[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,341)

Search Parameters:
Keywords = endocytosis

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 5171 KiB  
Article
Molecular Dynamics Reveal Key Steps in BAR-Related Membrane Remodeling
by Shenghan Song, Tongtong Li, Amy O. Stevens, Temair Shorty and Yi He
Pathogens 2024, 13(10), 902; https://doi.org/10.3390/pathogens13100902 - 15 Oct 2024
Viewed by 304
Abstract
Endocytosis plays a complex role in pathogen-host interactions. It serves as a pathway for pathogens to enter the host cell and acts as a part of the immune defense mechanism. Endocytosis involves the formation of lipid membrane vesicles and the reshaping of the [...] Read more.
Endocytosis plays a complex role in pathogen-host interactions. It serves as a pathway for pathogens to enter the host cell and acts as a part of the immune defense mechanism. Endocytosis involves the formation of lipid membrane vesicles and the reshaping of the cell membrane, a task predominantly managed by proteins containing BAR (Bin1/Amphiphysin/yeast RVS167) domains. Insights into how BAR domains can remodel and reshape cell membranes provide crucial information on infections and can aid the development of treatment. Aiming at deciphering the roles of the BAR dimers in lipid membrane bending and remodeling, we conducted extensive all-atom molecular dynamics simulations and discovered that the presence of helix kinks divides the BAR monomer into two segments—the “arm segment” and the “core segment”—which exhibit distinct movement patterns. Contrary to the prior hypothesis of BAR domains working as a rigid scaffold, we found that it functions in an “Arms-Hands” mode. These findings enhance the understanding of endocytosis, potentially advancing research on pathogen-host interactions and aiding in the identification of new treatment strategies targeting BAR domains. Full article
(This article belongs to the Special Issue Current Research on Host–Pathogen Interaction in 2024)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Root Mean Square Deviation (RMSD) averaged over 10 trajectories. (<b>b</b>) Root Mean Square Fluctuation (RMSF). (<b>c</b>) Secondary structure analysis of Helix 2 in the 4ATM BAR monomer. (<b>d</b>) Secondary structure analysis of Helix 3 in the 4ATM BAR monomer. (<b>e</b>) Secondary structure segmentation of the 4ATM BAR monomer.</p>
Full article ">Figure 2
<p>Hydrogen bonds. (<b>a</b>) Number of hydrogen bonds with error bands. Blue lines represent hydrogen bonds between the 4ATM BAR domain and the membrane; pink lines represent hydrogen bonds between positively charged amino acids and the membrane. Solid lines indicate the mean number of hydrogen bonds, while faded lines depict error bands. (<b>b</b>) Hydrogen bonds between lipid molecules and the membrane. (<b>c</b>) Hydrogen bonds between amino acid residue groups and the membrane. (<b>d</b>) Positions of amino acid residue groups and bound lipid molecules. Residues colored red have hydrogen bond interactions with lipids; residues colored black have minimal hydrogen bonding with lipids.</p>
Full article ">Figure 3
<p>Curvature of the membrane and 4ATM BAR dimer. (<b>a</b>). Red line and dots with error band: Curvature of the membrane with the 4ATM BAR over time. The results are the average of 10 trajectories. Black line: Curvature of the membrane without the 4ATM BAR. The curvature is the average of 300 ns simulations. Blue line: Curvature of the 4ATM BAR crystal for RCSB. (<b>b</b>–<b>d</b>): Curved membrane with the 4ATM BAR showing the fitting radiuses. Lipid molecules are shown in cyan, in line style, while phosphorus atoms are shown in yellow.</p>
Full article ">Figure 4
<p>Pearson correlation. (<b>a</b>). Segmentation of the 4ATM BAR. Red: “arm segments”. Blue: “core segments”. Black: Helix 2. Green: Span of the 4ATM BAR dimer. Distance between SER157 Chain A and SER157 Chain B. (<b>b</b>). Pearson correlation between the (<b><span class="html-italic">d</span></b>) distance and angles (<b><span class="html-italic">α</span>, <span class="html-italic">β</span>, and <span class="html-italic">γ</span></b>) and Pearson correlation between the “arm segment” (<b><span class="html-italic">α</span></b>) and “core segment” (<b><span class="html-italic">β</span></b>).</p>
Full article ">Figure 5
<p>Centripetal extrusion and upward bulge of the lipid membrane. (<b>a</b>). Definition of the horizontal distance from the lipid molecules bound at both ends to the central residue. The central residues are GLY69A and GLY69B, and the reference atom is Ca. The reference atom of lipid molecules is the phosphorus atom. (<b>b</b>). The 4ATM BAR with the membrane after 300ns of simulation. The orange atom marks the position of the membrane at 0 ns. (<b>c</b>). The average horizontal distance between lipid molecules and the center of the concave surface of the 4ATM BAR dimer. The calculated result is the average of ten trajectories. When calculating, the reference atom on the lipid molecule is a phosphorus atom. The average coordinates of GLY69A and GLY69B determine the center of the concave surface of the 4ATM BAR dimer.</p>
Full article ">
14 pages, 3997 KiB  
Article
Impact of Protein Coronas on Lipid Nanoparticle Uptake and Endocytic Pathways in Cells
by Rui Wang, Jing He, Yuhong Xu and Baowei Peng
Molecules 2024, 29(20), 4818; https://doi.org/10.3390/molecules29204818 - 11 Oct 2024
Viewed by 518
Abstract
Lipid nanoparticles (LNPs), widely used in disease diagnosis and drug delivery, face the challenge of being surrounded by biological macromolecules such as proteins upon entering the human body. These molecules compete for binding sites on the nanoparticle surfaces, forming a protein corona. The [...] Read more.
Lipid nanoparticles (LNPs), widely used in disease diagnosis and drug delivery, face the challenge of being surrounded by biological macromolecules such as proteins upon entering the human body. These molecules compete for binding sites on the nanoparticle surfaces, forming a protein corona. The impact of different types of protein coronas on LNP delivery remains unclear. In this study, we employed a newly developed, highly sensitive LNP labeling platform and analyzed the endocytosis of HeLa cells under different nutritional conditions using proteomics to address this critical issue. Our research found that under conditions of complete medium and amino acid starvation, most DNA-FITC vesicles in HeLa cells were located in the perinuclear region 4 h after transfection. In contrast, under serum starvation conditions, only a small portion of DNA-FITC vesicles were in the perinuclear region. On the other hand, through proteomics, we discovered that cells that were enriched in amino acids and complete medium contained more proteins, whereas those under serum starvation had relatively fewer enriched proteins. Through KEGG pathway enrichment analysis, we identified the phagosome and endocytosis pathways as particularly important. Lastly, differential analysis of proteins in these pathways revealed that proteins such as F-actin, Coronin, vATPase, TUBA, TUBB, MHCII, and TSP may have significant impacts on cellular endocytosis. Our research findings indicate that it is necessary to regulate cellular endocytosis based on different protein coronas to achieve optimal cytoplasmic release. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of intracellular uptake of dUTP-11-Biotin LNPs at different time points. (<b>A</b>). Endocytosis of LNPs DNA-FITC in HeLa cells pulsed with 2.5 µg DNA-FITC encapsulated in LNPs for 2 h, 4 h, 8 h, and 24 h. (<b>B</b>). HeLa cells were labeled with DAPI. (<b>C</b>). LNPs/lipids were labeled with DOPE-atto647. (<b>D</b>). LNPs/lipids were labeled with EEA1. Arrow heads point to LNP-DNA in perinuclear “cloud”, and arrows point to individual cytoplasm endosomes.</p>
Full article ">Figure 2
<p>The internalization of LNPs varies depending on different nutrients. (<b>A</b>). Correlation of peripheral LNP endosomes with endocytosis activity in AAs(−) HeLa cells and in Fed HeLa cells and in DMEM HeLa cells (Arrows point to individual cytoplasm endosomes.) (<b>B</b>). Quantification of SA-FITC intensity (<b>C</b>). Quantification of DOPE intensity (<b>D</b>). Quantification of EEA1 Area. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Protein corona analysis of LNPs. (<b>A</b>) SDS-PAGE gel indicating that the compositions of corona proteins DMEM, AAS(-), FED, AAS(-), AAS(-) + LNP, FED + LNP, and DMEM + LNP were co-incubated with serum for 8 h. (The bands within the red box are the protein bands sent for examination.) (<b>B</b>) The number of proteins. (<b>C</b>) A Venn diagram illustrating the unique and shared proteins between AAS(-) and AAS(-) + LNP. (<b>D</b>) A Venn diagram illustrating the unique and shared proteins between FED and AAS(-) + LNP. (<b>E</b>) A Venn diagram illustrating the unique and shared proteins between DMEM and DMEM + LNP (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>KEGG enrichment analysis; (<b>A</b>) KEGG analysis of the AAS(-) group; (<b>B</b>) KEGG analysis of the FED group; (<b>C</b>) KEGG analysis of the DMEM group (The red box contains the same signaling pathway).</p>
Full article ">Figure 5
<p>The signaling pathway diagram of phagosome (the images are sourced from the DAVID database 2024.05).</p>
Full article ">Figure 6
<p>Cluster heatmap analysis.</p>
Full article ">
23 pages, 6287 KiB  
Article
Leveraging the Aggregated Protein Dye YAT2150 for Malaria Chemotherapy
by Claudia Camarero-Hoyos, Inés Bouzón-Arnáiz, Yunuen Avalos-Padilla, Antonino Nicolò Fallica, Lucía Román-Álamo, Miriam Ramírez, Emma Portabella, Ona Cuspinera, Daniela Currea-Ayala, Marc Orozco-Quer, Maria Ribera, Inga Siden-Kiamos, Lefteris Spanos, Valentín Iglesias, Benigno Crespo, Sara Viera, David Andreu, Elena Sulleiro, Francesc Zarzuela, Nerea Urtasun, Sandra Pérez-Torras, Marçal Pastor-Anglada, Elsa M. Arce, Diego Muñoz-Torrero and Xavier Fernàndez-Busquetsadd Show full author list remove Hide full author list
Pharmaceutics 2024, 16(10), 1290; https://doi.org/10.3390/pharmaceutics16101290 - 30 Sep 2024
Viewed by 807
Abstract
Background/Objectives: YAT2150 is a first-in-class antiplasmodial compound that has been recently proposed as a new interesting drug for malaria therapy. Methods/Results: The fluorescence of YAT2150 rapidly increases upon its entry into Plasmodium, a property that can be of use for [...] Read more.
Background/Objectives: YAT2150 is a first-in-class antiplasmodial compound that has been recently proposed as a new interesting drug for malaria therapy. Methods/Results: The fluorescence of YAT2150 rapidly increases upon its entry into Plasmodium, a property that can be of use for the design of highly sensitive diagnostic approaches. YAT2150 blocks the in vitro development of the ookinete stage of Plasmodium and, when added to an infected blood meal, inhibits oocyst formation in the mosquito. Thus, the compound could possibly contribute to future transmission-blocking antimalarial strategies. Cell influx/efflux studies in Caco-2 cells suggest that YAT2150 is internalized by endocytosis and also through the OATP2B1 transporter, whereas its main export route would be via OSTα. YAT2150 has an overall favorable drug metabolism and pharmacokinetics profile, and its moderate cytotoxicity can be significantly reduced upon encapsulation in immunoliposomes, which leads to a dramatic increase in the drug selectivity index to values close to 1000. Although YAT2150 binds amyloid-forming peptides, its in vitro fluorescence emission is stronger upon association with peptides that form amorphous aggregates, suggesting that regions enriched in unstructured proteins are the preferential binding sites of the drug inside Plasmodium cells. The reduction of protein aggregation in the parasite after YAT2150 treatment, which has been suggested to be directly related to the drug’s mode of action, is also observed following treatment with quinoline antimalarials like chloroquine and primaquine. Conclusions: Altogether, the data presented here indicate that YAT2150 can represent the spearhead of a new family of compounds for malaria diagnosis and therapy due to its presumed novel mode of action based on the interaction with functional protein aggregates in the pathogen. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure of YAT2150.</p>
Full article ">Figure 2
<p>YAT2150 staining of clinical samples of <span class="html-italic">P. falciparum</span> and <span class="html-italic">P. ovale</span> infections. The merge panels refer to fluorescence images only. Arrowheads indicate the <span class="html-italic">Plasmodium</span>-infected red blood cells present in the microscope fields shown.</p>
Full article ">Figure 3
<p>Transmission blocking assays. (<b>A</b>,<b>B</b>) Ex vivo <span class="html-italic">P. berghei</span> ookinete maturation assay. (<b>A</b>) Scheme of the experimental protocol. (<b>B</b>) Effect of DONE3TCl and YAT2150 on ookinete development. Mean values ± standard deviations are indicated (n = 3). **: <span class="html-italic">p</span> &lt; 0.01 (one-way ANOVA, Dunnett’s post-hoc test). (<b>C</b>,<b>D</b>) Membrane feeding assay to test the effect of YAT2150 on <span class="html-italic">P. berghei</span> oocyst production. (<b>C</b>) Scheme of the experimental protocol. A total of 10 µM YAT2150 was added to blood infected with <span class="html-italic">P. berghei</span> and immediately fed to <span class="html-italic">A. gambiae</span> mosquitoes. Control feeding was the same blood mixed with DMSO at the same concentration as in the YAT2150-containing sample. The illustrative GFP-oocyst image is from Lantero et al. [<a href="#B33-pharmaceutics-16-01290" class="html-bibr">33</a>]. (<b>D</b>) Effect of YAT2150 on oocyst development in three independent experiments (Exp 1 to 3). Percentages in the graph indicate the prevalence of infection (% of infected mosquitoes). ****: <span class="html-italic">p</span> &lt; 0.0001 (Mann–Whitney non-parametric test).</p>
Full article ">Figure 4
<p>YAT2150 (<b>A</b>) influx and (<b>B</b>) efflux study in Caco-2 cells in FBS-free medium. Representative microscopy images of YAT2150 fluorescence from three to five independent experiments are shown. Each panel corresponds to a single inhibitor blocking a particular pathway (in parenthesis). The bar graphs show the quantification of YAT2150 accumulation as reported in Materials and Methods using a fluorescence imaging system and normalized to YAT2150 without inhibitors (YAT2150, red line). The negative control sample (control) corresponds to the fluorescence in cells cultured for the same time in the absence of YAT2150 (no image of this condition is provided because fluorescence was almost negligible). Results are the mean ± SEM from three to five independent experiments. Statistical significance relative to YAT2150 influx and efflux was determined by Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **<span class="html-italic">** p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Characterization of the fast-acting activity of YAT2150. (<b>A</b>) Stage arrest assay of 3D7 <span class="html-italic">P. falciparum</span> synchronized at schizont stage, treated for up to 72 h with the IC<sub>80</sub> of YAT2150 (200 nM). At the indicated times after treatment start, blood smears stained with Giemsa were prepared, and parasite population was noted for at least 100 pRBCs. Bars show the percentages of asexual blood stages present at each time; *: indicates the presence of only pyknotic and dead parasites of which the population could not be annotated (n = 2 independent experiments). (<b>B</b>) Parasite-killing profile of 3D7 <span class="html-italic">P. falciparum</span> parasites treated for 24 and 48 h with 10 times the IC<sub>50</sub> of YAT2150 or with fast- (in green, chloroquine and artesunate), moderate- (in orange, pyrimethamine), and slow-acting (in red, atovaquone) antimalarials. (<b>C</b>) Representative images of Giemsa-stained pRBCs in the stage arrest assay of panel (<b>A</b>).</p>
Full article ">Figure 6
<p>ThT analysis in <span class="html-italic">P. falciparum</span> asexual blood stage cultures of the effect on protein aggregation of four clinically used antimalarial drugs. ThT fluorescence assay of <span class="html-italic">P. falciparum</span> culture extracts normalized to have equal protein content, either non-treated or treated for 4 h with artemisinin, atovaquone, chloroquine, or primaquine at their respective in vitro IC<sub>50</sub> (10.8 nM, 1 nM, 7 nM, and 3 µM, respectively, as determined in our experimental setting). The <span class="html-italic">p</span>-values refer to the fluorescence intensity measured at the maximum emission wavelength.</p>
Full article ">Figure 7
<p>Interaction of YAT2150 with aggregative peptides. (<b>A</b>–<b>C</b>) Relative fluorescence emission intensity of solutions of non-disaggregated KDLLF, KVVNI, and LYWIYY peptides treated with YAT2150 and ThT. Peptides (12.5 µM each) were incubated in the presence of (<b>A</b>) 25 µM ThT or (<b>B</b>) 10 µM YAT2150 before proceeding to measuring fluorescence emission. The insets in panel (<b>A</b>) show representative TEM images of the aggregates formed by LYWIYY and KDLLF. (<b>C</b>) Fluorescence emission spectra of peptide-free 10 µM YAT2150 solutions in PBS (dashed line) and DMSO (solid line). (<b>D</b>) Transmission electron microscopy analysis of the effect of 100 nM YAT2150 on the aggregation of LYWIYY at 24 h and 48 h after undergoing a disaggregation process. Size bars: 1 µm (upper panels, 30,000×), 250 nm (lower panels, 120,000×).</p>
Full article ">
15 pages, 4617 KiB  
Article
Human Coronavirus 229E Uses Clathrin-Mediated Endocytosis as a Route of Entry in Huh-7 Cells
by Sabina Andreu, Inés Ripa, José Antonio López-Guerrero and Raquel Bello-Morales
Biomolecules 2024, 14(10), 1232; https://doi.org/10.3390/biom14101232 - 29 Sep 2024
Viewed by 507
Abstract
Human coronavirus 229E (HCoV-229E) is an endemic coronavirus responsible for approximately one-third of “common cold” cases. To infect target cells, HCoV-229E first binds to its receptor on the cell surface and then can follow different pathways, entering by direct fusion or by taking [...] Read more.
Human coronavirus 229E (HCoV-229E) is an endemic coronavirus responsible for approximately one-third of “common cold” cases. To infect target cells, HCoV-229E first binds to its receptor on the cell surface and then can follow different pathways, entering by direct fusion or by taking advantage of host cell mechanisms such as endocytosis. Based on the role of clathrin, the process can be classified into clathrin-dependent or -independent endocytosis. This study characterizes the role of clathrin-mediated endocytosis (CME) in HCoV-229E infection of the human hepatoma cell line Huh-7. Using specific CME inhibitory drugs, we demonstrated that blocking CME significantly reduces HCoV-229E infection. Additionally, CRISPR/Cas9-mediated knockout of the µ subunit of adaptor protein complex 2 (AP-2) further corroborated the role of CME, as KOs showed over a 50% reduction in viral infection. AP-2 plays an important role in clathrin recruitment and the maturation of clathrin-coated vesicles. Our study also confirmed that in Huh-7 cells, HCoV-229E requires endosomal acidification for successful entry, as viral entry decreased when treated with lysomotropic agents. Furthermore, the colocalization of HCoV-229E with early endosome antigen 1 (EEA-1), only present in early endosomes, suggested that the virus uses an endosomal route for entry. These findings highlight, for the first time, the role of CME in HCoV-229E infection and confirm previous data of the use of the endosomal route at a low pH in the experimental cell model Huh-7. Our results provide new insights into the mechanisms of entry of HCoV-229E and provide a new basis for the development of targeted antiviral therapies. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Viral Infections)
Show Figures

Figure 1

Figure 1
<p>Chemical inhibitors of CME block conjugate transferrin but not conjugate dextran internalization at non-cytotoxic doses. (<b>A</b>) Cell viability of Huh-7 cells treated with chlorpromazine, dynasore or pitstop 2 for 24 h. Viability was measured by a MTT assay and calculated as the percentage of cell viability compared to untreated cells; the columns represent the mean percentage of relative cellular viability ± S.D. (<span class="html-italic">n</span> = 4) after drug exposure. The dotted lines mark the area in which values are considered non cytotoxic. (<b>B</b>) The uptake of transferrin in Huh-7 cells is blocked by CME inhibitors. Cells were treated for 1 h with chlorpromazine, dynasore or pitstop 2. Then, they were maintained for 30 min on ice with Tf CF<sup>®</sup>543 (5 μg/mL). Finally, cells were incubated 5 min at 37 °C before fixation. (<b>C</b>) Quantification of Tf conjugate. ROIs from groups of 20 cells and three areas of each image were measured. Measurement of mean fluorescence intensity from the 555 channel in an ROI was performed. The mean percentage of fluorescence ± S.D. is shown. (<span class="html-italic">n</span> = 4); * <span class="html-italic">p</span> &lt; 0.05. (<b>D</b>) The uptake of dextran in the Huh-7 cells was not disrupted by the CME inhibitors. Cells were treated for 1 h with chlorpromazine, dynasore or pitstop 2. Then, they were maintained for 30 min on ice with CF<sup>®</sup>555 Labeled Dye Dextran 10,000 MW (5 μg/mL). Finally, the cells were incubated for 10 min at 37 °C before fixation. (<b>E</b>) Quantification of dextran conjugate. Same procedure as with Tf quantification. Scale bar = 20 µm. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>HCoV-229E infection is blocked by CME-inhibitory drugs. Huh-7 cells were treated with either 10 µM chlorpromazine, 100 µM dynasore or 50 µM pitstop 2 for 1 h and subsequently infected with HCoV-229E-GFP at an m.o.i. of 0.5. Cells were maintained in the presence of the drugs until they were collected at 20 h p.i. (<b>A</b>) Flow cytometry analysis: this graph shows the percentage of normalized infection at 20 h p.i. ± S.D. Triplicate experiments were performed (<span class="html-italic">n</span> = 4). (<b>B</b>) The fluorescence microscopy images show the GFP+ signal, which corresponds to viral infection. Cellular nuclei are stained with DAPI. Scale bar = 10 µm. (<b>C</b>) Infectious particles of the progeny virus were titrated in Huh-7 cells to determine the 50% tissue culture infective dose (TCID<sub>50</sub>)/mL. The graph shows the mean ± S.D. (<span class="html-italic">n</span> = 3) viral production. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Validation of the knockout of the AP2M1 gene in Huh-7 cells by CRISPR/Cas9. (<b>A</b>) AP2M1 expression was checked by immunoblotting. Western blot analysis of total cell lysates subjected to SDS-PAGE, showing AP2M1 for each subcloned cell. β-actin was chosen as the protein loading control. The black arrow points to the specific AP2M1 band. Values of immunoblot quantification are reported as mean ± S.D. (<span class="html-italic">n</span> = 3); * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Indel values of C5 and C7. Percentage of efficiency of CRISPR-Cas9 cut. (<b>C</b>) Sanger sequence view showing wild-type (control) and edited (C5, C7) sequences in the region around the guide sequence. The horizontal black underlined region represents the guide sequence. The horizontal red underline is the PAM site. The vertical black dotted line represents the actual cut site. Cutting and error-prone repair usually result in mixed sequencing bases after the cut. (<b>D</b>) Conjugate transferrin internalization is blocked in AP2M1-KO cells. The uptake of transferrin in KO AP2M1 Huh-7 cells C5 and C7 is blocked. Cells were maintained for 30 min on ice with Tf CF<sup>®</sup>543 (5 μg/mL). Finally, they were incubated 5 min at 37 °C before fixation. Nuclei were stained with DAPI. Scale bar = 20 µm. Original images can be found in <a href="#app1-biomolecules-14-01232" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 4
<p>The knockout of AP2M1 in Huh-7 cells contributed to the decrease in HCoV-229E infection. KO cells C5 and C7 and non-KO cells (wild-type) were infected with HCoV-229E-GFP at an m.o.i. of 0.5. At 20 h p.i., samples were processed, and infection was analyzed by the following techniques: (<b>A</b>) Flow cytometry data show the mean of the percentage of normalized infection 20 h p.i. (% GFP+ cells) ± S.D. (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05. The plots represent the histograms of GFP-positive (+) and GFP-negative (−) cells. (<b>B</b>) Infectious particles of the progeny virus for each condition were titrated in Huh-7 cells to determine the 50% tissue culture infective dose (TCID<sub>50</sub>)/mL. The graph shows the mean ± S.D. (<span class="html-italic">n</span> = 3) viral production. (<b>C</b>) Western blot analysis of total cell lysates subjected to SDS-PAGE, showing the viral GFP for each condition. β-actin was chosen as the protein loading control. Values of immunoblot quantification are reported as mean ± S.D. (<span class="html-italic">n</span> = 3); * <span class="html-italic">p</span> &lt; 0.05. Original images can be found in <a href="#app1-biomolecules-14-01232" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 5
<p>Inhibition of endosomal acidification prevents the entry of HCoV-229E in the Huh-7 cell line. Huh-7 cells were treated with NH<sub>4</sub>Cl at different concentrations for 1 h and during infection with HCoV-229E-GFP at an m.o.i. of 3. (<b>A</b>) Fluorescence microscopy images showing GFP-positive cells at 20 h p.i. and nuclei stained with DAPI for each condition. Scale bar = 100 µm. (<b>B</b>) The flow cytometry data report the mean percentage of GFP+ cells, normalized to the non-treated control. * <span class="html-italic">p</span> &lt; 0.05 (<span class="html-italic">n</span> = 4). (<b>C</b>) Huh-7 cells were treated with 20 mM NH<sub>4</sub>Cl from 1 h before infection to 6 h p.i. (short times) or from 6 to 24 h p.i. (long times), with HCoV-229E-GFP at an m.o.i. of 3. The flow cytometry analysis reports the mean percentage of GFP+ cells at 24 h p.i., normalized to the non-treated control. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 6
<p>Quantitative confocal analysis of the colocalization of HCoV-229E and EEA-1 signal hot spots. (<b>A</b>) Cells were maintained with HCoV-229E for 15 min on ice and then incubated for 30 min at 33 °C. Then, they were fixed and labeled with the EEA-1 primary antibody and AlexaFluor-555-conjugated secondary antibody (red). The nuclei were stained with DAPI (blue). In the far-right panel, all three signal bandwidth images are combined, with colocalized red and green signals seen as yellow. Scale bar = 20 µm. (<b>B</b>) Two lineal sections were selected for each signal (yellow lines). The red EEA-1 and green viral GFP signal intensities in this line were plotted as a function of X–Y distance across the cell. Scale bar = 10 µm.</p>
Full article ">
20 pages, 4296 KiB  
Article
Novel Anti-Trop2 Nanobodies Disrupt Receptor Dimerization and Inhibit Tumor Cell Growth
by Junwen Deng, Zhongmin Geng, Linli Luan, Dingwen Jiang, Jian Lu, Hanzhong Zhang, Bingguan Chen, Xinlin Liu and Dongming Xing
Pharmaceutics 2024, 16(10), 1255; https://doi.org/10.3390/pharmaceutics16101255 - 27 Sep 2024
Viewed by 530
Abstract
Background: Trop2 (trophoblast cell-surface antigen 2) is overexpressed in multiple malignancies and is closely associated with poor prognosis, thus positioning it as a promising target for pan-cancer therapies. Despite the approval of Trop2-targeted antibody–drug conjugates (ADCs), challenges such as side effects, drug resistance, [...] Read more.
Background: Trop2 (trophoblast cell-surface antigen 2) is overexpressed in multiple malignancies and is closely associated with poor prognosis, thus positioning it as a promising target for pan-cancer therapies. Despite the approval of Trop2-targeted antibody–drug conjugates (ADCs), challenges such as side effects, drug resistance, and limited efficacy persist. Recent studies have shown that the dimeric forms of Trop2 are crucial for its oncogenic functions, and the binding epitopes of existing Trop2-targeted drugs lie distant from the dimerization interface, potentially limiting their antitumor efficacy. Method: A well-established synthetic nanobody library was screened against Trop2-ECD. The identified nanobodies were extensively characterized, including their binding specificity and affinity, as well as their bioactivities in antigen-antibody endocytosis, cell proliferation, and the inhibition of Trop2 dimer assembly. Finally, ELISA based epitope analysis and AlphaFold 3 were employed to elucidate the binding modes of the nanobodies. Results: We identified two nanobodies, N14 and N152, which demonstrated high affinity and specificity for Trop2. Cell-based assays confirmed that N14 and N152 can facilitate receptor internalization and inhibit growth in Trop2-positive tumor cells. Epitope analysis uncovered that N14 and N152 are capable of binding with all three subdomains of Trop2-ECD and effectively disrupt Trop2 dimerization. Predictive modeling suggests that N14 and N152 likely target the epitopes at the interface of Trop2 cis-dimerization. The binding modality and mechanism of action demonstrated by N14 and N152 are unique among Trop2-targeted antibodies. Conclusions: we identified two novel nanobodies, N14 and N152, that specifically bind to Trop2. Importantly, these nanobodies exhibit significant anti-tumor efficacy and distinctive binding patterns, underscoring their potential as innovative Trop2-targeted therapeutics. Full article
(This article belongs to the Special Issue Nanosystems and Antibody/Peptide Modified Drugs for Cancer Treatment)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Screening and characterization of anti-Trop2 nanobodies. (<b>A</b>) Nanobodies were identified using phage ELISA. Fc fragment was employed as a negative control. (<b>B</b>) SDS-PAGE analysis of anti-Trop2 nanobodies. N13, N14, N124, N125, N128 and N152 exhibited good purity (approximately 95%). (<b>C</b>) SEC analysis of anti-Trop2 nanobodies. N14 and N152 are presented as single peaks and showed similar retention time (RT), demonstrating their great homogeneity.</p>
Full article ">Figure 2
<p>Specificity and binding activity of anti-Trop2 nanobodies. (<b>A</b>) The binding specificity of nanobodies identified by ELISA. hRS7 was used as a positive control, and an irrelevant nanobody (C5G2) served as a negative control. No antibodies showed an apparent reaction to Fc fragment and EpCAM. Both hRS7 and 10 anti-Trop2 nanobodies exhibited potent binding activity towards Trop2. The OD<sub>450</sub> values were shown as the mean ± SEM (n = 3). (<b>B</b>) EC<sub>50</sub> values of binding activity of nanobodies to Trop2. The values displayed in the heatmaps are the EC<sub>50</sub> values of antibody binding to Fc fragment, EpCAM, and Trop2. The EC<sub>50</sub> value of binding activity of N14 to Trop2 was 1.5 nM, while that of N152 was less than 1 nM.</p>
Full article ">Figure 3
<p>N14 and N152 bind to Trop2-positive tumor cells, mediate receptor internalization, and inhibit receptor recycling. (<b>A</b>,<b>B</b>) The cell-surface binding activity of N14 and N152 was quantified by flow cytometry in NCI-N87 and MDA-MB-231 cells. The majority of cells in the control group treated with the irrelevant nanobodies (C5G2) were unstained and migrated to the bottom left quadrant. In the experimental groups incubated with N14 and N152, the positive cells were located in the upper-left quadrant. Specifically, they represented 45% and 40% in NCI-N87 cells and 46.9% and 46.1% in MDA-MB-231 cells, respectively. The ridge plot characterizing mean fluorescence intensity is clearly shifted. (<b>C</b>,<b>D</b>) N14 and N152 mediate receptor internalization and inhibit receptor recycling in NCI-N87 cells. The irrelevant nanobody (C5G2) served as a negative control. N14 and N152 interfere Trop2 recycling on NCI-N87 and MDA-MB-231, resulting in sustained receptor internalization (2 h to 6 h). After 6 h, N14 and N152 mediate more significant receptor internalization in MDA-MB-231 cells, with internalization rates of 20.1% and 10.0% respectively, while in NCI-N87 they are 15% and 7.2%. The receptor internalization rate was shown as the mean ± SEM (n = 3). The statistical significance of the experiments was assessed using Student’s <span class="html-italic">t</span>-test (ns, no significant; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>N14 and N152 directly inhibit cell proliferation and migration. (<b>A</b>,<b>B</b>) N14 and N152 inhibit the proliferation of NCI-N87 and MDA-MB-231 cells. The data is presented through proliferation inhibition curves (Left) and the bar graphs (Right) at the maximum concentration. Both N14 and N152 inhibit the proliferation of NCI-N87 (<b>A</b>) and MDA-MB-231 (<b>B</b>), compared to the control (hRS7). The growth inhibition effect mediated by N14 is consistently higher than that of N152. The cell viability was reported as the mean ± SEM (n = 3). (<b>C</b>,<b>D</b>) Left: Western blot analysis of Trop2 expression on NCI-N87 and MDA-MB-231 after RNA interference. The Trop2 expression on NCI-N87 (<b>C</b>) and MDA-MB-231 (<b>D</b>) was significantly reduced after the action of Trop2-siRNA, compared with the control and NC-siRNA. Right: The proliferation inhibitory effects of N14 and N152 on Trop2-depleted NCI-N87 and Trop2-depleted MDA-MB-231. After Trop2 knockdown, N14 and N152 lost their growth inhibitory effect on NCI-N87 (<b>C</b>) and MDA-MB-231 (<b>D</b>). The cell viability was reported as the mean ± SEM (n = 3). (<b>E</b>,<b>F</b>) Inhibition of cell migration by N14 and N152 in Transwell assay. The number of migrating cells decreased after treatment with N14 and N152, compared with the control (hRS7). N14 and N152 had a significant migration inhibitory effect on NCI-N87. The number of migrating cells was reported as the mean ± SEM (n = 3). All statistical significance of the experiments was assessed using Student’s <span class="html-italic">t</span>-test (ns, no significant; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>N14 and N152 bind multiple domains of Trop2-ECD and inhibit Trop2 dimerization. (<b>A</b>) Schematic diagram of domain- or loop-substituted Trop2-ECD constructs. hTrop2-mCRD, hTrop2-mTY, hTrop2-mCPD, and hTrop2-mRCPD corresponded to the CRD (H27-L69), TY (T70-C145), CPD (D146-T274), and RCPD (Q237-Q252) of Trop2 replaced by that of murine homolog, respectively. All chimeric and wild-type Trop2 proteins were expressed in the 293T system. (<b>B</b>) ELISA analysis of the key domains of interaction between Trop2 and nanobodies. hRS7 recognized the RCPD, an exposed loop within the CPD of Trop2-ECD. N14 and N152 recognize multiple domains of Trop2-ECD. The OD<sub>450</sub> values are reported as the mean ± SEM (n = 3). (<b>C</b>) Heatmap of the epitope-mapping results observed for the anti-Trop2 nanobodies. The EC<sub>50</sub> value of nanobodies against different chimeric Trop2-ECD proteins was measured and displayed as a heatmap. (<b>D</b>) N14 and N152 inhibit Trop2 dimerization. Cells solely transfected with EV (lane 1) or co-transfected with EV (empty vector) and HA-Trop2 (lane 2) served as negative controls for dimerization. The cells co-transfected with Flag-Trop2 and HA-Trop2 as the experimental group (lanes 3-6). The cells treated with hRS7, N14, and N152 were lysed four hours later (Lanes 4, 5, and 6, respectively), with cells receiving no treatment serving as negative controls (Lane 3). The total lysate (used as an input) was probed with anti-Flag antibody and anti-HA antibody (Panels 1 and 2, respectively). The presence of Flag-Trop2 Please check that intended meaning is retained.in respective IP lanes was confirmed by probing with anti-Flag antibody (panel 3) and immune complexes pulled down by anti-Flag antibody were probed with anti-HA antibody for detection of co-immunoprecipitated HA-tagged monomers (panel 4).</p>
Full article ">Figure 6
<p>The binding sites of N14. (<b>A</b>) Left: Cartoon diagram illustrating the Trop2 <span class="html-italic">cis-</span>dimers. Right: The potential binding sites for N14 at <span class="html-italic">cis</span>-interfaces. The structural model of the <span class="html-italic">cis</span>-dimer with PDB ID 7E5N. The potential binding epitopes of N14 are indicated by the cyan lines, while the magenta circle encloses the binding epitopes (RCPD) of hRS7. The black lines represent two Trop2 monomers in <span class="html-italic">cis</span>-dimers. (<b>B</b>) The interaction between N14 and the Trop2-ECD monomer. The white segments in N14 represent the three CDRs. The enlarged area displays the details of the interaction interface. Residues that participate in hydrogen bond interactions are shown as sticks and labeled. Hydrogen bonds are shown as dashed yellow lines. (<b>C</b>) The interactions between N14 and the Trop2 <span class="html-italic">cis-</span>dimer. One subunit (Monomer B) in the <span class="html-italic">cis</span>-dimer structure is the cartoon representation, and the other subunit (Monomer C) presents a translucent surface format. The white segments in N14 represent the three CDRs. The enlarged area displays the details of the interaction interface. Residues that participate in hydrogen bond interactions are shown as sticks and labeled. Hydrogen bonds are shown as dashed yellow lines. Salt bonds are shown as dashed green lines. (<b>D</b>) The binding regions of hRS7 and N14 on the Trop2-ECD monomer (left) and <span class="html-italic">cis</span>-dimer (right). The control antibodies hRS7 are indicated in magenta, while the N14 are depicted in cyan. (<b>E</b>) Left: Compares the WT Trop2-ECD monomer (in gray) with the N14-bound Trop2-ECD monomer (in green). The RMSD was 0.525 Å. Right: Contrasts the WT <span class="html-italic">cis</span>-dimer (in gray) with the N14-bound <span class="html-italic">cis</span>-dimer (in green). The RMSD was 1.966 Å.</p>
Full article ">
17 pages, 4019 KiB  
Article
The Internalization Pathways of Liposomes, PLGA, and Magnetic Nanoparticles in Neutrophils
by Anastasiia Garanina, Daniil Vishnevskiy, Anastasia Chernysheva, Julia Malinovskaya, Polina Lazareva, Alevtina Semkina, Maxim Abakumov and Victor Naumenko
Biomedicines 2024, 12(10), 2180; https://doi.org/10.3390/biomedicines12102180 - 25 Sep 2024
Viewed by 425
Abstract
Background/Objectives: Neutrophils are emerging as promising candidates for cell-based nanodrug delivery to tumors due to their unique biological properties. This study aims to investigate the mechanisms of nanoparticle internalization by neutrophils, specifically focusing on liposomes, poly(lactic-co-glycolic acid) (PLGA), and magnetite nanoparticles. Understanding these [...] Read more.
Background/Objectives: Neutrophils are emerging as promising candidates for cell-based nanodrug delivery to tumors due to their unique biological properties. This study aims to investigate the mechanisms of nanoparticle internalization by neutrophils, specifically focusing on liposomes, poly(lactic-co-glycolic acid) (PLGA), and magnetite nanoparticles. Understanding these mechanisms could enhance the efficiency of neutrophil-based nanodrug delivery for cancer treatment. Methods: Neutrophils were isolated from the peripheral blood of mice bearing 4T1 mammary adenocarcinoma. Confocal microscopy, transmission electron microscopy, and flow cytometry were employed to evaluate the uptake of liposomes, PLGA, and magnetite nanoparticles by neutrophils. The effects of cultivation conditions, such as the presence or absence of plasma in the growth medium, were also examined. Additionally, the roles of immunoglobulins (IgG/IgM) and cell surface receptors (Fc and scavenger receptors) in nanoparticle internalization were explored. Results: All types of nanoparticles were successfully internalized by neutrophils, though the mechanisms of uptake varied. Plasma presence in the medium significantly influenced nanoparticle binding, particularly for PLGA nanoparticles. Internalization of PLGA nanoparticles was found to depend on the presence of IgG/IgM in the medium and Fc receptors on neutrophil surfaces, while scavenger receptors were not involved. Conclusions: Understanding the distinct endocytosis pathways for different nanoparticles can improve the efficacy of neutrophil loading with nanodrugs, potentially advancing the development of neutrophil-based cancer therapies. The findings underscore the importance of the extracellular environment in modulating nanoparticle uptake. Full article
Show Figures

Figure 1

Figure 1
<p>Nanoparticle (red) localization in neutrophils blood isolated from the 4T1 tumor-bearing mice after 1 h of co-incubation. Immunocytochemical staining with antibodies to the proteins EEA-1, Rab5, and Rab7 (green), cell nuclei are stained with DAPI (blue), confocal microscopy, Pearson correlation coefficient (R) values between the endosome markers and NPs are shown.</p>
Full article ">Figure 2
<p>MNP-Cy5 localization in neutrophil blood isolated from the 4T1 tumor-bearing mice: (<b>a</b>) microphotographs of control neutrophils cultivated in growth medium; (<b>b</b>–<b>d</b>) microphotographs of neutrophils incubated with MNPs-Cy5 for 1 h. TEM, green frames indicate areas of cells demonstrated at higher magnification, and arrows point to individual NPs (<b>b</b>) or NP conglomerates (<b>c</b>) inside vesicles or cytoplasm (<b>d</b>).</p>
Full article ">Figure 3
<p>Dynamics of nanoparticle accumulation in neutrophil blood isolated from the 4T1 tumor-bearing mice in the presence and absence of plasma in the culture medium: confocal microscopy results are presented as mean ± SEM, *** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05 (<span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p>Impact of different endocytosis pathways on nanoparticle uptake by neutrophils: graphs of fluorescence intensity from NPs accumulated in neutrophils (associated with <a href="#app1-biomedicines-12-02180" class="html-app">Figures S3–S5</a>). Results are presented as mean ± SEM, * <span class="html-italic">p</span> &lt; 0.05 (<span class="html-italic">t</span>-test). CME—clathrin-mediated endocytosis; CavME—caveolin-mediated endocytosis; MP—macropinocytosis.</p>
Full article ">Figure 5
<p>Effectiveness of PLGA-Cy5 NP interaction with isolated neutrophils after 1 h of co-incubation under various conditions: (<b>a</b>) fluorescence intensity from Cy5 dye, indicating the efficiency of NP binding by cells, confocal microscopy; (<b>b</b>) percentage of neutrophils associated with PLGA-Cy5 NPs relative to all cells of the subpopulation, flow cytometry. Results are presented as mean ± SD, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 (<span class="html-italic">t</span>-test).</p>
Full article ">
20 pages, 4706 KiB  
Article
Screening and Engineering Yeast Transporters to Improve Cellobiose Fermentation by Recombinant Saccharomyces cerevisiae
by Leonardo G. Kretzer, Marilia M. Knychala, Lucca C. da Silva, Isadora C. C. da Fontoura, Maria José Leandro, César Fonseca, Kevin J. Verstrepen and Boris U. Stambuk
Fermentation 2024, 10(9), 490; https://doi.org/10.3390/fermentation10090490 - 22 Sep 2024
Viewed by 528
Abstract
Developing recombinant Saccharomyces cerevisiae strains capable of transporting and fermenting cellobiose directly is a promising strategy for second-generation ethanol production from lignocellulosic biomass. In this study, we cloned and expressed in the S. cerevisiae CEN.PK2-1C strain an intracellular β-glucosidase (SpBGL7) from [...] Read more.
Developing recombinant Saccharomyces cerevisiae strains capable of transporting and fermenting cellobiose directly is a promising strategy for second-generation ethanol production from lignocellulosic biomass. In this study, we cloned and expressed in the S. cerevisiae CEN.PK2-1C strain an intracellular β-glucosidase (SpBGL7) from Spathaspora passalidarum and co-expressed the cellobiose transporter SiHXT2.4 from Scheffersomyces illinoinensis, and two putative transporters, one from Candida tropicalis (CtCBT1 gene), and one from Meyerozyma guilliermondii (MgCBT2 gene). While all three transporters allowed cell growth on cellobiose, only the MgCBT2 permease allowed cellobiose fermentation, although cellobiose consumption was incomplete. The analysis of the β-glucosidase and transport activities revealed that the cells stopped consuming cellobiose due to a drop in the transport activity. Since ubiquitinylation of lysine residues at the N- or C-terminal domains of the permease are involved in the endocytosis and degradation of sugar transporters, we constructed truncated versions of the permease lacking lysine residues at the C-terminal domain (MgCBT2ΔC), and at both the C- and N-terminal domain (MgCBT2ΔNΔC) and co-expressed these permeases with the SpBGL7 β-glucosidase in an industrial strain. While the strain harboring the MgCBT2ΔC transporter continued to produce incomplete cellobiose fermentations as the wild-type MgCBT2 permease, the strain with the MgCBT2ΔNΔC permease was able to consume and ferment all the cellobiose present in the medium. Thus, our results highlight the importance of expressing cellobiose transporters lacking lysine at the N- and C-terminal domains for efficient cellobiose fermentation by recombinant S. cerevisiae. Full article
Show Figures

Figure 1

Figure 1
<p>Cell growth (<b>A</b>) and cellobiose consumption (<b>B</b>) in YNB medium containing 20 g/L cellobiose by the indicated yeast strains harboring the intracellular <span class="html-italic">SpBGL7</span> β-glucosidase (strain B7), or this strain also transformed with the pGPD-426 plasmids containing the genes (<span class="html-italic">CtCBT1</span>, <span class="html-italic">MgCBT2</span>, and <span class="html-italic">SiHXT2.4</span>) encoding sugar transporters.</p>
Full article ">Figure 2
<p>Cellobiose consumption (<b>A</b>) and ethanol production (<b>B</b>) during fermentation of 20 g/L cellobiose in rich YP medium by 10 g dry cell weight/L of the indicated yeast strains harboring the <span class="html-italic">SpBGL7</span> β-glucosidase and the genes <span class="html-italic">CtCBT1</span>, <span class="html-italic">MgCBT2</span>, and <span class="html-italic">SiHXT2.4</span> encoding sugar transporters.</p>
Full article ">Figure 3
<p>Cellobiose consumption and ethanol production (<b>A</b>), and pNPβG intracellular hydrolysis and transport activities (<b>B</b>) by strain B7-CBT2 during fermentation of 20 g/L cellobiose in rich YP medium.</p>
Full article ">Figure 4
<p>Sequence alignment of the cytoplasmic N-terminal first 43 amino acids (<b>A</b>) and the cytoplasmic C-terminal last 43 amino acids of the protein, after the last transmembrane domain (<b>B</b>), deduced from the <span class="html-italic">MgCBT2</span> gene, and the truncated versions of <span class="html-italic">MgCBT2</span> at the C-terminal domain (<span class="html-italic">MgCBT2</span>ΔC), or at both the N- and C-terminal domains (<span class="html-italic">MgCBT2</span>ΔNΔC). The lysine residues with medium (blue) or high (red) ubiquitinylation potential were determined with the BDM-PUB [<a href="#B61-fermentation-10-00490" class="html-bibr">61</a>] and UbPred programs [<a href="#B62-fermentation-10-00490" class="html-bibr">62</a>].</p>
Full article ">Figure 5
<p>Cellobiose consumption (<b>A</b>) and ethanol production (<b>B</b>) during fermentation of 20 g/L cellobiose in rich YP medium by the indicated industrial yeast strains harboring the intracellular <span class="html-italic">SpBGL7</span> β-glucosidase and the <span class="html-italic">MgCBT2</span>, <span class="html-italic">MgCBT2</span>ΔC, and <span class="html-italic">MgCBT2</span>ΔNΔC cellobiose transporters.</p>
Full article ">Figure 6
<p>Co-fermentation of 20 g/L cellobiose plus 20 g/L xylose in rich YP medium by the recombinant industrial yeast strain MP-B7-CBT2ΔNΔC harboring the intracellular <span class="html-italic">SpBGL7</span> β-glucosidase and the <span class="html-italic">MgCBT2</span>ΔNΔC cellobiose transporter.</p>
Full article ">Figure 7
<p>Phylogenetic classification of β-glucosidases from various yeast and fungal hosts. The phylogenetic tree contains 15 enzyme sequences, and the numbers at the nodes represent percentage bootstrap values based on 1500 samplings. The abbreviation of each species is added after the name of the β-glucosidase genes: Ss = <span class="html-italic">Scheffersomyces stipitis</span>, Sp = <span class="html-italic">Spathaspora passalidarum</span>, Km = <span class="html-italic">Kluyveromyces marxianus</span>, Tn = <span class="html-italic">Thermotoga neapolitana</span>, Sv = <span class="html-italic">Streptomyces venezuelae</span>, Aa = <span class="html-italic">Aspergillus aculeatus</span>, Tr = <span class="html-italic">Trichoderma reesei</span>, Yl = <span class="html-italic">Yarrowia lipolytica</span>, Ma = <span class="html-italic">Moesziomyces antarcticus</span>, Nc = <span class="html-italic">Neurospora crassa</span>, and Bs = <span class="html-italic">Bacillus subtilus</span>.</p>
Full article ">Figure 8
<p>Phylogenetic classification of glucose, maltose, and cellobiose transporters from various yeast and fungal hosts. The phylogenetic tree contains 35 transporter sequences, and the numbers at the nodes represent percentage bootstrap values based on 1500 samplings. The abbreviation of each species is added after the name of the transporter genes: As = <span class="html-italic">Aspergillus nidulans</span>, Tt = <span class="html-italic">Thielavia terrestris</span>, Af = <span class="html-italic">Aspergillus flavus</span>, Pp = <span class="html-italic">Postia placenta</span>, Nc = <span class="html-italic">Neurospora crassa</span>, Fg = <span class="html-italic">Fusarium graminearum</span>, Ls = <span class="html-italic">Lipomyces starkeyi</span>, Kd = <span class="html-italic">Kluyveromyces dobzhanskii</span>, Ss = <span class="html-italic">Scheffersomyces stipitis</span>, Ct = <span class="html-italic">Candida tropicalis</span>, Mg = <span class="html-italic">Meyerozyma guilliermondii</span>, Km = <span class="html-italic">Kluyveromyces marxianus</span>, Kw = <span class="html-italic">Kluyveromyces wickerhamii</span>, Ka = <span class="html-italic">Kluyveromyces aestuarii</span>, Kl = <span class="html-italic">Kluyveromyces lactis</span>, Po = <span class="html-italic">Penicillium oxalicum</span>, An = <span class="html-italic">Aspergillus niger</span>, Tr = <span class="html-italic">Trichoderma reesei</span>, Pc = <span class="html-italic">Penicillium chrysogenum</span>, and Sc = <span class="html-italic">Saccharomyces cerevisiae</span>.</p>
Full article ">
12 pages, 1162 KiB  
Article
Gallium Uncouples Iron Metabolism to Enhance Glioblastoma Radiosensitivity
by Stephenson B. Owusu, Amira Zaher, Stephen Ahenkorah, Darpah N. Pandya, Thaddeus J. Wadas and Michael S. Petronek
Int. J. Mol. Sci. 2024, 25(18), 10047; https://doi.org/10.3390/ijms251810047 - 18 Sep 2024
Viewed by 568
Abstract
Gallium-based therapy has been considered a potentially effective cancer therapy for decades and has recently re-emerged as a novel therapeutic strategy for the management of glioblastoma tumors. Gallium targets the iron-dependent phenotype associated with aggressive tumors by mimicking iron in circulation and gaining [...] Read more.
Gallium-based therapy has been considered a potentially effective cancer therapy for decades and has recently re-emerged as a novel therapeutic strategy for the management of glioblastoma tumors. Gallium targets the iron-dependent phenotype associated with aggressive tumors by mimicking iron in circulation and gaining intracellular access through transferrin-receptor-mediated endocytosis. Mechanistically, it is believed that gallium inhibits critical iron-dependent enzymes like ribonucleotide reductase and NADH dehydrogenase (electron transport chain complex I) by replacing iron and removing the ability to transfer electrons through the protein secondary structure. However, information regarding the effects of gallium on cellular iron metabolism is limited. As mitochondrial iron metabolism serves as a central hub of the iron metabolic network, the goal of this study was to investigate the effects of gallium on mitochondrial iron metabolism in glioblastoma cells. Here, it has been discovered that gallium nitrate can induce mitochondrial iron depletion, which is associated with the induction of DNA damage. Moreover, the generation of gallium-resistant cell lines reveals a highly unstable phenotype characterized by impaired colony formation associated with a significant decrease in mitochondrial iron content and loss of the mitochondrial iron uptake transporter, mitoferrin-1. Moreover, gallium-resistant cell lines are significantly more sensitive to radiation and have an impaired ability to repair any sublethal damage and to survive potentially lethal radiation damage when left for 24 h following radiation. These results support the hypothesis that gallium can disrupt mitochondrial iron metabolism and serve as a potential radiosensitizer. Full article
Show Figures

Figure 1

Figure 1
<p>Ga(NO<sub>3</sub>)<sub>3</sub> causes mitochondrial iron depletion associated with DNA damage induction. (<b>A</b>) Cell counts in U251 and U87 GBM cells following a 48 h, 500 µM Ga(NO<sub>3</sub>)<sub>3</sub> treatment. Error bars represent mean ± SD of three measures with * <span class="html-italic">p</span> &lt; 0.05 using Welch’s <span class="html-italic">t</span>-test. (<b>B</b>) Western blot analysis of transferrin receptor (TfR) and ferritin heavy chain (FtH) expression following a 24 h, 500 µM Ga(NO<sub>3</sub>)<sub>3</sub> treatment. β-actin is used as a loading control. (<b>C</b>) Baseline mitochondrial iron content in U251 and U87 GBM cells (normalized to U251 cells). Error bars represent mean ± SD of three measures with * <span class="html-italic">p</span> &lt; 0.05 using Welch’s <span class="html-italic">t</span>-test. (<b>D</b>) Temporal effects of 500 µM Ga(NO<sub>3</sub>)<sub>3</sub> on mitochondrial iron content measured at 3 and 24 h. Error bars represent mean ± SD of three measures with * <span class="html-italic">p</span> &lt; 0.05 using a two-way ANOVA test. (<b>E</b>) Single-stranded DNA damage analyzed using an alkaline comet assay in U251 cells following a 24 h treatment of Ga(NO<sub>3</sub>)<sub>3</sub>. Error bars represent mean ± SD of three measures with * <span class="html-italic">p</span> &lt; 0.05 using Welch’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>Gallium-resistant cells are highly unstable and have impaired mitochondrial iron uptake. (<b>A</b>) Gallium-resistant U251 cell lines were generated via three separate clonal selections following a 72 h, 500 µM treatment of Ga(NO<sub>3</sub>)<sub>3</sub> for generating stable cell lines. (<b>B</b>) Ga(NO<sub>3</sub>)<sub>3</sub> resistance was confirmed using a clonogenic survival assay with three separate clonally selected cell lines (GaR1,2,3) and parental U251 cells being treated with a 500 µM treatment of Ga(NO<sub>3</sub>)<sub>3</sub> for 72 h. Error bars represent mean ± SEM (n = 3) with * <span class="html-italic">p</span> &lt; 0.05 using Welch’s <span class="html-italic">t</span>-test. (<b>C</b>,<b>D</b>) Clonally selected cell lines (GaR1,2,3) with confirmed resistance to Ga(NO<sub>3</sub>)<sub>3</sub> exhibit noticeable decreases in colony formation when plated as single cells (500 cells, <b>C</b>). This translates to a significant decrease in plating efficiency where plating efficiency (%) = # colonies counted/# cells plated (<b>D</b>). To combat this experimental difference, GaR1/2 were plated at higher concentrations (1000–2000 cells) for subsequent studies. Error bars represent mean ± SEM (n = 3) with * <span class="html-italic">p</span> &lt; 0.05 using a one-way ANOVA with a post hoc Tukey’s test for multiple comparisons. (<b>E</b>) Parental U251, GaR1, and GaR2 cells were incubated with 200 µCi <sup>67</sup>Ga-citrate for 2 h prior to harvesting, lysing, and analysis with a gamma counter to evaluate gallium uptake. (<b>F</b>) Analysis of basal mitochondrial iron content using MitoFerroGreen flow cytometry in Parental U251, GaR1, and GaR2 cells. Error bars represent mean ± SEM (n = 3) with * <span class="html-italic">p</span> &lt; 0.05 using a one-way ANOVA with a post hoc Tukey’s test for multiple comparisons. (<b>G</b>) Western blot analysis of mitoferrin-1 expression (Mfrn-1) in the 3 separate generations of GaR1 and GaR2 cells following each individual round of clonal selection compared to parental U251 cells show an evolutionary loss of Mfrn-1.</p>
Full article ">Figure 3
<p>Mitochondrial iron metabolism modulates cell radiosensitivity. (<b>A</b>) Clonogenic survival analysis of radiation dose-dependent cell killing in parental U251, GaR1, and GaR2 cell lines. (<b>B</b>) Clonogenic survival analysis of U251, GaR1, and GaR2 cell lines treated with 2 Gy radiation and left for 24 h prior to plating. Error bars represent mean ± SEM (n = 3) with * <span class="html-italic">p</span> &lt; 0.05 using a two-way ANOVA test with a post hoc test for multiple comparisons.</p>
Full article ">
20 pages, 10279 KiB  
Article
Exploration into Galectin-3 Driven Endocytosis and Lattices
by Massiullah Shafaq-Zadah, Estelle Dransart, Satish Kailasam Mani, Julio Lopes Sampaio, Lydia Bouidghaghen, Ulf J. Nilsson, Hakon Leffler and Ludger Johannes
Biomolecules 2024, 14(9), 1169; https://doi.org/10.3390/biom14091169 - 18 Sep 2024
Viewed by 591
Abstract
Essentially all plasma membrane proteins are glycosylated, and their activity is regulated by tuning their cell surface dynamics. This is achieved by glycan-binding proteins of the galectin family that either retain glycoproteins within lattices or drive their endocytic uptake via the clathrin-independent glycolipid-lectin [...] Read more.
Essentially all plasma membrane proteins are glycosylated, and their activity is regulated by tuning their cell surface dynamics. This is achieved by glycan-binding proteins of the galectin family that either retain glycoproteins within lattices or drive their endocytic uptake via the clathrin-independent glycolipid-lectin (GL-Lect) mechanism. Here, we have used immunofluorescence-based assays to analyze how lattice and GL-Lect mechanisms affect the internalization of the cell adhesion and migration glycoprotein α5β1 integrin. In retinal pigment epithelial (RPE-1) cells, internalized α5β1 integrin is found in small peripheral endosomes under unperturbed conditions. Pharmacological compounds were used to competitively inhibit one of the galectin family members, galectin-3 (Gal3), or to inhibit the expression of glycosphingolipids, both of which are the fabric of the GL-Lect mechanism. We found that under acute inhibition conditions, endocytic uptake of α5β1 integrin was strongly reduced, in agreement with previous studies on the GL-Lect driven internalization of the protein. In contrast, upon prolonged inhibitor treatment, the uptake of α5β1 integrin was increased, and the protein was now internalized by alternative pathways into large perinuclear endosomes. Our findings suggest that under these prolonged inhibitor treatment conditions, α5β1 integrin containing galectin lattices are dissociated, leading to an altered endocytic compartmentalization. Full article
(This article belongs to the Special Issue Cell Biology and Biomedical Application of Galectins)
Show Figures

Figure 1

Figure 1
<p>Gal3-based duality in β<sub>1</sub> integrin dynamics. (<b>A</b>) Schematic representation of the molecular organization of Gal3 where the C-terminal carbohydrate recognition domain (CRD) and the N-terminal oligomerization domain are indicated. (<b>B</b>) <b>Left:</b> Representative region of interest (leading edge) of a 2D STORM image of an RPE-1 cell showing surface-bound Gal3 (<b>top</b>) and the corresponding clusters obtained after segmentation (<b>bottom</b>). <b>Right</b>: The occurrence of each type of Gal3 cluster is shown in function of their surface area (×10<sup>3</sup> nm<sup>2</sup>). Means ± SEM, one-way ANOVA; ns = <span class="html-italic">p</span> &gt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001. Scale bar = 5 μm. (<b>C</b>) The percentage of Gal3 molecules is shown in function of the different Gal3 cluster populations. Means ± SEM, one-way ANOVA; ns = <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.002, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>D</b>) 2D STORM image of Gal3 and anti-β<sub>1</sub> integrin antibodies on RPE-1 cells (4 °C co-binding). A zoom of the leading edge is shown to illustrate the extensive level of colocalization between Gal3 and β<sub>1</sub> integrin. Gal3 clusters with variable size and shape are detected (yellow arrowheads in the zoom). Scale bars = 5 μm. (E) Quantification of the probability of proximity (colocalization) between Gal3 and β<sub>1</sub> integrin for each class of Gal3 clusters, compared to random non-clustered Gal3. Means ± SEM, one-way ANOVA; ns = <span class="html-italic">p</span> &gt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>F</b>) Top: Anti-β<sub>1</sub> integrin antibody uptake assay in RPE-1 cells. Internalized antibody is immunolabeled, imaged by confocal microscopy, and quantified. Bottom: Scheme of the protocol detailing the use of the Gal3 inhibitor I3 (10 μM) in acute versus prolonged incubation conditions. (<b>G</b>) Anti-β<sub>1</sub> integrin antibody uptake assay as in (<b>F</b>). Note the shift from both peripheral (green arrowheads) and perinuclear (red arrowheads) distribution of internalized β<sub>1</sub> integrin in control cells to exclusive perinuclear localization in the prolonged incubation condition. (<b>H</b>) Transferrin (Tf) internalization is very little affected by I3. In (<b>G</b>) and (<b>H</b>): Dashed lines indicate the contour of individual cells. Scale bars = 10 μm. Nuclei in blue (DAPI). Quantification of fluorescence intensities as means ± SEM, one-way ANOVA; ns = <span class="html-italic">p</span> &gt; 0.05, ** <span class="html-italic">p</span> &lt; 0.002, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>GSL-based duality in β<sub>1</sub> integrin dynamics. (<b>A</b>) Simplified schematic representation of the early steps of GSL synthesis. The reaction inhibited by Genz-123346 is indicated. (<b>B</b>) Analysis of cellular levels of the indicated GSL in function of incubation time with Genz-123346. Note that the most important drop occurs up to day 3. Means ± SEM, unpaired <span class="html-italic">t</span>-test; ns = <span class="html-italic">p</span> &gt; 0.05, ** <span class="html-italic">p</span> &lt; 0.002, **** <span class="html-italic">p</span> &lt; 0.0001. (<b>C</b>) Scheme of experimental procedure detailing how GSL inhibition has been set up either in acute (3 days) or prolonged (5 days) incubation conditions, prior to cargo protein internalization for 10 min. (<b>D</b>) Anti-β<sub>1</sub> integrin antibody uptake assay as in (<b>C</b>). Note that β<sub>1</sub> integrin uptake is inhibited upon acute Genz-123346 treatment and increased upon prolonged treatment. In the latter condition, the intracellular accumulation of β<sub>1</sub> integrin is massively perinuclear (red arrowheads), compared to control cells where peripheral localizations are also observed (green arrowheads). Means ± SEM, unpaired <span class="html-italic">t</span>-test; **** <span class="html-italic">p</span> &lt; 0.0001. (<b>E</b>) Transferrin (Tf) internalization (10 min) is only mildly affected in all conditions. Means ± SEM, unpaired <span class="html-italic">t</span>-test; ns = <span class="html-italic">p</span> &gt; 0.05, ** <span class="html-italic">p</span> &lt; 0.002. (<b>F</b>) Internalization of exogenous Gal3 (10 min). Similar to β<sub>1</sub> integrin, Gal3 endocytosis is significantly inhibited upon acute Genz-123346 treatment, and increased with perinuclear accumulation upon prolonged treatment (red arrowheads). Means ± SEM, unpaired <span class="html-italic">t</span>-test; **** <span class="html-italic">p</span> &lt; 0.0001. In (<b>D</b>–<b>F</b>): Yellow dashed lines indicate contours of cells; scale bars = 10 μm, nuclei in blue (DAPI).</p>
Full article ">Figure 3
<p>Characterization of sites of perinuclear β<sub>1</sub> integrin accumulation. (<b>A</b>,<b>B</b>) Anti-β<sub>1</sub> integrin or (<b>C</b>) Gal3 uptake assay (10 min) under acute or prolonged I3 (<b>A</b>) or Genz-123346 (<b>B</b>,<b>C</b>) treatment followed by immunolabeling for EEA1. The colocalization of β<sub>1</sub> integrin (<b>A</b>,<b>B</b>) or Gal3 (<b>C</b>) with EEA1 as well as the fluorescent intensity of EEA1 signal were quantified (<b>right</b>). Note the increased colocalization of internalized β<sub>1</sub> integrin (<b>A</b>,<b>B</b>) or Gal3 (<b>C</b>) with EEA1 and increased EEA1 signal intensity, notably in the prolonged treatment conditions. Means ± SEM, one-way ANOVA (<b>A</b>,<b>B</b>), or unpaired <span class="html-italic">t</span>-test (<b>C</b>); **** <span class="html-italic">p</span> &lt; 0.0001. Yellow dashed lines indicate contours of cells. Scale bars = 10 μm, nuclei in blue (DAPI).</p>
Full article ">Figure 3 Cont.
<p>Characterization of sites of perinuclear β<sub>1</sub> integrin accumulation. (<b>A</b>,<b>B</b>) Anti-β<sub>1</sub> integrin or (<b>C</b>) Gal3 uptake assay (10 min) under acute or prolonged I3 (<b>A</b>) or Genz-123346 (<b>B</b>,<b>C</b>) treatment followed by immunolabeling for EEA1. The colocalization of β<sub>1</sub> integrin (<b>A</b>,<b>B</b>) or Gal3 (<b>C</b>) with EEA1 as well as the fluorescent intensity of EEA1 signal were quantified (<b>right</b>). Note the increased colocalization of internalized β<sub>1</sub> integrin (<b>A</b>,<b>B</b>) or Gal3 (<b>C</b>) with EEA1 and increased EEA1 signal intensity, notably in the prolonged treatment conditions. Means ± SEM, one-way ANOVA (<b>A</b>,<b>B</b>), or unpaired <span class="html-italic">t</span>-test (<b>C</b>); **** <span class="html-italic">p</span> &lt; 0.0001. Yellow dashed lines indicate contours of cells. Scale bars = 10 μm, nuclei in blue (DAPI).</p>
Full article ">Figure 4
<p>Exogenous Gal3 and dextran 70K uptake upon prolonged GSL depletion. After prolonged (5 days) treatment with Genz-123346, RPE-1 cells were continuously co-incubated (10 min) with exogenous Gal3 and dextran 70K. Note the increased perinuclear accumulation of Gal3 and its increased overlap with dextran 70K under these conditions. Means ± SEM, unpaired <span class="html-italic">t</span>-test; **** <span class="html-italic">p</span> &lt; 0.0001. Yellow dashed lines indicate contours of cells. Scale bars = 10 μm, nuclei in blue (DAPI).</p>
Full article ">Figure 5
<p>Role of clathrin in endocytic uptake under prolonged treatment conditions. (<b>A</b>–<b>C</b>) Uptake assays (10 min) of anti-β<sub>1</sub> integrin antibodies (<b>A</b>,<b>B</b>) or Gal3 (<b>C</b>) upon prolonged I3 (<b>A</b>) or Genz-123346 (<b>B</b>,<b>C</b>) treatment. When indicated (siCHC), clathrin heavy chain was depleted (<b>right</b> images). The perinuclear accumulation of β<sub>1</sub> integrin (<b>A</b>,<b>B</b>) or that of Gal3 (<b>C</b>) as observed in the prolonged treatment conditions (red or white arrowheads) is strongly inhibited upon clathrin depletion. Means ± SEM, one-way ANOVA; **** <span class="html-italic">p</span> &lt; 0.0001. Yellow dashed lines indicate contours of cells. Scale bars = 10 μm, nuclei in blue (DAPI).</p>
Full article ">Figure 5 Cont.
<p>Role of clathrin in endocytic uptake under prolonged treatment conditions. (<b>A</b>–<b>C</b>) Uptake assays (10 min) of anti-β<sub>1</sub> integrin antibodies (<b>A</b>,<b>B</b>) or Gal3 (<b>C</b>) upon prolonged I3 (<b>A</b>) or Genz-123346 (<b>B</b>,<b>C</b>) treatment. When indicated (siCHC), clathrin heavy chain was depleted (<b>right</b> images). The perinuclear accumulation of β<sub>1</sub> integrin (<b>A</b>,<b>B</b>) or that of Gal3 (<b>C</b>) as observed in the prolonged treatment conditions (red or white arrowheads) is strongly inhibited upon clathrin depletion. Means ± SEM, one-way ANOVA; **** <span class="html-italic">p</span> &lt; 0.0001. Yellow dashed lines indicate contours of cells. Scale bars = 10 μm, nuclei in blue (DAPI).</p>
Full article ">Figure 6
<p>Continuum model between lattices and GL-Lect driven endocytosis. (<b>A</b>) Unperturbed condition. A glycoprotein cargo, here α<sub>5</sub>β<sub>1</sub> integrin, is either recruited into galectin lattices (<b>left</b>, underlined in red) or internalized by GL-Lect driven endocytosis (<b>right</b>, underlined in blue). (<b>B</b>) Acute treatment conditions. Since tubular endocytic pits for GL-Lect driven endocytosis are built de novo, acute interference with Gal3 activity or GSL expression prevents their formation. In contrast, preassembled galectin lattices resist under these conditions. (<b>C</b>) Prolonged treatment conditions. Even galectin lattices are disassembled. With GL-Lect driven endocytosis being inhibited, α<sub>5</sub>β<sub>1</sub> integrin is now internalized by alternative endocytic pathways, i.e., clathrin-mediated endocytosis and macropinocytosis.</p>
Full article ">
20 pages, 4220 KiB  
Review
Upgrading In Vitro Digestion Protocols with Absorption Models
by Otilia Antal, István Dalmadi and Krisztina Takács
Appl. Sci. 2024, 14(18), 8320; https://doi.org/10.3390/app14188320 - 15 Sep 2024
Viewed by 692
Abstract
Intestinal digestion and absorption are complex processes; thus, it is a challenge to imitate them realistically. There are numerous approaches available, with different disadvantages and advantages. The simplest methods to mimic absorption are the non-cell-based transport models but these lack important characteristics of [...] Read more.
Intestinal digestion and absorption are complex processes; thus, it is a challenge to imitate them realistically. There are numerous approaches available, with different disadvantages and advantages. The simplest methods to mimic absorption are the non-cell-based transport models but these lack important characteristics of enterocytes of the intestine. Therefore, the most often used method is to measure absorption through viable mammalian cells (most commonly Caco-2 cells, cultured on membrane insert plates), which not only assures the incorporation of brush border enzymes (responsible for the final digestion of peptides and disaccharides), it also simulates the absorption process. This means that influx/efflux transporter-facilitated transport, carrier-mediated transport, endocytosis, and transcytosis is also imitated besides passive diffusion. Still, these also lack the complexity of intestinal epithelium. Organoids or ex vivo models are a better approach if we want to attain precision but the highest accuracy can be achieved with microfluidic systems (gut-on-a-chip models). We propose that more research is necessary, and food absorption should also be studied on gut-on-a-chips, especially with fragmented organoids. Our review supports the choices of a proper intestinal epithelium model, which may have a key role in functional food development, nutrition studies, and toxicity assessment. Full article
(This article belongs to the Special Issue Feature Review Papers in Section ‘Food Science and Technology')
Show Figures

Figure 1

Figure 1
<p>The small intestinal epithelium cells and their localization [<a href="#B6-applsci-14-08320" class="html-bibr">6</a>].</p>
Full article ">Figure 2
<p>Transport processes in the intestinal epithelium [<a href="#B21-applsci-14-08320" class="html-bibr">21</a>].</p>
Full article ">Figure 3
<p>Approaches for the investigation of the absorption of food components. In Transwell<sup>®</sup> systems, primary cells, cell lines, or fragmented organoids derived from stem cells can be cultured.</p>
Full article ">Figure 4
<p>Schematic of transepithelial electrical resistance (TEER) measurements [<a href="#B58-applsci-14-08320" class="html-bibr">58</a>].</p>
Full article ">Figure 5
<p>The structure of organoids compared to the intestinal epithelium [<a href="#B64-applsci-14-08320" class="html-bibr">64</a>].</p>
Full article ">Figure 6
<p>The best-known method to simulate peristaltic motility in a gut-on-a-chip [<a href="#B42-applsci-14-08320" class="html-bibr">42</a>].</p>
Full article ">Figure 7
<p>The schematic organization of the “Intestine Chip” [<a href="#B73-applsci-14-08320" class="html-bibr">73</a>], based on [<a href="#B65-applsci-14-08320" class="html-bibr">65</a>].</p>
Full article ">Figure 8
<p>Illustration of the Ussing chamber [<a href="#B47-applsci-14-08320" class="html-bibr">47</a>].</p>
Full article ">Figure 9
<p>Illustration of everted gut sac [<a href="#B47-applsci-14-08320" class="html-bibr">47</a>].</p>
Full article ">
20 pages, 14709 KiB  
Article
Characterizing Extracellular Vesicles Generated from the Integra CELLine Culture System and Their Endocytic Pathways for Intracellular Drug Delivery
by Tianjiao Geng, Lei Tian, Song Yee Paek, Euphemia Leung, Lawrence W. Chamley and Zimei Wu
Pharmaceutics 2024, 16(9), 1206; https://doi.org/10.3390/pharmaceutics16091206 - 13 Sep 2024
Viewed by 729
Abstract
Extracellular vesicles (EVs) have attracted great attention as promising intracellular drug delivery carriers. While the endocytic pathways of small EVs (sEVs, <200 nm) have been reported, there is limited understanding of large EVs (lEVs, >200 nm), despite their potential applications for drug delivery. [...] Read more.
Extracellular vesicles (EVs) have attracted great attention as promising intracellular drug delivery carriers. While the endocytic pathways of small EVs (sEVs, <200 nm) have been reported, there is limited understanding of large EVs (lEVs, >200 nm), despite their potential applications for drug delivery. Additionally, the low yield of EVs during isolation remains a major challenge in their application. Herein, we aimed to compare the endocytic pathways of sEVs and lEVs using MIA PaCa-2 pancreatic cancer cell-derived EVs as models and to explore the efficiency of their production. The cellular uptake of EVs by MIA PaCa-2 cells was assessed and the pathways were investigated with the aid of endocytic inhibitors. The yield and protein content of sEVs and lEVs from the Integra CELLine culture system and the conventional flasks were compared. Our findings revealed that both sEVs and lEVs produced by the Integra CELLine system entered their parental cells via multiple routes, including caveolin-mediated endocytosis, clathrin-mediated endocytosis, and actin-dependent phagocytosis or macropinocytosis. Notably, caveolin- and clathrin-mediated endocytosis were more prominent in the uptake of sEVs, while actin-dependent phagocytosis and macropinocytosis were significant for both sEVs and lEVs. Compared with conventional flasks, the Integra CELLine system demonstrated a 9-fold increase in sEVs yield and a 6.5-fold increase in lEVs yield, along with 3- to 4-fold higher protein content per 1010 EVs. Given that different endocytic pathways led to distinct intracellular trafficking routes, this study highlights the unique potentials of sEVs and lEVs for intracellular cargo delivery. The Integra CELLine proves to be a highly productive and cost-effective system for generating EVs with favourable properties for drug delivery. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the Integra CELLine culture system (<b>A</b>) and conventional flasks (<b>B</b>) and the set-up to generate EVs. With the Integra CELLine culture system, cells can be continuously used for up to 300 days while medium containing EVs was collected twice per week. However, conventional flasks only can be used for one week and medium containing EVs was collected once.</p>
Full article ">Figure 2
<p>The structure of sulfo-cyanine5 NHS ester (Cy5) (MW 777.49 g/mol), a red fluorescent dye for EVs labelling. Labelling of EVs was achieved through covalent coupling between NHS ester groups and amine groups on proteins within the EVs. Excitation maximum ~646 nm; emission maximum ~662 nm.</p>
Full article ">Figure 3
<p>The EVs yield from the Integra CELLine culture system compared with that from conventional flasks. Data are expressed as EVs number per mL medium (<b>A</b>) and EVs protein (μg) per mL medium (<b>B</b>) (mean ± SD, <span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005, **** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Characterisation of EVs isolated from the Integra CELLine culture system. Nanoparticle tracking analysis (NTA) measurement of sEVs (<b>A</b>) and lEVs (<b>B</b>) isolated from MIA PaCa-2 cells. Size distribution of sEVs (<b>C</b>) and lEVs (<b>D</b>) measured by dynamic light scattering (DLS). Representative EVs samples in tubes (black arrows) and cryogenic transmission electron microscopic (cryo-TEM) micrographs of sEVs (<b>E</b>) and lEVs (red arrows) (<b>F</b>). Scale bars: 200 nm.</p>
Full article ">Figure 5
<p>The cellular uptake of Cy5- or PKH67-labelled sEVs and lEVs by their donor MIA PaCa-2 cells. (<b>A</b>) Fluorescence intensity of Cy5 (relative fluorescence units, RFU) in sEVs and lEVs cellular uptake at different time points measured by a fluorescence plate reader. Fluorescence images of sEVs (<b>B</b>) and lEVs (<b>C</b>) labelled by PKH67 internalised by MIA PaCa-2 cells for 1 h and 2 h, respectively. Scale bars: 20 μm. Cells were exposed to the same number of EVs in both experiments. Data represent the means ± SD, and <span class="html-italic">n</span> = 3 cells each data point. ** <span class="html-italic">p</span> &lt; 0.01. For clarity cells within the red boxed areas are highlighted in the upper panels and the arrows show that internalized EVs were prone to being entrapped in organelles.</p>
Full article ">Figure 6
<p>The effect of inhibitors concentrations on MIA PaCa-2 cells viability. Left: Images showing cell morphology after 3 h pre-treatment under light microscope (40×). Right: An MTT assay was carried out to evaluate the cell viability in the presence of inhibitors with different concentrations (mean ± SD, <span class="html-italic">n</span> = 3 batches). * <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.001, ns = nonsignificant.</p>
Full article ">Figure 7
<p>Representative fluorescence images of MIA PaCa-2 cells treated with MIA PaCa-2 cell-derived sEVs (<b>A</b>) or lEVs (<b>C</b>) for 1 h and 2 h in the presence of genistein (50 μg/mL), CPZ (10 μg/mL), or CytoD (10 μg/mL). Representative cells within the red boxed area are displayed in the upper panels for clarity. The cellular uptake (% of control) was measured by Fiji software (<b>B</b>,<b>D</b>) (data are presented as means ± SD, <span class="html-italic">n</span> = 3 cells in each case). Cells without pre-treatment with inhibitors were used as controls. * <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; 0.005, **** <span class="html-italic">p</span> &lt; 0.001. Scale bars: 20 μm. Images of the control groups were also used in <a href="#pharmaceutics-16-01206-f005" class="html-fig">Figure 5</a> as they were captured at the same time as those from the other groups.</p>
Full article ">Figure 7 Cont.
<p>Representative fluorescence images of MIA PaCa-2 cells treated with MIA PaCa-2 cell-derived sEVs (<b>A</b>) or lEVs (<b>C</b>) for 1 h and 2 h in the presence of genistein (50 μg/mL), CPZ (10 μg/mL), or CytoD (10 μg/mL). Representative cells within the red boxed area are displayed in the upper panels for clarity. The cellular uptake (% of control) was measured by Fiji software (<b>B</b>,<b>D</b>) (data are presented as means ± SD, <span class="html-italic">n</span> = 3 cells in each case). Cells without pre-treatment with inhibitors were used as controls. * <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; 0.005, **** <span class="html-italic">p</span> &lt; 0.001. Scale bars: 20 μm. Images of the control groups were also used in <a href="#pharmaceutics-16-01206-f005" class="html-fig">Figure 5</a> as they were captured at the same time as those from the other groups.</p>
Full article ">Figure 8
<p>Schematic representation of the pathways of cellular internalisation of EVs based on their size. This includes phagocytosis, macropinocytosis, clathrin-mediated endocytosis, and caveolin-mediated endocytosis.</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 741
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 ">
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 566
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 ">
13 pages, 2555 KiB  
Article
Trivalent Disulfide Unit-Masked System Efficiently Delivers Large Oligonucleotide
by Lei Wang, Xiao Liu, Yiliang Wu, Zhaoyan Ye, Yiru Wang, Shengshu Gao, Hao Gong and Yong Ling
Molecules 2024, 29(17), 4223; https://doi.org/10.3390/molecules29174223 - 5 Sep 2024
Viewed by 492
Abstract
Oligonucleotide drugs are shining in clinical therapeutics, but efficient and safe delivery systems severely limit their widespread use. A disulfide unit technology platform based on dynamic thiol exchange chemistry at the cell membrane has the potential for drug delivery. However, the alteration of [...] Read more.
Oligonucleotide drugs are shining in clinical therapeutics, but efficient and safe delivery systems severely limit their widespread use. A disulfide unit technology platform based on dynamic thiol exchange chemistry at the cell membrane has the potential for drug delivery. However, the alteration of the disulfide unit CSSC dihedral angle induced by different substituents directly affects the effectiveness of this technology and its stability. Previously, we constructed a trivalent low dihedral angle disulfide unit that can effectively promote the cellular uptake of small molecules. Here, we constructed a novel disulfide unit-masked oligonucleotide hybrid based on a low dihedral angle disulfide unit, motivated by prodrug design. Cellular imaging results showed that such a system exhibited superior cellular delivery efficiency than the commercial Lipo2000 without cytotoxicity. The thiol reagents significantly reduced its cellular uptake (57–74%), which proved to be endocytosis-independent. In addition, in vivo distribution experiments in mice showed that such systems can be rapidly distributed in liver tissues with a duration of action of more than 24 h, representing a potential means of silencing genes involved in the pathogenesis of liver-like diseases. In conclusion, this trivalent disulfide unit-masked system we constructed can effectively deliver large oligonucleotide drugs. Full article
(This article belongs to the Section Chemical Biology)
Show Figures

Figure 1

Figure 1
<p>Brief description of disulfide unit-masked oligonucleotide (SS-ODN) strategy used. The red marks represent disulfide unit reaction parts. The blue marks represent prodrug release parts.</p>
Full article ">Figure 2
<p>(<b>A</b>) CLSM images of HeLa S3 cells after incubation with 1 μM SS-ODN-FAM probe and Lipo2000/ODN-FAM (1 μM) complex for 4 h. Nuclei were stained with DAPI. (<b>B</b>) Quantitative analysis of fluorescence intensity. Scar bar: 8 µm. The blue marks represent DAPI staining. The green marks represent the fluorescent signal of FAM.</p>
Full article ">Figure 3
<p>(<b>A</b>) CLSM images of HeLa S3 cells after different incubation times with 1 μM SS-ODN-FAM probe. Nuclei were stained with DAPI. (<b>B</b>) Quantitative analysis of fluorescence intensity. Scar bar: 8 µm. The blue marks represent DAPI staining. The green marks represent the fluorescent signal of FAM.</p>
Full article ">Figure 4
<p>(<b>A</b>) CLSM images of HeLa S3 cells after 1 h incubation with different concentrations of SS-ODN-FAM probe. Nuclei were stained with DAPI. (<b>B</b>) Quantitative analysis of fluorescence intensity. Scar bar: 8 µm. The blue marks represent DAPI staining. The green marks represent the fluorescent signal of FAM.</p>
Full article ">Figure 5
<p>The cytotoxicity of disulfide unit system for intermediate <b>7</b> and Lipo2000 using CCK-8 assay.</p>
Full article ">Figure 6
<p>(<b>A</b>) CLSM images of HeLa S3 cells that were preincubated with thiol reagent (1.2 mM, 0.5 h), then incubated with an SS-ODN-FAM probe (1 μM, 1 h). The nuclei were stained with DAPI. (<b>B</b>) The detailed structure of different thiol reagents, such as ethylmaleimide, sodium iodoacetate, and DTNB. (<b>C</b>) Quantitative analysis of fluorescence intensity. Scar bar: 8 µm. The blue marks represent DAPI staining. The green marks represent the fluorescent signal of FAM.</p>
Full article ">Figure 7
<p>Distribution of SS-ODN-FAM probe in vivo.</p>
Full article ">Scheme 1
<p>Synthesis of disulfide unit-masked phosphoramidite monomer.</p>
Full article ">Scheme 2
<p>Synthesis of disulfide unit-conjugated oligonucleotide (SS-ODN-FAM).</p>
Full article ">
30 pages, 2340 KiB  
Review
Bio-Pathological Functions of Posttranslational Modifications of Histological Biomarkers in Breast Cancer
by Anca-Narcisa Neagu, Claudiu-Laurentiu Josan, Taniya M. Jayaweera, Hailey Morrissiey, Kaya R. Johnson and Costel C. Darie
Molecules 2024, 29(17), 4156; https://doi.org/10.3390/molecules29174156 - 2 Sep 2024
Viewed by 1104
Abstract
Proteins are the most common types of biomarkers used in breast cancer (BC) theranostics and management. By definition, a biomarker must be a relevant, objective, stable, and quantifiable biomolecule or other parameter, but proteins are known to exhibit the most variate and profound [...] Read more.
Proteins are the most common types of biomarkers used in breast cancer (BC) theranostics and management. By definition, a biomarker must be a relevant, objective, stable, and quantifiable biomolecule or other parameter, but proteins are known to exhibit the most variate and profound structural and functional variation. Thus, the proteome is highly dynamic and permanently reshaped and readapted, according to changing microenvironments, to maintain the local cell and tissue homeostasis. It is known that protein posttranslational modifications (PTMs) can affect all aspects of protein function. In this review, we focused our analysis on the different types of PTMs of histological biomarkers in BC. Thus, we analyzed the most common PTMs, including phosphorylation, acetylation, methylation, ubiquitination, SUMOylation, neddylation, palmitoylation, myristoylation, and glycosylation/sialylation/fucosylation of transcription factors, proliferation marker Ki-67, plasma membrane proteins, and histone modifications. Most of these PTMs occur in the presence of cellular stress. We emphasized that these PTMs interfere with these biomarkers maintenance, turnover and lifespan, nuclear or subcellular localization, structure and function, stabilization or inactivation, initiation or silencing of genomic and non-genomic pathways, including transcriptional activities or signaling pathways, mitosis, proteostasis, cell–cell and cell–extracellular matrix (ECM) interactions, membrane trafficking, and PPIs. Moreover, PTMs of these biomarkers orchestrate all hallmark pathways that are dysregulated in BC, playing both pro- and/or antitumoral and context-specific roles in DNA damage, repair and genomic stability, inactivation/activation of tumor-suppressor genes and oncogenes, phenotypic plasticity, epigenetic regulation of gene expression and non-mutational reprogramming, proliferative signaling, endocytosis, cell death, dysregulated TME, invasion and metastasis, including epithelial–mesenchymal/mesenchymal–epithelial transition (EMT/MET), and resistance to therapy or reversal of multidrug therapy resistance. PTMs occur in the nucleus but also at the plasma membrane and cytoplasmic level and induce biomarker translocation with opposite effects. Analysis of protein PTMs allows for the discovery and validation of new biomarkers in BC, mainly for early diagnosis, like extracellular vesicle glycosylation, which may be considered as a potential source of circulating cancer biomarkers. Full article
Show Figures

Figure 1

Figure 1
<p>PTMs of ERα in BC.</p>
Full article ">Figure 2
<p>PTMs of p53 tumor-suppressor protein in BC.</p>
Full article ">
Back to TopTop