-italic">p</span> &lt; 0.05 (Tukey post hoc test).</p>
Full article ">Figure 3
<p>FTIR spectrum of pequi oil nanoemulsion (PeNE) (i), pequi oil nanoemulsion associated to anacardic acid (PAA) (ii), pequi oil nanoemulsion associated to docetaxel (PDTX) (iii), blank formulation (only lecithin) (iv), pure pequi oil (v), free anacardic acid (vi), and free docetaxel (vii).</p>
Full article ">Figure 4
<p>(<b>a</b>) FTIR spectrum in the region of 2800 to 3050 cm<sup>−1</sup> of the pequi oil nanoemulsion (PENE—black line), pequi oil nanoemulsion associated to anacardic acid (PAA—blue line), and pequi oil nanoemulsion associated to docetaxel (PDTX—red line). Magnification of spectral regions around 2852 cm<sup>−1</sup> (<b>b</b>) and 2922 cm<sup>−1</sup> (<b>c</b>). Dependence on vibrational energies (<b>d</b>) and the ratios between the intensities of the absorption bands <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> <mo>(</mo> <msub> <mrow> <mi>C</mi> <mi>H</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>a</mi> <mi>s</mi> </mrow> </msub> <mo>(</mo> <msub> <mrow> <mi>C</mi> <mi>H</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> as a function of the different nanoemulsions (<b>e</b>).</p>
Full article ">Figure 5
<p>Morphologic evaluation of triple-negative breast cancer cells (4T1) by contrast phase microscopy (<b>A</b>) and flow cytometry (<b>B</b>) after 24 h of exposure with pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (PAA), and docetaxel (PDTX), free anarcadic acid (AA), free docetaxel (DTX), association of PAA and PDTX (P.AA + P.DTX), and association of free AA and free DTX (AA + DTX) at 180 µg/mL of pequi oil, 10 µg/mL of AA, and 16 µg/mL of DTX. Control group was treated with phosphate buffer. (<b>A</b>) Scale bar = 100 µm. (<b>B</b>) FSC-H = size and SSC-H = granularity. One-way ANOVA: significant difference between groups <span class="html-italic">p</span> &lt; 0.05 (Tukey post hoc test). Different letters indicate statistically significant differences between groups.</p>
Full article ">Figure 5 Cont.
<p>Morphologic evaluation of triple-negative breast cancer cells (4T1) by contrast phase microscopy (<b>A</b>) and flow cytometry (<b>B</b>) after 24 h of exposure with pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (PAA), and docetaxel (PDTX), free anarcadic acid (AA), free docetaxel (DTX), association of PAA and PDTX (P.AA + P.DTX), and association of free AA and free DTX (AA + DTX) at 180 µg/mL of pequi oil, 10 µg/mL of AA, and 16 µg/mL of DTX. Control group was treated with phosphate buffer. (<b>A</b>) Scale bar = 100 µm. (<b>B</b>) FSC-H = size and SSC-H = granularity. One-way ANOVA: significant difference between groups <span class="html-italic">p</span> &lt; 0.05 (Tukey post hoc test). Different letters indicate statistically significant differences between groups.</p>
Full article ">Figure 6
<p>Assessment of cytotoxic effect on key organelles, plasma membrane, and intracellular physiology of triple-negative breast cancer cells (4T1) by flow cytometry. The cells were exposed to pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (P.AA), and docetaxel (P.DTX), free anarcadic acid (AA), free docetaxel (DTX), association of PAA e PDTX, and association of free AA and free DTX (AA + DTX) at 180 µg/mL of pequi oil, 10 µg/mL of AA and 16 µg/mL of DTX for 24 h. (<b>A</b>) Lysosomal membrane permeability. (<b>B</b>) Mitochondrial membrane potential. (<b>C</b>) Fragmentation of DNA. (<b>D</b>) Membrane integrity by trypan blue assay. (<b>E</b>) Intracellular Reactive Oxygen Species (ROS) analysis with CM-H<sub>2</sub>DCFDA. One-way and two-way ANOVA: significant difference between groups with *: <span class="html-italic">p</span> &lt; 0.05, ***: <span class="html-italic">p</span> &lt; 0.001 and ****: <span class="html-italic">p</span> &lt; 0.0001 (Tukey post hoc test). Different letters indicate statistically significant differences between groups.</p>
Full article ">Figure 6 Cont.
<p>Assessment of cytotoxic effect on key organelles, plasma membrane, and intracellular physiology of triple-negative breast cancer cells (4T1) by flow cytometry. The cells were exposed to pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (P.AA), and docetaxel (P.DTX), free anarcadic acid (AA), free docetaxel (DTX), association of PAA e PDTX, and association of free AA and free DTX (AA + DTX) at 180 µg/mL of pequi oil, 10 µg/mL of AA and 16 µg/mL of DTX for 24 h. (<b>A</b>) Lysosomal membrane permeability. (<b>B</b>) Mitochondrial membrane potential. (<b>C</b>) Fragmentation of DNA. (<b>D</b>) Membrane integrity by trypan blue assay. (<b>E</b>) Intracellular Reactive Oxygen Species (ROS) analysis with CM-H<sub>2</sub>DCFDA. One-way and two-way ANOVA: significant difference between groups with *: <span class="html-italic">p</span> &lt; 0.05, ***: <span class="html-italic">p</span> &lt; 0.001 and ****: <span class="html-italic">p</span> &lt; 0.0001 (Tukey post hoc test). Different letters indicate statistically significant differences between groups.</p>
Full article ">Figure 7
<p>Impact on cell proliferation. The cells were exposed to pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (PAA) and docetaxel (PDTX), free anacardic acid (AA), free docetaxel (DTX), association of PAA + P. DTX, and association of free AA and free DTX (AA + DTX) at 180 µg/mL of pequi oil, 10 µg/mL of AA and 16 µg/mL of DTX for 24 h. (<b>A</b>) Total cell number by trypan blue assay. (<b>B</b>) Cell cycle using propidium iodide by flow cytometry. (<b>C</b>) Clonogenic assay for formation of colonies. (<b>A</b>,<b>C</b>) One-way ANOVA: significant difference between groups <span class="html-italic">p</span> &lt; 0.05 (Tukey post hoc test). Different letters indicate statistically significant differences between groups. (<b>B</b>) Two-way ANOVA: significant difference between groups **** <span class="html-italic">p</span> &lt; 0.001 (Tukey post hoc test).</p>
Full article ">Figure 8
<p>Exposition of phosphatidylserine and caspase activation. The cells were exposed to pequi oil nanoemulsion (PeNE), pequi oil nanoemulsion associated with anacardic acid (PAA) and docetaxel (PDTX), free anarcadic acid (AA), free docetaxel (DTX), association of PAA + PDTX and association of free AA and free DTX (AA + DTX) at 180 µg/mL of pequi oil, 10 µg/mL of AA, and 16 µg/mL of DTX for 24 h. (<b>A</b>) Exposition of phosphatidylserine (stained with annexin V-FITC and propidium iodide by flow cytometry). (<b>B</b>) Multicaspase activity. One-way ANOVA: significant difference between groups <span class="html-italic">p</span> &lt; 0.05 (Tukey post hoc test). Different letters indicate statistically significant differences between groups.</p>
Full article ">Figure 9
<p>Suggested cell targets involved in the cytotoxicity of pequi oil nanoemulsion associated to docetaxel (PDTX) and anacardic acid (PAA) on breast cancer cells (4T1). Source: own authorship created in BioRender.com.</p>
Full article ">
25 pages, 788 KiB  
Review
Subcellular Drug Distribution: Exploring Organelle-Specific Characteristics for Enhanced Therapeutic Efficacy
by Xin Liu, Miaomiao Li and Sukyung Woo
Pharmaceutics 2024, 16(9), 1167; https://doi.org/10.3390/pharmaceutics16091167 - 4 Sep 2024
Viewed by 687
Abstract
The efficacy and potential toxicity of drug treatments depends on the drug concentration at its site of action, intricately linked to its distribution within diverse organelles of mammalian cells. These organelles, including the nucleus, endosome, lysosome, mitochondria, endoplasmic reticulum, Golgi apparatus, lipid droplets, [...] Read more.
The efficacy and potential toxicity of drug treatments depends on the drug concentration at its site of action, intricately linked to its distribution within diverse organelles of mammalian cells. These organelles, including the nucleus, endosome, lysosome, mitochondria, endoplasmic reticulum, Golgi apparatus, lipid droplets, exosomes, and membrane-less structures, create distinct sub-compartments within the cell, each with unique biological features. Certain structures within these sub-compartments possess the ability to selectively accumulate or exclude drugs based on their physicochemical attributes, directly impacting drug efficacy. Under pathological conditions, such as cancer, many cells undergo dynamic alterations in subcellular organelles, leading to changes in the active concentration of drugs. A mechanistic and quantitative understanding of how organelle characteristics and abundance alter drug partition coefficients is crucial. This review explores biological factors and physicochemical properties influencing subcellular drug distribution, alongside strategies for modulation to enhance efficacy. Additionally, we discuss physiologically based computational models for subcellular drug distribution, providing a quantifiable means to simulate and predict drug distribution at the subcellular level, with the potential to optimize drug development strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of a mammalian cell highlighting the differential accumulation of drugs within various subcellular organelles and structures. Italics indicate the factors affecting accumulation.</p>
Full article ">
25 pages, 4438 KiB  
Article
Methamphetamine Increases Tubulo-Vesicular Areas While Dissipating Proteins from Vesicles Involved in Cell Clearance
by Gloria Lazzeri, Paola Lenzi, Carla L. Busceti, Stefano Puglisi-Allegra, Michela Ferrucci and Francesco Fornai
Int. J. Mol. Sci. 2024, 25(17), 9601; https://doi.org/10.3390/ijms25179601 - 4 Sep 2024
Viewed by 397
Abstract
Cytopathology induced by methamphetamine (METH) is reminiscent of degenerative disorders such as Parkinson’s disease, and it is characterized by membrane organelles arranged in tubulo-vesicular structures. These areas, appearing as clusters of vesicles, have never been defined concerning the presence of specific organelles. Therefore, [...] Read more.
Cytopathology induced by methamphetamine (METH) is reminiscent of degenerative disorders such as Parkinson’s disease, and it is characterized by membrane organelles arranged in tubulo-vesicular structures. These areas, appearing as clusters of vesicles, have never been defined concerning the presence of specific organelles. Therefore, the present study aimed to identify the relative and absolute area of specific membrane-bound organelles following a moderate dose (100 µM) of METH administered to catecholamine-containing PC12 cells. Organelles and antigens were detected by immunofluorescence, and they were further quantified by plain electron microscopy and in situ stoichiometry. This analysis indicated an increase in autophagosomes and damaged mitochondria along with a decrease in lysosomes and healthy mitochondria. Following METH, a severe dissipation of hallmark proteins from their own vesicles was measured. In fact, the amounts of LC3 and p62 were reduced within autophagy vacuoles compared with the whole cytosol. Similarly, LAMP1 and Cathepsin-D within lysosomes were reduced. These findings suggest a loss of compartmentalization and confirm a decrease in the competence of cell clearing organelles during catecholamine degeneration. Such cell entropy is consistent with a loss of energy stores, which routinely govern appropriate subcellular compartmentalization. Full article
Show Figures

Figure 1

Figure 1
<p>In PC12 cells, a moderate dose of METH (100 µM) induces moderate cell death and noticeable cell damage. (<b>A</b>) Representative pictures of H&amp;E-stained cells from controls and METH-treated cells, respectively. Apart from decreased cell number (63% of controls), there is evident cell pathology induced by METH, with pale vacuolar cytosolic areas, irregular shape and smaller size compared with the homogenously stained cytosol and regular cell shape of the control. (<b>B</b>) FJ-B histofluorescence from controls and METH-treated cells, respectively, where degenerating cells are evident mostly among METH-treated cells. (<b>C</b>) The graph reports the percentage of H&amp;E-stained viable cells following METH compared with control conditions according to the criteria expressed in the Methods, extended statistics. (<b>D</b>) The graph reports that FJ-B-positive cells (absolute number), occur in excess (almost twelvefold) in METH-treated cells compared with controls. Values in C are given as the mean percentage of controls (where controls = 100%) ± S.E.M. Values in D are given as the mean ± S.E.M. of controls and METH-treated cells, respectively. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bars (<b>A</b>,<b>B</b>) = 15 µm.</p>
Full article ">Figure 2
<p>Specific vesicular markers and their overlap with LC3 increase following METH (100 µM). As shown in representative pictures in <a href="#ijms-25-09601-f002" class="html-fig">Figure 2</a>, in PC12 cells, METH markedly increases the amount of LC3 compared with control conditions both alone and in combination with various antigens, which are also modified: (<b>A</b>) p62; (<b>B</b>) P20S; (<b>C</b>) LAMP-1; (<b>D</b>) Cat-D and (<b>E</b>) Mitogreen. All these antigens increase following METH (except p20S), and an increase is measured in their area of overlap with LC3 (including P20S). It is important to emphasize that Mitogreen is a non-specific mitochondrial marker, since it stains both healthy and damaged mitochondria. It is difficult to interpret the significance of this latter merging based solely on light microscopy. In fact, this could be due to either the placement of LC3 over the mitochondrial structure or the mitophagy of altered mitochondria within engulfed autophagosomes. Again, this might be due to disrupted mitochondrial structures co-localizing with LC3 outside of any specific compartment. (<b>F</b>) reports the graph showing the area of overlapping immunofluorescence between each antigen and LC3. As reported, all antigens produce a remarkable area of overlap with LC3 following METH (100 µM) compared with control conditions. Values are given as the mean percentage compared with controls, assuming controls = 100% ± S.E.M. Please refer also to <a href="#ijms-25-09601-t001" class="html-table">Table 1</a>, which reports the absolute overlapping area, expressed in µm<sup>2</sup>). Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bars = 12 µm.</p>
Full article ">Figure 3
<p>METH alters tubulo-vesicular domains in PC12 cells on TEM. (<b>A</b>) Representative pictures from control (upper lane)- and METH (lower lane)-treated PC12 cells. In each of these pictures, the inset provides a higher magnification. (<b>B</b>) The graph reports the area covered by tubulo-vesicular structures, which is expressed as the mean percentage ± S.E.M. of the area of the whole cytosol (n = 30). Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Picture scale bars = 689 nm. Inset scale bars = 229 nm (Control); 300 nm (METH).</p>
Full article ">Figure 4
<p>In PC12 cells, METH increases LC3 immunogold in the whole cytosol and dissipates LC3 from vacuoles. (<b>A</b>) Representative pictures showing Control and METH, respectively. Arrows indicate LC3 immunogold. METH markedly increases LC3 in the whole cytosol, although the vacuolar amount of LC3 does not change. (<b>B</b>) A graph reporting the number of LC3 particles in the whole cytosol. (<b>C</b>) A graph reporting the amount of LC3 within vacuoles. (<b>D</b>) A graph reporting the decrease in the ratio of vacuolar to cytosolic LC3 produced by METH. This indicates that, although METH increases LC3 in the whole cell, this autophagosome-preferring protein is dissipated from autophagosome vacuoles towards the non-particulate cytosol. Values are expressed as the mean + S.E.M. of n = 30 cells per group. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bar (<b>A</b>): 125 nm.</p>
Full article ">Figure 5
<p>In PC12 cells, METH increases p62 immunogold in the whole cytosol, while decreases p62 from vacuoles. (<b>A</b>) Representative pictures showing Control and METH, respectively. Arrows indicate p62 immunogold. The amount markedly increases in the whole cytosol although vacuolar amount of p62 was reduced following METH exposure. (<b>B</b>) A graph reporting the number of p62-stained immunogold particles in the whole cytosol. (<b>C</b>) A graph reporting the amount of p62 within vacuoles. (<b>D</b>) A graph reporting the decrease in the ratio of vacuolar to cytosolic p62 produced by METH. This indicates that, although METH increases p62 in the whole cell, this autophagosome-preferring protein is dissipated from autophagosome vacuoles towards the non-particulate peri-vacuolar cytosol. Values are expressed as the mean + S.E.M. of n = 30 cells per group. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bar (<b>A</b>): 172 nm.</p>
Full article ">Figure 6
<p>In PC12 cells, METH does not alter cytosolic P20S despite reducing vacuolar P20S. (<b>A</b>) Representative pictures showing Control and METH, respectively. Arrows indicate P20S immunogold. The amount varies following METH when counted in the whole cytosol, although vacuolar amount of p20S is following METH exposure. (<b>B</b>) A graph reporting the number of P20S-stained immunogold particles in the whole cytosol. (<b>C</b>) A graph reporting the amount of P20S within vacuoles. (<b>D</b>) A graph reporting the decrease in the ratio of vacuolar to cytosolic P20S produced by METH. This indicates that, despite METH does not alter the amount of P20S in the whole cytosol, this proteasome-preferring protein is dissipated from autophagoproteasome vacuoles [<a href="#B16-ijms-25-09601" class="html-bibr">16</a>] towards the non-particulate peri-vacuolar cytosol. Values are expressed as the mean + S.E.M. of n = 30 cells per group. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bar (<b>A</b>): 172 nm.</p>
Full article ">Figure 7
<p>METH does not alter cytosolic LAMP1 despite reducing LAMP1 vacuolar compartmentalization in PC12 cells. (<b>A</b>) Representative pictures showing Control and METH, respectively. Arrows indicate LAMP1 immunogold. The quantity does not change following METH when counted in the whole cytosol, although the vacuolar amount of LAMP1 is markedly reduced following METH exposure. (<b>B</b>) A graph reporting the number of LAMP1-stained immunogold particles in the whole cytosol. (<b>C</b>) A graph reporting the amount of LAMP1 within vacuoles. (<b>D</b>) A graph reporting the decrease in the ratio of vacuolar to cytosolic LAMP1 produced by METH. This indicates that, although METH does not alter the amount of LAMP1 in the whole cytosol, this lysosome-preferring protein is dissipated from lysosome/autophagolysosome vacuoles towards the non-particulate peri-vacuolar cytosol. Values are expressed as the mean + S.E.M. of n = 30 cells per group. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bar (<b>A</b>): 125 nm.</p>
Full article ">Figure 8
<p>METH increases cytosolic Cat-D while reducing the amount of Cat-D in the vacuoles of PC12 cells. (<b>A</b>) Representative pictures showing Control and METH, respectively. Arrows indicate Cat-D immunogold. The amount strongly increases following METH when counted in the whole cytosol, although the vacuolar amount of Cat-D is markedly reduced following METH exposure. (<b>B</b>) A graph reporting the number of Cat-D-stained immunogold particles in the whole cytosol. (<b>C</b>) A graph reporting the amount of Cat-D within vacuoles. (<b>D</b>) A graph reporting the dramatic decrease in the ratio of vacuolar to cytosolic Cat-D produced by METH. This indicates that, METH severely decreases Cat-D compartmentalization both as a consequence of increased cytosolic Cat-D and reduced vacuolar Cat-D. Thus, this lysosome-preferring enzyme is profoundly dissipated from lysosome/autophagolysosome vacuoles towards the non-particulate peri-vacuolar cytosol. Values are expressed as the mean + S.E.M. of n = 30 cells per group. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bar (<b>A</b>): 172 nm.</p>
Full article ">Figure 9
<p>In PC12 cells METH dissipates all double immunogold keeping constant LC3 immunocytochemistry. When double immunogold was carried out, antibodies against LC3 (arrows) were challenged with all other antibodies (arrowheads). The representative pictures (<b>A</b>) and graph (<b>B</b>) report double staining of LC3 and p62, which indicate that METH produces a marked loss of LC3+p62 positive vacuoles. Representative pictures (<b>C</b>) and graph (<b>D</b>) show double immunogold for LC3+P20S, which provides similar results since METH decreases the vacuolar amount of both antigens. The representative pictures (<b>E</b>) and graph (<b>F</b>) report double immunogold of LC3+LAMP1, which indicates a severe loss of autophagolysosomes. This is likely to be due to impaired merging between autophagosomes and lysosomes. The representative pictures (<b>G</b>) and graph (<b>H</b>) report double immunogold of LC3+Cat-D. Again, the decrease is consistent with previous findings indicating a loss of autophagolysosomes. Values are expressed as the mean + S.E.M. of n = 30 cells per group. Comparisons between groups were carried out through one-way ANOVA with Scheffe’s post hoc analysis. The null hypothesis was rejected for <span class="html-italic">p</span> &lt; 0.05. * <span class="html-italic">p</span> &lt; 0.05 compared with controls. Scale bars: 125 nm.</p>
Full article ">
14 pages, 19599 KiB  
Article
Lysosome-Associated Membrane Protein Targeting Strategy Improved Immunogenicity of Glycoprotein-Based DNA Vaccine for Marburg Virus
by Xiyang Zhang, Yubo Sun, Junqi Zhang, Hengzheng Wei, Jing Wang, Chenchen Hu, Yang Liu, Sirui Cai, Qinghong Yuan, Yueyue Wang, Yuanjie Sun, Shuya Yang, Dongbo Jiang and Kun Yang
Vaccines 2024, 12(9), 1013; https://doi.org/10.3390/vaccines12091013 - 4 Sep 2024
Viewed by 424
Abstract
Marburg hemorrhagic fever (MHF) is a fatal infectious disease caused by Marburg virus (MARV) infection, and MARV has been identified as a priority pathogen for vaccine development by the WHO. The glycoprotein (GP) of MARV mediates viral adhesion and invasion of host cells [...] Read more.
Marburg hemorrhagic fever (MHF) is a fatal infectious disease caused by Marburg virus (MARV) infection, and MARV has been identified as a priority pathogen for vaccine development by the WHO. The glycoprotein (GP) of MARV mediates viral adhesion and invasion of host cells and therefore can be used as an effective target for vaccine development. Moreover, DNA vaccines have unique advantages, such as simple construction processes, low production costs, and few adverse reactions, but their immunogenicity may decrease due to the poor absorption rate of plasmids. Lysosome-associated membrane protein 1 (LAMP1) can direct antigens to lysosomes and endosomes and has great potential for improving the immunogenicity of nucleic acid vaccines. Therefore, we constructed a DNA vaccine based on a codon-optimized MARV GP (ID MF939097.1) fused with LAMP1 and explored the effect of a LAMP targeting strategy on improving the immunogenicity of the MARV DNA vaccine. ELISA, ELISpot, and flow cytometry revealed that the introduction of LAMP1 into the MARV DNA candidate vaccine improved the humoral and cellular immune response, enhanced the secretion of cytokines, and established long-term immune protection. Transcriptome analysis revealed that the LAMP targeting strategy significantly enriched antigen processing and presentation-related pathways, especially the MHC class II-related pathway, in the candidate vaccine. Our study broadens the strategic vision for enhanced DNA vaccine design and provides a promising candidate vaccine for MHF prevention. Full article
(This article belongs to the Special Issue Advances in Vaccines against Infectious Diseases)
Show Figures

Figure 1

Figure 1
<p>The construction and verification of the plasmids. (<b>A</b>) The codon-optimized MARV GP gene sequence was used to construct the pVAX1-GP<sub>MARV</sub> plasmids. pVAX1-LAMP/GP<sub>MARV</sub> was constructed by inserting the GP gene into the pVAX1-LAMP vector. (<b>B</b>) The sizes of the recombinant plasmids were verified by agarose gel electrophoresis. (<b>C</b>) The relative mRNA expression of the MARV GP was detected by qPCR. (<b>D</b>) The GP in the transfected cells was verified by Western blotting. (<b>E</b>) Immunofluorescence images of the MARV GP in transfected cells (40×) (* <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>The evaluation of the humoral immune response induced by the candidate vaccines. (<b>A</b>) The schedule of immunization, sample collection, and experiments. (<b>B</b>) The MARV GP-specific antibody titers were detected by ELISA after each immunization. (<b>C</b>) The cross-neutralizing antibody titers for EBOV were detected after each booster immunization (* <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 3
<p>The production of IFN-γ and IL-4 was evaluated by ELISpot after each booster immunization. (<b>A</b>) Representative images of IFN-γ and IL-4 spots after booster immunization. (<b>B</b>) The levels of cytokines induced by the candidate vaccines were detected under stimulation with different peptide pools. (<b>C</b>) The stimulation effect of a single peptide is also shown as a heatmap. (<b>D</b>) Representative images of spots after long-term booster immunization. (<b>E</b>) The secretion of cytokines was further detected in mice that received long-term booster immunization. (<b>F</b>) The stimulatory effect of a single peptide after long-term booster immunization (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>The T-cell response was evaluated in mice that received the candidate vaccines by flow cytometry. (<b>A</b>) The gating graphs for CD4<sup>+</sup> and CD8<sub>+</sub> T cells (using IFN-γ as an example). (<b>B,C</b>) The secretion of IL-2, IL-4, and IFN-γ by CD4<sup>+</sup> and CD8<sub>+</sub> T cells was detected after each booster immunization. (<b>D</b>) Gating graphs for the CD4<sup>+</sup> Tem cells. (<b>E,F</b>) MARV GP-specific CD4<sup>+</sup> and CD8<sup>+</sup> Tem cells were observed after each booster immunization (* <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>Transcriptome analysis of immune response-related pathways. (<b>A</b>) Volcano plots showing the upregulated and downregulated genes in different groups under stimulation with different peptide pools. (<b>B</b>) KEGG enrichment analysis revealed that both of the candidate vaccines activated immune response-related pathways, especially the B-cell receptor signaling pathway and NK cell-mediated cytotoxicity pathways. (<b>C</b>) Both of the candidate vaccines affected the overall immune response and inflammation-related pathways (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Preliminary safety assessment of candidate vaccines. (<b>A</b>) H&amp;E staining was performed on the main organs of mice after long-term booster immunization. (<b>B</b>) Representative heatmap of animal behavior recorded by an automatic analysis system within 24 h. (<b>C</b>) The activity time and distance, eating and drinking, grooming, and climbing data were collected and analyzed.</p>
Full article ">
15 pages, 729 KiB  
Review
Mucopolysaccharidosis-Plus Syndrome: Is This a Type of Mucopolysaccharidosis or a Separate Kind of Metabolic Disease?
by Zuzanna Cyske, Lidia Gaffke, Karolina Pierzynowska and Grzegorz Węgrzyn
Int. J. Mol. Sci. 2024, 25(17), 9570; https://doi.org/10.3390/ijms25179570 - 4 Sep 2024
Viewed by 679
Abstract
Several years ago, dozens of cases were described in patients with symptoms very similar to mucopolysaccharidosis (MPS). This new disease entity was described as mucopolysaccharidosis-plus syndrome (MPSPS). The name of the disease indicates that in addition to the typical symptoms of conventional MPS, [...] Read more.
Several years ago, dozens of cases were described in patients with symptoms very similar to mucopolysaccharidosis (MPS). This new disease entity was described as mucopolysaccharidosis-plus syndrome (MPSPS). The name of the disease indicates that in addition to the typical symptoms of conventional MPS, patients develop other features such as congenital heart defects and kidney and hematopoietic system disorders. The symptoms are highly advanced, and patients usually do not survive past the second year of life. MPSPS is inherited in an autosomal recessive manner and is caused by a homozygous-specific mutation in the gene encoding the VPS33A protein. To date, it has been described in 41 patients. Patients with MPSPS exhibited excessive excretion of glycosaminoglycans (GAGs) in the urine and exceptionally high levels of heparan sulfate in the plasma, but the accumulation of substrates is not caused by a decrease in the activity of any lysosomal enzymes. Here, we discuss the pathomechanisms and symptoms of MPSPS, comparing them to those of MPS. Moreover, we asked the question whether MPSPS should be classified as a type of MPS or a separate disease, as contrary to ‘classical’ MPS types, despite GAG accumulation, no defects in lysosomal enzymes responsible for degradation of these compounds could be detected in MPSPS. The molecular mechanism of the appearance of GAG accumulation in MPSPS is suggested on the basis of results available in the literature. Full article
Show Figures

Figure 1

Figure 1
<p>The hypothesis of the mechanisms of GAG accumulation in MPSPS. In classical MPS types (upper panel), GAG(s) accumulate(s) due to mutation(s) in one of genes coding for enzymes involved in degradation of this/these compound(s). GAG storage causes various secondary effects, among others, dysregulation of expression of many genes, resulting in pathological changes in different cellular processes. Disturbed vesicular trafficking is among them, leading to impaired transportation of GAGs into lysosomes. This makes GAG accumulation even more pronounced (as GAGs cannot be effectively degraded outside of lysosomes), which drives a spiral of above-described reactions within a positive feedback loop, enhancing the pathological processes. In MPSPS (lower panel), a similar spiral of reactions resulting in GAG accumulation occurs. However, it is initiated by impaired vesicular transport of GAGs (due to deficiency in VPS33A activity, which is required in this process) rather than dysfunction of an enzyme involved in GAG degradation. Nevertheless, the final effect (GAG accumulation) is similar in both classical MPS and MPSPS.</p>
Full article ">
23 pages, 10754 KiB  
Review
Mitochondrial Dysfunction in Glycogen Storage Disorders (GSDs)
by Kumudesh Mishra and Or Kakhlon
Biomolecules 2024, 14(9), 1096; https://doi.org/10.3390/biom14091096 - 1 Sep 2024
Viewed by 443
Abstract
Glycogen storage disorders (GSDs) are a group of inherited metabolic disorders characterized by defects in enzymes involved in glycogen metabolism. Deficiencies in enzymes responsible for glycogen breakdown and synthesis can impair mitochondrial function. For instance, in GSD type II (Pompe disease), acid alpha-glucosidase [...] Read more.
Glycogen storage disorders (GSDs) are a group of inherited metabolic disorders characterized by defects in enzymes involved in glycogen metabolism. Deficiencies in enzymes responsible for glycogen breakdown and synthesis can impair mitochondrial function. For instance, in GSD type II (Pompe disease), acid alpha-glucosidase deficiency leads to lysosomal glycogen accumulation, which secondarily impacts mitochondrial function through dysfunctional mitophagy, which disrupts mitochondrial quality control, generating oxidative stress. In GSD type III (Cori disease), the lack of the debranching enzyme causes glycogen accumulation and affects mitochondrial dynamics and biogenesis by disrupting the integrity of muscle fibers. Malfunctional glycogen metabolism can disrupt various cascades, thus causing mitochondrial and cell metabolic dysfunction through various mechanisms. These dysfunctions include altered mitochondrial morphology, impaired oxidative phosphorylation, increased production of reactive oxygen species (ROS), and defective mitophagy. The oxidative burden typical of GSDs compromises mitochondrial integrity and exacerbates the metabolic derangements observed in GSDs. The intertwining of mitochondrial dysfunction and GSDs underscores the complexity of these disorders and has significant clinical implications. GSD patients often present with multisystem manifestations, including hepatomegaly, hypoglycemia, and muscle weakness, which can be exacerbated by mitochondrial impairment. Moreover, mitochondrial dysfunction may contribute to the progression of GSD-related complications, such as cardiomyopathy and neurocognitive deficits. Targeting mitochondrial dysfunction thus represents a promising therapeutic avenue in GSDs. Potential strategies include antioxidants to mitigate oxidative stress, compounds that enhance mitochondrial biogenesis, and gene therapy to correct the underlying mitochondrial enzyme deficiencies. Mitochondrial dysfunction plays a critical role in the pathophysiology of GSDs. Recognizing and addressing this aspect can lead to more comprehensive and effective treatments, improving the quality of life of GSD patients. This review aims to elaborate on the intricate relationship between mitochondrial dysfunction and various types of GSDs. The review presents challenges and treatment options for several GSDs. Full article
(This article belongs to the Special Issue Molecular Diagnosis and Regulation of Mitochondrial Dysfunction)
Show Figures

Figure 1

Figure 1
<p>Glycogen metabolism pathway’s enzymes defects and GSDs; (defective enzymes are in cyan color and resulting GSDs are in dark brown circle, metabolism products are shown in light brown color (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 8 August 2024).</p>
Full article ">Figure 2
<p>Major mitochondrial dysfunctions in glycogen storage disorders (GSDs), (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 17 August 2024).</p>
Full article ">Figure 3
<p>Mitochondrial dysfunction and tumorigenesis by excessive ROS formation, (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 17 August 2024).</p>
Full article ">Figure 4
<p>(<b>A</b>) Mitochondria biogenesis (<b>B</b>) PGC-1α downregulation impaired Mitochondrial ROS defense (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 27 August 2024).</p>
Full article ">Figure 5
<p>Mitophagy starts with stressed and dysfunctional mitochondria; impaired mitochondria loss membrane potential, PINK1/parkin RBR E3 ubiquitin protein ligase complex formation, phagophore formation and lysosomal degradation (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 17 August 2024).</p>
Full article ">Figure 6
<p>Mitochondrial apoptotic pathway (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 22 August 2024).</p>
Full article ">Figure 7
<p>OxPHOS complex proteins imbalance and myopathy (created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 21 August 2024).</p>
Full article ">
22 pages, 13561 KiB  
Article
pH-Sensitive Fluorescent Probe in Nanogel Particles as Theragnostic Agent for Imaging and Elimination of Latent Bacterial Cells Residing Inside Macrophages
by Igor D. Zlotnikov, Alexander A. Ezhov, Natalya G. Belogurova and Elena V. Kudryashova
Gels 2024, 10(9), 567; https://doi.org/10.3390/gels10090567 - 30 Aug 2024
Viewed by 371
Abstract
Rhodamine 6G (R6G) and 4-nitro-2,1,3-benzoxadiazole (NBD) linked through a spacer molecule spermidine (spd), R6G-spd-NBD, produces a fluorescent probe with pH-sensitive FRET (Förster (fluorescence) resonance energy transfer) effect that can be useful in a variety of diagnostic applications. Specifically, cancer cells can be spotted [...] Read more.
Rhodamine 6G (R6G) and 4-nitro-2,1,3-benzoxadiazole (NBD) linked through a spacer molecule spermidine (spd), R6G-spd-NBD, produces a fluorescent probe with pH-sensitive FRET (Förster (fluorescence) resonance energy transfer) effect that can be useful in a variety of diagnostic applications. Specifically, cancer cells can be spotted due to a local decrease in pH (Warburg effect). In this research, we applied this approach to intracellular infectious diseases—namely, leishmaniasis, brucellosis, and tuberculosis, difficult to treat because of their localization inside macrophages. R6G-spd-NBD offers an opportunity to detect such bacteria and potentially deliver therapeutic targets to treat them. The nanogel formulation of the R6G-spd-NBD probe (nanoparticles based on chitosan or heparin grafted with lipoic acid residues, Chit-LA and Hep-LA) was obtained to improve the pH sensitivity in the desired pH range (5.5–7.5), providing selective visualization and targeting of bacterial cells, thereby enhancing the capabilities of CLSM (confocal laser scanning microscopy) imaging. According to AFM (atomic force microscopy) data, nanogel particles containing R6G-spd-NBD of compact structure and spherical shape are formed, with a diameter of 70–100 nm. The nanogel formulation of the R6G-spd-NBD further improves absorption and penetration into bacteria, including those located inside macrophages. Due to the negative charge of the bacteria surface, the absorption of positively charged R6G-spd-NBD, and even more so in the chitosan derivatives’ nanogel particles, is pronounced. Additionally, with a pH-sensitive R6G-spd-NBD fluorescent probe, the macrophages’ lysosomes can be easily distinguished due to their acidic pH environment. CLSM was used to visualize samples of macrophage cells containing absorbed bacteria. The created nanoparticles showed a significant selectivity to model E. coli vs. Lactobacillus bacterial cells, and the R6G-spd-NBD agent, being a mild bactericide, cleared over 50% E.coli in conditions where Lactobacillus remained almost unaffected. Taken together, our data indicate that R6G-spd-NBD, as well as similar compounds, can have value not only for diagnostic, but also for theranostic applications. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic representation of the polysaccharide nanogel particles loaded with theranostic agent R6G-spd-NBD with a pH-sensitivity function. Left panel delineates the architectures of amphiphilic polymers and the fluorescent probe, the central panel presents a schematic depiction of nanogel particles, and the right panel illustrates the correlation between the fluorescence intensity at a wavelength of 550 nm and pH for a therapeutic formulation based on R6G-spd-NBD. The ellipsoidal shape is a schematic representation of the polar head of a polymer forming a gel nanoparticle composed of polymer chains derived from chitosan or heparin. The lines depict the hydrophobic tails of the acidic residues.</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectra and structure of R6G-spd-NBD. PBS (0.01 M, pH = 7.4). (<b>b</b>) The <sup>1</sup>H NMR spectra of R6G-spd-NBD. (<b>c</b>) The <sup>13</sup>C NMR spectra of R6G-spd-NBD. T = 22 °C. DMSO-d<sub>6</sub>. Working frequencies: 500.13 MHz (<sup>1</sup>H) and 125.76 MHz (<sup>13</sup>C).</p>
Full article ">Figure 3
<p>(<b>a</b>) Atomic force microscopy images of non-loaded Chit5-LA in nanoparticles in 2D view. (<b>b</b>) Atomic force microscopy images of Chit5-LA nanoparticles loaded with R6G-spd-NBD (10 mass.%). (<b>c</b>) Atomic force microscopy images of non-loaded Chit5-LA nanoparticles in 3D view. (<b>d</b>) The corresponding height section of (<b>b</b>) through the main diagonal from top to bottom.</p>
Full article ">Figure 4
<p>Fluorescence emission spectra of R6G (1 µM) and the pH sensor R6G-spd-NBD (1 µM): (1) in PBS buffer (<b>a</b>) prior to and (<b>b</b>) following a 2 h incubation at 37 °C; (2) mixed with an <span class="html-italic">E. coli</span> bacterial suspension (10<sup>7</sup> CFU/mL) in PBS (<b>c</b>) prior to and (<b>d</b>) following a 2 h incubation at the same temperature; (3) mixed with a <span class="html-italic">lactobacillus</span> suspension (10<sup>7</sup> CFU/mL) in PBS (<b>e</b>) prior to and (<b>f</b>) following the same incubation period at 37 °C. λ<sub>exci</sub> = 460 nm. PBS (0.01 M, pH = 7.4). T = 37 °C.</p>
Full article ">Figure 5
<p>(<b>a</b>) Kinetic parameters for the 1 h interaction between R6G or R6G-spd-NBD pH sensors with <span class="html-italic">E. coli</span> and <span class="html-italic">lactobacillus</span> suspensions (10<sup>7</sup> CFU/mL): the slopes of the fluorescence emission intensity curves at 490 nm (NBD &gt; R6G) and 550 nm (R6G &gt; NBD). The ratio of tangents of kinetic lines at 490 nm and 550 nm for <span class="html-italic">E. coli</span> and <span class="html-italic">lactobacillus</span>. λ<sub>exci</sub> = 460 nm. PBS (0.01 M, pH = 7.4). T = 37 °C. (<b>b</b>) Fluorescent image of a Petri dish with <span class="html-italic">E. coli</span> cells (10<sup>7</sup> CFU/0.5 mL placed onto 20 mL of agar medium) when exposed to R6G or R6G-spd-NBD in free form and in nanogel particles for 12 h at 37 °C. The fluorophores are placed in wells with a diameter of 9 mm. The concentration of fluorophores is 0.1 mg/mL, the concentration of amphiphilic polymers is 0.5 mg/mL. λ<sub>exci, max</sub> = 480 nm. λ<sub>emi</sub> = 515–570 nm.</p>
Full article ">Figure 6
<p>Confocal laser scanning microscopy images of (<b>a</b>,<b>b</b>) <span class="html-italic">E. coli</span> cells and (<b>c</b>,<b>d</b>) <span class="html-italic">Lactobacilli</span> cells, labelled with NBD-spd, R6G, or R6G-spd-NBD (1 µg/mL for all markers) in free or nanogel form. λ<sub>exci</sub> = 488 nm, λ<sub>emi</sub> = 510–560 nm (green channel), λ<sub>emi</sub> = 560–610 nm (red channel). The scale bar is 20 µm. C<sub>mic</sub> = 0.1 mg/mL.</p>
Full article ">Figure 6 Cont.
<p>Confocal laser scanning microscopy images of (<b>a</b>,<b>b</b>) <span class="html-italic">E. coli</span> cells and (<b>c</b>,<b>d</b>) <span class="html-italic">Lactobacilli</span> cells, labelled with NBD-spd, R6G, or R6G-spd-NBD (1 µg/mL for all markers) in free or nanogel form. λ<sub>exci</sub> = 488 nm, λ<sub>emi</sub> = 510–560 nm (green channel), λ<sub>emi</sub> = 560–610 nm (red channel). The scale bar is 20 µm. C<sub>mic</sub> = 0.1 mg/mL.</p>
Full article ">Figure 7
<p>Fluorescence microscopy images of macrophage (large) and bacterial (small) cells from human bronchoalveolar lavage (BAL) labelled with (<b>a</b>) R6G-spd-NBD or (<b>b</b>) R6G (1 µg/mL for all fluorophores) in nanogel form. Blue channel: λ<sub>exci, max</sub> = 360 nm, λ<sub>emi</sub> = 425–700 nm. Green channel: λ<sub>exci, max</sub> = 475 nm, λ<sub>emi</sub> = 515–700 nm. Additionally, the blue and green overlay channels and the fluorescence overlay channel on brightfield are shown. The scale bar is 20 µm. C<sub>mic</sub> = 0.1 mg/mL.</p>
Full article ">
Back to TopTop