[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (26,703)

Search Parameters:
Keywords = CD166

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 22317 KiB  
Article
Biomimetic Cooling: Functionalizing Biodegradable Chitosan Films with Saharan Silver Ant Microstructures
by Markus Zimmerl, Richard W. van Nieuwenhoven, Karin Whitmore, Wilfried Vetter and Ille C. Gebeshuber
Biomimetics 2024, 9(10), 630; https://doi.org/10.3390/biomimetics9100630 (registering DOI) - 17 Oct 2024
Abstract
The increasing occurrence of hot summer days causes stress to both humans and animals, particularly in urban areas where temperatures can remain high, even at night. Living nature offers potential solutions that require minimal energy and material costs. For instance, the Saharan silver [...] Read more.
The increasing occurrence of hot summer days causes stress to both humans and animals, particularly in urban areas where temperatures can remain high, even at night. Living nature offers potential solutions that require minimal energy and material costs. For instance, the Saharan silver ant (Cataglyphis bombycina) can endure the desert heat by means of passive radiative cooling induced by its triangular hairs. The objective of this study is to transfer the passive radiative cooling properties of the micro- and nanostructured chitin hairs of the silver ant body to technically usable, biodegradable and bio-based materials. The potential large-scale transfer of radiative cooling properties, for example, onto building exteriors such as house facades, could decrease the need for conventional cooling and, therefore, lower the energy demand. Chitosan, a chemically altered form of chitin, has a range of medical uses but can also be processed into a paper-like film. The procedure consists of dissolving chitosan in diluted acetic acid and uniformly distributing it on a flat surface. A functional structure can then be imprinted onto this film while it is drying. This study reports the successful transfer of the microstructure-based structural colors of a compact disc (CD) onto the film. Similarly, a polyvinyl siloxane imprint of the silver ant body shall make it possible to transfer cooling functionality to technically relevant surfaces. FTIR spectroscopy measurements of the reflectance of flat and structured chitosan films allow for a qualitative assessment of the infrared emissivity. A minor decrease in reflectance in a relevant wavelength range gives an indication that it is feasible to increase the emissivity and, therefore, decrease the surface temperature purely through surface-induced functionalities. Full article
(This article belongs to the Special Issue The Latest Progress in Bionics Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Incoming and outgoing radiation energy intensity and the absorption spectrum of the atmosphere. The bulk of the outgoing energy lies within the atmospheric window from 8 to 13 µm [<a href="#B17-biomimetics-09-00630" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM micrograph of cross-sections cut with a focused ion beam (FIB) through triangular chitin hairs of a Sahara silver ant gaster (hind part). Scale bar—2 µm. (<b>b</b>) Illustration of the triangular cross-section of a silver ant hair. Incoming solar radiation undergoes Mie scattering at the small indentations of the top sides. The light that enters the silver ant hair can be reflected on the bottom side when the conditions for total reflection are met (incidence angle and difference in refractive index between silver ant hair and air gap) [<a href="#B14-biomimetics-09-00630" class="html-bibr">14</a>]. Scale bar—1 µm.</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM sample holder with exposed and unexposed shrimp shell sample, as well as silver ant gaster (rear segment of the silver ant). Scale bar—1 cm. (<b>b</b>) Climate chamber cycles (programmed).</p>
Full article ">Figure 4
<p>Process of creating a copy of the silver ant surface structure in chitosan with the help of a PVS stamp.</p>
Full article ">Figure 5
<p>(<b>a</b>) Confocal image of scratched shrimp shell before and after exposure in the climate chamber. Scale bar—200 µm. (<b>b</b>) Chitosan film with iridescent microstructures transferred from a CD. Scale bar—0.5 cm.</p>
Full article ">Figure 6
<p>Confocal images of (<b>a</b>) an ’unstructured’ area of the PVS stamp, which shows the structure of the cardboard that surrounded the silver ant gaster. (<b>b</b>) The PVS–cardboard structure transferred onto chitosan. (<b>c</b>) A structured area of the PVS stamp structured with an ant gaster. (<b>d</b>) The PVS–ant structure transferred onto chitosan. Scale bars—200 µm.</p>
Full article ">Figure 7
<p>(<b>a</b>) Average reflectance (* in relation to a reference gold mirror) of structured and unstructured chitosan films. The structured areas in both samples exhibit a slightly higher reflectance for wavelengths greater than 6 µm. However, this difference is less than the calculated standard deviation. Below 6 µm, the two samples feature inconsistent differences in reflectance. (<b>b</b>) Zoom into the respective region of the atmospheric window in (<b>a</b>).</p>
Full article ">
14 pages, 3363 KiB  
Article
Activation-Induced Marker Assay to Identify and Isolate HCV-Specific T Cells for Single-Cell RNA-Seq Analysis
by Mohamed Eisa, Nicol Flores, Omar Khedr, Elsa Gomez-Escobar, Nathalie Bédard, Nourtan F. Abdeltawab, Julie Bruneau, Arash Grakoui and Naglaa H. Shoukry
Viruses 2024, 16(10), 1623; https://doi.org/10.3390/v16101623 (registering DOI) - 17 Oct 2024
Abstract
Identification and isolation of antigen-specific T cells for downstream transcriptomic analysis is key for various immunological studies. Traditional methods using major histocompatibility complex (MHC) multimers are limited by the number of predefined immunodominant epitopes and MHC matching of the study subjects. Activation-induced markers [...] Read more.
Identification and isolation of antigen-specific T cells for downstream transcriptomic analysis is key for various immunological studies. Traditional methods using major histocompatibility complex (MHC) multimers are limited by the number of predefined immunodominant epitopes and MHC matching of the study subjects. Activation-induced markers (AIM) enable highly sensitive detection of rare antigen-specific T cells irrespective of the availability of MHC multimers. Herein, we have developed an AIM assay for the detection, sorting and subsequent single-cell RNA sequencing (scRNA-seq) analysis of hepatitis C virus (HCV)-specific T cells. We examined different combinations of the activation markers CD69, CD40L, OX40, and 4-1BB at 6, 9, 18 and 24 h post stimulation with HCV peptide pools. AIM+ CD4 T cells exhibited upregulation of CD69 and CD40L as early as 6 h post-stimulation, while OX40 and 4-1BB expression was delayed until 18 h. AIM+ CD8 T cells were characterized by the coexpression of CD69 and 4-1BB at 18 h, while the expression of CD40L and OX40 remained low throughout the stimulation period. AIM+ CD4 and CD8 T cells were successfully sorted and processed for scRNA-seq analysis examining gene expression and T cell receptor (TCR) usage. scRNA-seq analysis from this one subject revealed that AIM+ CD4 T (CD69+ CD40L+) cells predominantly represented Tfh, Th1, and Th17 profiles, whereas AIM+ CD8 T (CD69+ 4-1BB+) cells primarily exhibited effector and effector memory profiles. TCR analysis identified 1023 and 160 unique clonotypes within AIM+ CD4 and CD8 T cells, respectively. In conclusion, this approach offers highly sensitive detection of HCV-specific T cells that can be applied for cohort studies, thus facilitating the identification of specific gene signatures associated with infection outcome and vaccination. Full article
(This article belongs to the Special Issue Hepatitis Viral Infections, Pathogenesis and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Expression of CD69, CD40L and OX40 at 18 h defines AIM<sup>+</sup> CD4 T cells. (<b>A</b>) Representative gating strategy used for the detection of AIM<sup>+</sup> cells among total CD4<sup>+</sup> and CD8<sup>+</sup> T cells. (<b>B</b>–<b>D</b>) Representative staining for different AIM marker pairs. PBMC from a resolver of HCV reinfection were stimulated with either: (<b>B</b>) no peptide (unstimulated control), (<b>C</b>) staphylococcal enterotoxin B (SEB) (positive control), or (<b>D</b>) HCV non-structural protein (NS5B; aa 2708–3014) peptides pool for 6, 9, 18, or 24 h followed by the detection of AIM markers upregulation in CD4<sup>+</sup> T cells using flow cytometry. AIM<sup>+</sup> CD4<sup>+</sup> T cells are defined by the expression of CD69 in conjunction with either CD40L, OX40, or 4.1BB and presented as a percentage of total CD4<sup>+</sup> T cells. (<b>E</b>) Bar charts depicting the frequencies of AIM<sup>+</sup> CD4<sup>+</sup> T cells at the indicated time points post-stimulation. Data points represent different experimental replicates and bars represent median with interquartile range after background staining subtraction from unstimulated samples. Statistical analysis was performed using a two-way ANOVA followed by Tukey’s multiple-comparison posttest. * = <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>AIM<sup>+</sup> CD8<sup>+</sup> T cells primarily express CD69 and 4-1BB at 18 h post-activation. (<b>A</b>–<b>C</b>) Representative staining demonstrating the expression of various AIM marker pairs by CD8<sup>+</sup> T cells. PBMC from a resolver of HCV reinfection were stimulated with either: (<b>A</b>) no peptide (unstimulated control), (<b>B</b>) staphylococcal enterotoxin B (SEB) (positive control), or (<b>C</b>) HCV non-structural protein (NS5B; aa 2708–3014) peptides pool for 6, 9, 18, or 24 h followed by the detection of AIM markers upregulation in CD8<sup>+</sup> T cells using flow cytometry. Cells were gated as in <a href="#viruses-16-01623-f001" class="html-fig">Figure 1</a>A. AIM<sup>+</sup> CD8<sup>+</sup> T cells are defined by the coexpression of CD69 and 4.1BB and presented as a percentage of total CD8<sup>+</sup> T cells. (<b>D</b>) Bar charts depicting the frequencies of AIM<sup>+</sup> CD8<sup>+</sup> T cells at the indicated time points post-stimulation. Data points represent different experimental replicates and bars represent median with interquartile range after background staining subtraction from unstimulated samples. Two-way ANOVA followed by Tukey’s multiple-comparison posttest. * = <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Gene expression analysis of AIM<sup>+</sup> CD4 T cells. (<b>A</b>) Uniform Manifold Approximation and Projection (UMAP) plot of scRNA-seq data from AIM<sup>+</sup> CD4 T cells that FACS sorted from an HCV-resolver subject. (<b>B</b>) Heatmap showing the expression of the top marker genes of each cluster. (<b>C</b>) Feature plots displaying the expression of representative markers of different CD4 subsets. (<b>D</b>) Dot plot presenting the expression of canonical CD4 markers in each cluster. Color represents the average expression and dot size represents the percentage of cells that express the gene.</p>
Full article ">Figure 4
<p>Gene expression analysis of AIM<sup>+</sup> CD8 T cells. (<b>A</b>) Uniform Manifold Approximation and Projection (UMAP) plot of scRNA-seq data from AIM<sup>+</sup> CD8 T cells derived from an HCV-infected subject. (<b>B</b>) Heatmap showing the expression of the top marker genes of each cluster. (<b>C</b>) Feature plots displaying the expression of representative markers of different CD8 subsets. (<b>D</b>) Dot plot presenting the expression of canonical CD8 markers in each cluster. Color represents the average expression and dot size represents the percentage of cells that express the gene. Eff: Effector, EM: Effector memory, Exh: Exhausted, Exh-Pre: Precursor exhausted, Exh-Term: Terminally exhausted, ISG: Interferon-stimulated genes, TEMRA: Terminally differentiated EM.</p>
Full article ">Figure 5
<p>T cell receptor analysis of AIM<sup>+</sup> CD4 and CD8 T cells. (<b>A</b>,<b>B</b>) Uniform Manifold Approximation and Projection (UMAP) plots and bar charts showing the relative frequency of TCR clonotypes within the indicated categories in (<b>A</b>) AIM<sup>+</sup> CD4 T cells and (<b>B</b>) AIM<sup>+</sup> CD8 T cells. (<b>C</b>,<b>D</b>) Bar charts depicting the usage of (<b>C</b>) T-cell receptor alpha variable region (TRAV) and (<b>D</b>) T-cell receptor beta variable region (TRBV) genes in AIM<sup>+</sup> CD4 T cells (<b>left</b>) and AIM<sup>+</sup> CD8 T cells (<b>right</b>).</p>
Full article ">
13 pages, 889 KiB  
Article
Antiretroviral Therapy with Ritonavir-Boosted Atazanavir- and Lopinavir-Containing Regimens Correlates with Diminished HIV-1 Neutralization
by Eloisa Yuste, Horacio Gil, Felipe Garcia and Victor Sanchez-Merino
Vaccines 2024, 12(10), 1176; https://doi.org/10.3390/vaccines12101176 (registering DOI) - 17 Oct 2024
Abstract
Background/Objectives: The impact of virion maturation on neutralizing antibody responses in HIV treatment is not fully understood. This study examines whether antiretroviral regimens (ART) with boosted protease inhibitors (b-PI), which increase exposure to immature virions, affect neutralization capacity compared to Non-b-PI regimens. Methods: [...] Read more.
Background/Objectives: The impact of virion maturation on neutralizing antibody responses in HIV treatment is not fully understood. This study examines whether antiretroviral regimens (ART) with boosted protease inhibitors (b-PI), which increase exposure to immature virions, affect neutralization capacity compared to Non-b-PI regimens. Methods: Neutralization activity was assessed in 45 HIV-infected individuals on b-PI regimens and 56 on Non-b-PI regimens, adjusting for factors like infection duration, ART initiation, and immune markers. Individuals on b-PI regimens had significantly lower neutralization scores [mean: 6.1, 95% Confidence Interval (CI): 5.3–6.9] than those on Non-b-PI regimens (mean: 8.9, 95% CI: 8.0–9.9; p < 0.0001). This difference was not explained by infection duration or CD4+ counts. CD4+/CD8+ ratios were positively associated with neutralization, while b-PI use was negatively associated. A regression model indicated that b-PI use significantly predicted lower neutralization scores (beta = −0.30, p = 0.049). Conclusions: These findings suggest that exposure to immature virions via b-PI use reduces neutralizing antibody responses, highlighting the importance of virion maturation in antibody induction. ART regimens promoting exposure to mature virions may enhance neutralization, with potential implications for HIV vaccine design. Further research is needed to explore implications for HIV vaccine design, especially using virus-like particles. Full article
Show Figures

Figure 1

Figure 1
<p>Individual selection and stratification. Serum-purified IgGs from 364 individuals were included. b-PI: ART regimen including boosted protease inhibitor (protease inhibitor + ritonavir). Non-b-PI: ART regimen not including boosted protease inhibitor. NRTI: nucleoside analog reverse-transcriptase inhibitor; NNRTI: non-nucleoside analog reverse-transcriptase inhibitor; PI: protease inhibitor; and INSTI: integrase strand transfer inhibitor.</p>
Full article ">Figure 2
<p>Comparison of neutralizing antibody responses (neutralization scores) in individuals treated with and without boosted protease inhibitors. HIV-1 infected individuals were on treatment regimens that either included boosted protease inhibitors (b-PI) or did not (Non-b-PI). Horizontal bars in the box plots indicate the mean for each group, and the standard errors of the means (SEMs) are shown.</p>
Full article ">Figure 3
<p>Comparison of cumulative time post-infection, time to ART initiation, CD4+ and CD8+ T-cell counts, CD4+/CD8+ T-lymphocyte ratios, and nadir CD4+ T-lymphocyte counts between individuals treated with and without b-PI. (<b>a</b>) Time post-infection comparison. (<b>b</b>) Time to ART initiation comparison. (<b>c</b>) CD4+ T-cell count comparison. (<b>d</b>) CD8+ T-cell count comparison. (<b>e</b>) CD4+/CD8+ T-lymphocyte ratio comparison. (<b>f</b>) Nadir CD4+ T-lymphocyte count comparison. Horizontal bars within the box plots indicate the mean for each group, and standard errors of means (SEMs) are shown. Significance levels between groups are indicated. Mann–Whitney U tests were used for comparisons between groups. Simple comparisons were made with two-sided alpha level of 0.05.</p>
Full article ">
16 pages, 717 KiB  
Article
Optimization of Desalting Conditions for the Green Seaweed Codium fragile for Use as a Functional Food with Hypnotic Effects
by Sohong Park, Duhyeon Kim, Seonghui Kim, Gibeom Choi, Hodeung Yoo, Serim Park and Suengmok Cho
Foods 2024, 13(20), 3287; https://doi.org/10.3390/foods13203287 (registering DOI) - 16 Oct 2024
Abstract
Codium fragile (CF) contains various bioactive compounds, but its high salt content (39.8%) makes its use as a functional food challenging. Here, we aimed to optimize the desalination process and verify changes in functionality based on variations in salt and total phenolic contents. [...] Read more.
Codium fragile (CF) contains various bioactive compounds, but its high salt content (39.8%) makes its use as a functional food challenging. Here, we aimed to optimize the desalination process and verify changes in functionality based on variations in salt and total phenolic contents. To optimize the CF immersion conditions for the lowest salt content and monitor the total phenolic content, a response surface methodology was used. The optimal immersion conditions were as follows: X1 (immersion temperature) = 42.8 °C; X2 (immersion time) = 1.0 h. An inverse correlation was noted between salt content and total phenolic content. Among the post-desalination processes, desalination with centrifugal dehydration (CD) significantly reduced salt content. CD ethanol extract (CD-E) induced the longest sleep duration in the pentobarbital-induced sleep test in ethanol extracts. Moreover, 1000 mg/kg CD-E had a significant effect on non-rapid eye movement sleep but did not affect delta activity. These findings highlight the potential of industrializing CF as a functional food through desalination and its promise as a natural aid for sleep promotion. Full article
(This article belongs to the Section Nutraceuticals, Functional Foods, and Novel Foods)
15 pages, 24149 KiB  
Communication
Neutrophil Extracellular Traps in Pediatric Inflammatory Bowel Disease: A Potential Role in Ulcerative Colitis
by Rachel Shukrun, Victoria Fidel, Szilvia Baron, Noga Unger, Yoav Ben-Shahar, Shlomi Cohen, Ronit Elhasid and Anat Yerushalmy-Feler
Int. J. Mol. Sci. 2024, 25(20), 11126; https://doi.org/10.3390/ijms252011126 - 16 Oct 2024
Abstract
Inflammatory bowel disease (IBD), encompassing Crohn’s disease (CD) and ulcerative colitis (UC), is a chronic inflammatory condition of the gut affecting both adults and children. Neutrophil extracellular traps (NETs) are structures released by activated neutrophils, potentially contributing to tissue damage in various diseases. [...] Read more.
Inflammatory bowel disease (IBD), encompassing Crohn’s disease (CD) and ulcerative colitis (UC), is a chronic inflammatory condition of the gut affecting both adults and children. Neutrophil extracellular traps (NETs) are structures released by activated neutrophils, potentially contributing to tissue damage in various diseases. This study aimed to explore the presence and role of NETs in pediatric IBD. We compared intestinal biopsies and peripheral blood from 20 pediatric IBD patients (UC and CD) to controls. Biopsy staining and techniques for neutrophil activation were used to assess neutrophil infiltration and NET formation. We also measured the enzymatic activity of key NET proteins and evaluated NET formation in UC patients in remission. Both UC and CD biopsies showed significantly higher levels of neutrophils and NETs compared to controls (p < 0.01), with UC exhibiting the strongest association. Peripheral blood neutrophils from UC patients at diagnosis displayed increased NET formation compared to controls and CD patients. Interestingly, NET formation normalized in UC patients following remission-inducing treatment. This pilot study suggests a potential role for NETs in pediatric IBD, particularly UC. These findings warrant further investigation into the mechanisms of NET involvement and the potential for targeting NET formation as a therapeutic strategy. Full article
Show Figures

Figure 1

Figure 1
<p>Elevated neutrophils and NETs in pediatric IBD tissue samples. (<b>A</b>) The number of neutrophils and (<b>B</b>) NET formation was significantly elevated in both UC (N = 9) and CD (N = 11) compared to controls (N = 9) (*** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>) Representative images of NETs in the tissue samples of patients with UC and CD. The tissues were stained with NE (green) and H3 (red). NET-forming cells are identified by their high H3 signal and NE co-localization (yellow). Scale bar: 50 μm. On the right bottom panel, there is a higher magnification of the selected area (marked by dashed lines), scale bar: 20 μm (right).</p>
Full article ">Figure 2
<p>Neutrophil infiltration and NET formation in inflamed vs. non-inflamed intestinal tissue. (<b>A</b>,<b>B</b>) Significantly higher levels of infiltrating neutrophils were demonstrated in the inflamed intestine of UC (<b>A</b>) and CD (<b>B</b>) compared to the non-inflamed intestine (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>,<b>D</b>) Significantly higher levels of NET formation were demonstrated in the inflamed intestine of UC (<b>C</b>) and CD (<b>D</b>) compared to the non-inflamed tissue (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>NET formation in inflamed vs. non-inflamed intestinal tissue. Representative images of NETs in the inflamed intestine of UC and CD, compared to the non-inflamed tissue. Tissues were stained with NE (green) and H3 (red). NET-forming cells are identified by their high H3 signal and NE co-localization (yellow). Scale bar: 50 μm. On the right bottom panel, there is a higher magnification of the selected area (marked by dashed lines), scale bar: 20 μm (right).</p>
Full article ">Figure 4
<p>Increased NET formation by neutrophils from pediatric UC patients. (<b>A</b>) Neutrophils isolated from patients with UC at diagnosis displayed significantly elevated NET formation upon stimulation with PMA compared to both healthy controls and CD patients (** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05, respectively). (<b>B</b>,<b>C</b>) Enzymatic activity of neutrophil elastase (NE) and myeloperoxidase (MPO) in UC and CD patients at diagnosis showed no significant difference compared to controls. (<b>D</b>) Representative images of NET formation in UC at diagnosis. Cells were stained with Sytox Green (green) and Hoechst 33342 (blue) dyes to identify decondensed chromatin. White arrows indicate NET-forming cells. Scale bar: 50 μm.</p>
Full article ">Figure 5
<p>Normalization of NET formation in UC with immunomodulatory therapy. (<b>A</b>) Neutrophils from UC patients in remission displayed significantly lower NET formation upon stimulation compared to UC patients at diagnosis (** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05, respectively). NET formation in remission was comparable to healthy controls (<span class="html-italic">p</span> = 0.962, ns—not significant). (<b>B</b>,<b>C</b>) Enzymatic activity of neutrophil elastase (NE) and myeloperoxidase (MPO) showed no significant difference between UC patients in remission, at diagnosis, and healthy controls. (<b>D</b>) Representative images confirm lower NET formation in UC patients in remission compared to diagnosis. Cells were stained with Sytox Green (green) and Hoechst 33342 (blue) dyes to identify decondensed chromatin. White arrows indicate NET-forming cells. Scale bar: 50 μm.</p>
Full article ">
12 pages, 2239 KiB  
Article
Some Glycoproteins Expressed on the Surface of Immune Cells and Cytokine Plasma Levels Can Be Used as Potential Biomarkers in Patients with Colorectal Cancer
by Tsvetelina Batsalova, Denitsa Uzunova, Gergana Chavdarova, Tatyana Apostolova and Balik Dzhambazov
Biomolecules 2024, 14(10), 1314; https://doi.org/10.3390/biom14101314 - 16 Oct 2024
Abstract
Colorectal cancer (CRC) is a leading cause of mortality worldwide. Its incidence holds a major position among the most common life-threatening diseases. Hence, the early identification and precise characterization of disease activity based on proper biomarkers are of utmost importance for therapeutic strategy [...] Read more.
Colorectal cancer (CRC) is a leading cause of mortality worldwide. Its incidence holds a major position among the most common life-threatening diseases. Hence, the early identification and precise characterization of disease activity based on proper biomarkers are of utmost importance for therapeutic strategy and patient survival. The identification of new biomarkers for colorectal cancer or disease-specific levels/combinations of biomarkers will significantly contribute to precise diagnosis and improved personalized treatment of patients. Therefore, the present study aims to identify colorectal cancer-specific immunological biomarkers. The plasma levels of several cytokines (interleukin-1β /IL-1β/, IL-2, IL-4, IL-10, IL-12, IL-15, TGFβ and IFNγ) of 20 patients with colorectal cancer and 21 healthy individuals were determined by ELISA. The expression of several types of glycoproteins on the surface of peripheral blood leukocytes isolated from CRC patients and healthy volunteers was evaluated by flow cytometry. Correlations between cytokine levels and cell surface glycoprotein expression were analyzed. The obtained results demonstrated significantly elevated levels of CD80, CD86, CD279 and CD274 expressing leukocyte populations in the cancer patient group, while the numbers of NK cells and CD8- and CD25-positive cells were decreased. Based on these data and the correlations with cytokine levels, it can be concluded that CD25, CD80, CD86, CD274 and CD279 glycoproteins combined with specific plasma levels of IL-1β, IL-2, IL-15 and TGFβ could represent potential biomarkers for colorectal cancer. Full article
(This article belongs to the Special Issue Immune-Related Biomarkers: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Plasma concentrations of inflammatory cytokines in stage IV CRC patients and control healthy volunteers. Levels of IL-1β (<b>A</b>), IL-2 (<b>B</b>), IL-4 (<b>C</b>), IL-10 (<b>D</b>), IL-12 (<b>E</b>), IL-13 (<b>F</b>), IL-15 (<b>G</b>), TGFβ (<b>H</b>) and IFNγ (<b>I</b>). Data represent ± standard error of the mean (±SEM). A Mann–Whitney <span class="html-italic">U</span> test was used for statistical analyses. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Flow cytometry analyses of the expression of T-cell-specific markers on peripheral blood leukocytes derived from stage IV CRC patients and healthy individuals. CD4<sup>+</sup> T cells (<b>A</b>), CD8<sup>+</sup> T cells (<b>B</b>), CD25<sup>+</sup> T cells (<b>C</b>), CD28<sup>+</sup> T cells (<b>D</b>), CD152<sup>+</sup> T cells (<b>E</b>), and CD279<sup>+</sup> T cells (<b>F</b>). The results are presented as ±SEM. Statistical significance was defined by the Mann–Whitney <span class="html-italic">U</span> test. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Levels of CD20<sup>+</sup> B cells, NK cells and expression of some B7 family molecules in the peripheral blood of stage IV CRC patients and healthy controls. Percentage of CD20<sup>+</sup> B cells (<b>A</b>), CD80<sup>+</sup> B cells (<b>B</b>), CD86<sup>+</sup> B cells (<b>C</b>), CD273<sup>+</sup> lymphocytes (<b>D</b>), CD274<sup>+</sup> lymphocytes (<b>E</b>), and CD56<sup>+</sup>/CD16<sup>+</sup> NK cells (<b>F</b>). Data are shown as ±SEM. Statistical analyses were performed using the Mann–Whitney <span class="html-italic">U</span> test; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Correlations between IL-2 and IL-15 plasma concentrations with other cytokines. First column: IL-2 vs. IL-1β (<b>A</b>), IL-2 vs. TGFβ (<b>C</b>), IL-2 vs. IFNγ (<b>E</b>), IL-2 vs. IL-15 (<b>G</b>), IL-2 vs. IL-12 (<b>I</b>); Second column: IL-15 vs. IL-12 (<b>B</b>), IL-15 vs. IL-1β (<b>D</b>), IL-15 vs. IL-13 (<b>F</b>), IL-15 vs. TGFβ (<b>H</b>), and IL-2 vs. IL-12 only for CRC patients (<b>J</b>). The charts indicate Spearman’s rank correlation coefficient (ρ). * <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>Correlations between circulating cytokines and immunomodulatory molecule expression levels determined for the stage IV CRC patient group. IL-15 vs. CD279<sup>+</sup> cells (<b>A</b>), IL-1β vs. CD86<sup>+</sup> cells (<b>B</b>) and CD8<sup>+</sup> vs. CD279<sup>+</sup> cells (<b>C</b>). The graphs show Spearman’s rank correlation coefficient (ρ). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
16 pages, 2591 KiB  
Article
Short Link N Modulates Inflammasome Activity in Intervertebral Discs Through Interaction with CD14
by Muskan Alad, Michael P. Grant, Laura M. Epure, Sunny Y. Shih, Geraldine Merle, Hee-Jeong Im, John Antoniou and Fackson Mwale
Biomolecules 2024, 14(10), 1312; https://doi.org/10.3390/biom14101312 - 16 Oct 2024
Abstract
Intervertebral disc degeneration and pain are associated with the nucleotide-binding domain, leucine-rich repeat, and pyrin domain-containing 3 (NLRP3) inflammasome activation and the processing of interleukin-1 beta (IL-1β). Activation of thehm inflammasome is triggered by Toll-like receptor stimulation and requires the cofactor receptor cluster [...] Read more.
Intervertebral disc degeneration and pain are associated with the nucleotide-binding domain, leucine-rich repeat, and pyrin domain-containing 3 (NLRP3) inflammasome activation and the processing of interleukin-1 beta (IL-1β). Activation of thehm inflammasome is triggered by Toll-like receptor stimulation and requires the cofactor receptor cluster of differentiation 14 (CD14). Short Link N (sLN), a peptide derived from link protein, has been shown to modulate inflammation and pain in discs in vitro and in vivo; however, the underlying mechanisms remain elusive. This study aims to assess whether sLN modulates IL-1β and inflammasome activity through interaction with CD14. Disc cells treated with lipopolysaccharides (LPS) with or without sLN were used to assess changes in Caspase-1, IL-1β, and phosphorylated nuclear factor kappa-light-chain-enhancer of activated B cells (NFκB). Peptide docking of sLN to CD14 and immunoprecipitation were performed to determine their interaction. The results indicated that sLN inhibited LPS-induced NFκB and Caspase-1 activation, reducing IL-1β maturation and secretion in disc cells. A significant decrease in inflammasome markers was observed with sLN treatment. Immunoprecipitation studies revealed a direct interaction between sLN and the LPS-binding pocket of CD14. Our results suggest that sLN could be a potential therapeutic agent for discogenic pain by mitigating IL-1β and inflammasome activity within discs. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Immunohistochemistry of NLRP3 expression in a degenerative IVD model. Rabbit IVDs from a puncture model were given a single injection of either saline or sLN. Sham animals were used as controls. Twelve weeks following injection, IVDs were processed for immunohistochemistry and the detection of NLRP3. Higher-magnification inset images are included and demarcated to highlight specific areas of interest. Immunohistochemistry of NLRP3 in (<b>A</b>) sham and degenerative IVDS treated with (<b>B</b>) saline or (<b>C</b>) sLN. (<b>D</b>) Densitometry of NLRP3 expression in IVDs presented in (<b>A</b>). Statistical significance was assessed using Student’s <span class="html-italic">t</span>-test (comparison between saline and sLN); ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 4. (IVD—intervertebral disc, NLRP3—nucleotide-binding domain, leucine-rich repeat, and pyrin domain-containing 3, yellow dotted line—endplate).</p>
Full article ">Figure 2
<p>Evaluation of NF-κB activation in human nucleus pulposus cells treated with sLN and LPS. (<b>A</b>) Human nucleus pulposus (NP) cells were treated with LPS alone or in combination with sLN with the indicated concentrations (0.05, 0.5, or 5.0 μg/mL) for 45 min. Western blotting was performed to detect activation of P-NF-κB. Blots were normalized to GAPDH for loading by densitometry and calculated as a fold increase over control. The densitometry of the blots is presented in the graph below. (<b>B</b>) HNP cells were incubated with LPS alone or (<b>C</b>) LPS and 0.5 μg/mL sLN for the indicated times (0–180 min). P-NF-κB signal was normalized to GAPDH and calculated as a fold-over control. Statistical significance was assessed using ANOVA and post hoc Dunnett’s test (comparison to control); ****, <span class="html-italic">p</span> &lt; 0.0001; <span class="html-italic">n</span> = 4. Original Western blot images are available in <a href="#app1-biomolecules-14-01312" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 3
<p>Suppression of inflammasome markers by sLN. HNP cells were incubated for 48 h in medium supplemented with LPS or LPS and 0.5 μg/mL sLN. Control samples were incubated with medium alone. RNA expression is shown for (<b>A</b>) <span class="html-italic">NLRP3</span>, (<b>B</b>) <span class="html-italic">PYC</span>, (<b>C</b>) <span class="html-italic">CASP1</span>, (<b>D</b>) <span class="html-italic">IL1B</span>, and (<b>E</b>) <span class="html-italic">TNFA</span>. Treatments were compared to controls. ANOVA and post hoc Dunnett’s test; *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001; ****, <span class="html-italic">p</span> &lt; 0.0001; <span class="html-italic">n</span> = 4.</p>
Full article ">Figure 4
<p>Assessment of Caspase-1 activation and IL-1β secretion in hNP cells. HNP cells were treated with LPS alone or in combination with sLN at varying concentrations (0.05, 0.5 and 5.0 μg/mL) for 48 h. (<b>A</b>) Western blotting and densitometry of Caspase-1 demonstrating Pro-Caspase-1, Caspase-1, and GAPDH as loading control. (<b>B</b>) Western blotting and densitometry of pro- and mature forms of IL-1β. Data are expressed as fold-over controls. Statistical significance was assessed using ANOVA and post hoc Dunnett’s test (treatments compared to LPS alone); **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 4. Original Western blot images are available in <a href="#app1-biomolecules-14-01312" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 5
<p>Inflammasome activation and hNP-induced macrophage polarization. RAW macrophages and hNP cells were co-cultured to determine the effects of hNP on macrophage polarization. HNP cells were stimulated with LPS and incubated for 48 hrs with RAW cells. Expression of M1 (<b>A</b>–<b>C</b>) and M2 (<b>D</b>–<b>F</b>) markers in RAW macrophages were measured by qPCR. Plots represent fold-over control. ANOVA and post hoc Dunnett’s test (treatments compared to control); *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 4.</p>
Full article ">Figure 6
<p>Interaction and immunoprecipitation of sLN and CD14. (<b>A</b>) Schematic on the co-immunoprecipitation of CD14 and sLN and dot-blot showing enrichment of sLN following pull-down with CD14. (<b>B</b>) Schematic demonstrating the competitive co-immunoprecipitation of biotinylated LPS for CD14 in the presence of unlabeled LPS and sLN. Western blot showing the detection of CD14 following the indicated pull-downs. Lanes from left to right: control (CD14 only), LPS-treated, sLN-treated. Original Western blot images are available in <a href="#app1-biomolecules-14-01312" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 7
<p>Mechanism of sLN inhibition of inflammasome activation in disc cells. This schematic illustrates the process by which sLN inhibits inflammasome activation in nucleus pulposus cells, detailing the interactions and molecular pathways involved.</p>
Full article ">
25 pages, 638 KiB  
Review
Elastography as a Discriminator Between Fibrotic and Inflammatory Strictures in Crohn’s Disease: A Dead End or Bright Future in Clinical Decision-Making? Critical Review
by Maryla Kuczyńska, Monika Zbroja and Anna Drelich-Zbroja
Diagnostics 2024, 14(20), 2299; https://doi.org/10.3390/diagnostics14202299 - 16 Oct 2024
Abstract
Background: Crohn’s disease (CD) is a complex systemic entity, characterized by the progressive and relapsing inflammatory involvement of any part of the gastrointestinal tract. Its clinical pattern may be categorized as penetrating, stricturing or non-penetrating non-stricturing. Methods: In this paper, we performed a [...] Read more.
Background: Crohn’s disease (CD) is a complex systemic entity, characterized by the progressive and relapsing inflammatory involvement of any part of the gastrointestinal tract. Its clinical pattern may be categorized as penetrating, stricturing or non-penetrating non-stricturing. Methods: In this paper, we performed a database search (Pubmed, MEDLINE, Mendeley) using combinations of the queries “crohn”, “stricture” and “elastography” up to 19 June 2024 to summarize current knowledge regarding the diagnostic utility of ultrasound (US) and magnetic resonance (MR) elastography techniques in the evaluation of stricturing CD by means of an assessment of the transmural intestinal fibrosis. We decided to include papers published since 1 January 2017 for further evaluation (n = 24). Results: Despite growing collective and original data regarding numerous applications of mostly ultrasound elastography (quantification of fibrosis, distinguishing inflammatory from predominantly fibrotic strictures, assessment of treatment response, predicting disease progression) constantly emerging, to date, we are still lacking a uniformization in both cut-off values and principles of measurements, i.e., reference tissue in strain elastography (mesenteric fat, abdominal muscles, unaffected bowel segment), units, not to mention subtle differences in technical background of SWE techniques utilized by different vendors. All these factors imply that ultrasound elastography techniques are hardly translatable throughout different medical centers and practitioners, largely depending on the local experience. Conclusions: Nonetheless, the existing medical evidence is promising, especially in terms of possible longitudinal comparative studies (follow-up) of patients in the course of the disease, which seems to be of particular interest in children (lack of radiation, less invasive contrast media) and terminal ileal disease (easily accessible). Full article
(This article belongs to the Special Issue Advances in Ultrasound)
24 pages, 10877 KiB  
Article
Modeling Temperature-Dependent Photoluminescence Dynamics of Colloidal CdS Quantum Dots Using Long Short-Term Memory (LSTM) Networks
by Ivan Malashin, Daniil Daibagya, Vadim Tynchenko, Vladimir Nelyub, Aleksei Borodulin, Andrei Gantimurov, Alexandr Selyukov, Sergey Ambrozevich, Mikhail Smirnov and Oleg Ovchinnikov
Materials 2024, 17(20), 5056; https://doi.org/10.3390/ma17205056 - 16 Oct 2024
Abstract
This study addresses the challenge of modeling temperature-dependent photoluminescence (PL) in CdS colloidal quantum dots (QD), where PL properties fluctuate with temperature, complicating traditional modeling approaches. The objective is to develop a predictive model capable of accurately capturing these variations using Long Short-Term [...] Read more.
This study addresses the challenge of modeling temperature-dependent photoluminescence (PL) in CdS colloidal quantum dots (QD), where PL properties fluctuate with temperature, complicating traditional modeling approaches. The objective is to develop a predictive model capable of accurately capturing these variations using Long Short-Term Memory (LSTM) networks, which are well suited for managing temporal dependencies in time-series data. The methodology involved training the LSTM model on experimental time-series data of PL intensity and temperature. Through numerical simulation, the model’s performance was assessed. Results demonstrated that the LSTM-based model effectively predicted PL trends under different temperature conditions. This approach could be applied in optoelectronics and quantum dot-based sensors for enhanced forecasting capabilities. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) TEM image of the synthesized semiconductor colloidal CdS QD. (<b>b</b>) Absorption (dashed curve) and luminescence (solid curve) spectra of CdS QD.</p>
Full article ">Figure 2
<p>Workflow of the experimental and analytical approach used in this study.</p>
Full article ">Figure 3
<p>Examples of images of spectra with approximations at different temperatures. (<b>a</b>) 85.6 K, (<b>b</b>) 182.6 K, (<b>c</b>) 300.2 K.</p>
Full article ">Figure 4
<p>The PL spectra of CdS quantum dots (QDs) were measured as the sample was heated from 84 K to 306 K, with curves recorded at intervals of 7 K.</p>
Full article ">Figure 5
<p>Temperature-dependence of trap state luminescence peak energy (<b>a</b>), FWHM (<b>b</b>), and integrated intensity (<b>c</b>) for CdS QDs. The approximation of the absorption spectrum by four Gaussian functions (G1–G4); the red dashed line shows the fitting result (<b>d</b>).</p>
Full article ">Figure 6
<p>LSTM prediction results.</p>
Full article ">Figure 7
<p>Overview of the LSTM model for predicting temperature-dependent PL of CdS QDs.</p>
Full article ">
17 pages, 4221 KiB  
Article
Prognostic Impact of Acute and Chronic Inflammatory Interleukin Signatures in the Tumor Microenvironment of Early Breast Cancer
by Anne-Sophie Heimes, Ina Shehaj, Katrin Almstedt, Slavomir Krajnak, Roxana Schwab, Kathrin Stewen, Antje Lebrecht, Walburgis Brenner, Annette Hasenburg and Marcus Schmidt
Int. J. Mol. Sci. 2024, 25(20), 11114; https://doi.org/10.3390/ijms252011114 - 16 Oct 2024
Abstract
Interleukins play dual roles in breast cancer, acting as both promoters and inhibitors of tumorigenesis within the tumor microenvironment, shaped by their inflammatory functions. This study analyzed the subtype-specific prognostic significance of an acute inflammatory versus a chronic inflammatory interleukin signature using microarray-based [...] Read more.
Interleukins play dual roles in breast cancer, acting as both promoters and inhibitors of tumorigenesis within the tumor microenvironment, shaped by their inflammatory functions. This study analyzed the subtype-specific prognostic significance of an acute inflammatory versus a chronic inflammatory interleukin signature using microarray-based gene expression analysis. Correlations between these interleukin signatures and immune cell markers (CD8, IgKC, and CD20) and immune checkpoints (PD-1) were also evaluated. This study investigated the prognostic significance of an acute inflammatory IL signature (IL-12, IL-21, and IFN-γ) and a chronic inflammatory IL signature (IL-4, IL-5, IL-10, IL-13, IL-17, and CXCL1) for metastasis-free survival (MFS) using Kaplan–Meier curves and Cox regression analyses in a cohort of 461 patients with early breast cancer. Correlations were analyzed using the Spearman–Rho correlation coefficient. Kaplan–Meier curves revealed that the prognostic significance of the acute inflammatory IL signature was specifically pronounced in the basal-like subtype (p = 0.004, Log Rank). This signature retained independent prognostic significance in multivariate Cox regression analysis (HR 0.463, 95% CI 0.290–0.741; p = 0.001). A higher expression of the acute inflammatory IL signature was associated with longer MFS. The chronic inflammatory IL signature showed a significant prognostic effect in the whole cohort, with higher expression associated with shorter MFS (p = 0.034). Strong correlations were found between the acute inflammatory IL signature and CD8 expression (ρ = 0.391; p < 0.001) and between the chronic inflammatory IL signature and PD-1 expression (ρ = 0.627; p < 0.001). This study highlights the complex interaction between acute and chronic inflammatory interleukins in breast cancer progression and prognosis. These findings provide insight into the prognostic relevance of interleukin expression patterns in breast cancer and may inform future therapeutic strategies targeting the immune–inflammatory axis. Full article
Show Figures

Figure 1

Figure 1
<p>Kaplan–Meier diagram of the acute inflammatory IL signature (IL-12, IL-21, and IFN-<math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math>) in the basal-like subtype (<span class="html-italic">p</span> = 0.004, Log Rank).</p>
Full article ">Figure 2
<p>Kaplan–Meier diagram of the acute inflammatory IL signature (IL-12, IL-21, and IFN-<math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math>) in the whole cohort (<span class="html-italic">p</span> = 0.205, Log Rank).</p>
Full article ">Figure 3
<p>Kaplan–Meier diagram of the chronic inflammatory IL signature (IL-4, IL-5, IL-10, IL-13, IL-17, and CXCL1) in the whole cohort (<span class="html-italic">p</span> = 0.034, Log Rank).</p>
Full article ">Figure 4
<p>Correlation acute inflammatory IL signature/CD8 expression (ρ = 0.391; <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Correlation chronic inflammatory IL signature/PD-1 expression (ρ = 0.627; <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure A1
<p>Kaplan–Meier plot of the anti-inflammatory IL signature in terms of MFS of an unselected breast cancer cohort using publicly available gene expression data showing that a higher expression of the anti-inflammatory signature was significantly associated with an improved outcome (longer MFS).</p>
Full article ">Figure A2
<p>Kaplan–Meier plot of the pro-inflammatory IL signature in terms of MFS of an unselected breast cancer cohort using publicly available gene expression data showing that a higher expression of the pro-inflammatory signature was significantly associated with worse prognosis (shorter MFS).</p>
Full article ">Figure A3
<p>Correlation acute inflammatory IL signature/IgKC mRNA expression (ρ = 0.157; <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure A4
<p>Correlation acute inflammatory IL signature/CD20 mRNA expression (ρ = 0.160; <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
9 pages, 1318 KiB  
Article
Effects of Sublethal Exposure to Three Water Pollutants on Scototaxis in Rare Minnow (Gobiocypris rarus)
by Ning Qiu, Wenjing Li, Jianna Jia, Guoqiang Ma and Shitao Peng
Water 2024, 16(20), 2948; https://doi.org/10.3390/w16202948 - 16 Oct 2024
Abstract
The biological early warning system with fish behavior as the detection index is an efficient and rapid early warning technology for the ecological damage caused by water pollutants. However, the attempt to apply the scototaxis (dark preference) behavior of fish to biological early [...] Read more.
The biological early warning system with fish behavior as the detection index is an efficient and rapid early warning technology for the ecological damage caused by water pollutants. However, the attempt to apply the scototaxis (dark preference) behavior of fish to biological early warning is still relatively lacking. In this study, we delved into the dark and light preferences of the rare minnows (Gobiocypris rarus), employing three distinct tank configurations. Additionally, we systematically examined the modulating effects of environmental illumination, nutritional status, and the number of test subjects on this behavior, aiming to establish optimal experimental parameters for its observation. Furthermore, cadmium ions [Cd2+], tricaine methanesulfonate [MS222], and p-chloroaniline were employed as representative heavy metal ions, neuroactive agents, and organic toxicants, respectively, to test the impact of chemicals on scototaxis in gradient concentrations. The results demonstrated that the rare minnow exhibited a clear scototaxis (dark preference), and this behavior was not affected by the nutritional status of the test fish, the illumination, or the number of subjects. While the dark chamber was consistently the preferred location of rare minnows during the chemical exposure tests, the degree of scototaxis by the rare minnow significantly decreased at Cd2+ ≥ 3 mg/L, MS222 ≥ 11 mg/L, and p-chloroaniline ≥ 29 mg/L, suggesting a potential disruption of their innate behavioral patterns by these chemicals. These findings underscore the sensitivity of rare minnows to water pollutants. Therefore, the scototaxis behavior of rare minnows can be a potential and useful behavioral indicator for biological early warning, which can be used for early biological warning of sudden water pollution caused by chemicals such as Cd2+, MS222, and p-chloroaniline. Full article
(This article belongs to the Special Issue Monitoring and Modelling of Contaminants in Water Environment)
Show Figures

Figure 1

Figure 1
<p>Experiment apparatus. Half of Tank 1 (three walls and the top) was surrounded by gray, opaque cardboard. A lamp at the top of the room provided 500 lux of light (<b>a</b>). Tank 2 was used in a completely dark environment, and a lamp was placed over one side of the tank to create a gradient of light. The average illumination of the chamber close to the lamp and the chamber far away was 500 lux and 50 lux, respectively (<b>b</b>). For Tank 3, the walls and floor of one chamber were painted black and the other white (<b>c</b>).</p>
Full article ">Figure 2
<p>Quantitative distribution of rare minnow in two chambers in three apparatuses. Mean ± SD are shown; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Quantitative distribution of rare minnow in two chambers with different nutritional status of test fish (<b>a</b>), illumination (<b>b</b>), and number of test fish (<b>c</b>). Mean ± SD are shown; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Quantitative distribution of rare minnow in two chambers (<b>a</b>–<b>c</b>) and average number of entries to dark chamber (<b>d</b>–<b>f</b>) upon exposure to Cd<sup>2+</sup>, MS222, and <span class="html-italic">p</span>-chloroaniline. Mean ± SD are shown; * <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 ">
21 pages, 2075 KiB  
Review
Mpox and Lessons Learned in the Light of the Recent Outbreak: A Narrative Review
by Konstantinos Protopapas, Dimitra Dimopoulou, Nikolaos Kalesis, Karolina Akinosoglou and Charalampos D. Moschopoulos
Viruses 2024, 16(10), 1620; https://doi.org/10.3390/v16101620 - 16 Oct 2024
Abstract
According to the WHO, more than 90,000 cases of mpox have been reported since the 2022 worldwide outbreak, which resulted in 167 deaths, while a new outbreak in Africa since 2023 has resulted in over 18,000 cases and 617 deaths. Mpox is a [...] Read more.
According to the WHO, more than 90,000 cases of mpox have been reported since the 2022 worldwide outbreak, which resulted in 167 deaths, while a new outbreak in Africa since 2023 has resulted in over 18,000 cases and 617 deaths. Mpox is a zoonosis caused by the monkeypox virus, a double-stranded DNA virus belonging to the Orthopoxvirus genus, which causes smallpox-like illness. Until 2022, cases were predominately located in West and Central Africa, with only sporadic cases and outbreaks reported in other parts of the world. During the 2022 outbreak, the primary mode of transmission was sexual contact among men who have sex with men. The changing epidemiology of mpox resulted in new disease phenotypes and populations at risk, disproportionally affecting people who live with HIV. Commonly presenting as a mild, self-limiting illness, mpox can cause severe and protracted disease in people with HIV with a CD4 count < 200 cell/mm3. The global emergence of mpox that followed and intersected with COVID-19 mobilized the scientific community and healthcare stakeholders to provide accurate diagnostics, preventive vaccines and treatment to those most affected. Despite existing gaps, this rapid response helped to contain the outbreak, but challenges remain as new variants emerge. Preparedness and readiness to respond to the next outbreak is crucial in order to minimize the impact to the most vulnerable. Full article
(This article belongs to the Section Viral Immunology, Vaccines, and Antivirals)
Show Figures

Figure 1

Figure 1
<p>Structure of monkeypox virus (created using <a href="https://www.biorender.com" target="_blank">https://www.biorender.com</a> (accessed on 7 July 2024)).</p>
Full article ">Figure 2
<p>Manifestations of mpox (created using <a href="https://www.biorender.com" target="_blank">https://www.biorender.com</a> (accessed on 7 July 2024)).</p>
Full article ">Figure 3
<p>(<b>a</b>) Temporal evolution of the characteristic rash in a patient with mpox. Typical evolution of a cutaneous mpox lesion (from left to right): well-demarcated papule on an erythematous base; vesicle, pustular umbilicated lesion on an erythematous base; ulcerated lesion; crusted lesion. (<b>b</b>) Temporal evolution of the characteristic rash of mpox.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) Temporal evolution of the characteristic rash in a patient with mpox. Typical evolution of a cutaneous mpox lesion (from left to right): well-demarcated papule on an erythematous base; vesicle, pustular umbilicated lesion on an erythematous base; ulcerated lesion; crusted lesion. (<b>b</b>) Temporal evolution of the characteristic rash of mpox.</p>
Full article ">Figure 4
<p>Global response components to effectively battle the current mpox outbreak and future epidemics from emerging pathogens (created using <a href="https://www.biorender.com" target="_blank">https://www.biorender.com</a> (accessed on 12 July 2024)).</p>
Full article ">
22 pages, 4378 KiB  
Article
Characterization of EpCAM-Positive and EpCAM-Negative Tumor Cells in Early-Stage Breast Cancer
by Vladimir M. Perelmuter, Evgeniya S. Grigoryeva, Vladimir V. Alifanov, Anna Yu. Kalinchuk, Elena S. Andryuhova, Olga E. Savelieva, Ivan A. Patskan, Olga D. Bragina, Evgeniy Yu. Garbukov, Mariya A. Vostrikova, Marina V. Zavyalova, Evgeny V. Denisov, Nadezhda V. Cherdyntseva and Liubov A. Tashireva
Int. J. Mol. Sci. 2024, 25(20), 11109; https://doi.org/10.3390/ijms252011109 - 16 Oct 2024
Abstract
Most studies on CTCs have focused on isolating cells that express EpCAM. In this study, we emphasize the presence of EpCAM-negative and EpCAMlow CTCs, in addition to EpCAMhigh CTCs, in early BC. We evaluated stem cell markers (CD44/CD24 and CD133) and [...] Read more.
Most studies on CTCs have focused on isolating cells that express EpCAM. In this study, we emphasize the presence of EpCAM-negative and EpCAMlow CTCs, in addition to EpCAMhigh CTCs, in early BC. We evaluated stem cell markers (CD44/CD24 and CD133) and EMT markers (N-cadherin) in each subpopulation. Our findings indicate that all stemness variants were present in both EpCAMhigh and EpCAM-negative CTCs, whereas only one variant of stemness (nonCD44+CD24−/CD133+) was observed among EpCAMlow CTCs. Nearly all EpCAMhigh CTCs were represented by CD133+ stem cells. Notably, the hybrid EMT phenotype was more prevalent among EpCAM-negative CTCs. scRNA-seq of isolated CTCs and primary tumor partially confirmed this pattern. Therefore, further investigation is imperative to elucidate the prognostic significance of EpCAM-negative and EpCAMlow CTCs. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>The proportion (<b>A</b>,<b>B</b>) and count (<b>C</b>) of EpCAM<sup>high</sup>, EpCAM<sup>low</sup>, and EpCAM-negative CTCs in BC patients. When determining the proportion of each CTC type, the total sum of EpCAM<sup>high</sup>, EpCAM<sup>low</sup>, and EpCAM-negative CTCs was taken as 100%.</p>
Full article ">Figure 2
<p>Flow cytometry analysis of EpCAM<sup>high</sup> CTCs. (<b>A</b>) Evaluation of CD44/CD24 and CD133 stem markers expression. (<b>B</b>) Evaluation of epithelial cell markers cytokeratin 7/8 (CK7/8), panCK (panCK), E-cadherin, and mesenchymal cell marker N-cadherin expression in EpCAM<sup>high</sup> CTCs.</p>
Full article ">Figure 3
<p>Correlation analysis between the subsets of CTCs taking into account stem features among EpCAM<sup>high</sup> (<b>A</b>), EpCAM<sup>low</sup> (<b>B</b>), and EpCAM-negative (<b>C</b>) tumor cells. Red asterisk indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Correlation analysis between the subsets of CTCs taking into account EMT features among EpCAM<sup>high</sup> (<b>A</b>), EpCAM<sup>low</sup> (<b>B</b>), and EpCAM-negative (<b>C</b>) tumor cells. Red asterisk indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>EpCAM staining of FFPE tissue slides of breast cancer patients by IHC analysis. The red arrow indicates EpCAM<sup>high</sup> expression in tumor cells, the yellow arrow points to EpCAM<sup>low</sup> expression, and the green arrow corresponds to the absence of EpCAM expression in tumor cells.</p>
Full article ">Figure 6
<p>The proportion of EpCAM<sup>high</sup>, EpCAM<sup>low</sup>, and EpCAM-negative tumor cells in primary tumor of BC patients. (<b>A</b>) Interpersonal heterogeneity of EpCAM<sup>high</sup>, EpCAM<sup>low</sup>, and EpCAM-negative tumor cells in each patient. (<b>B</b>) The comparison of the proportion of EpCAM<sup>high</sup>, EpCAM<sup>low</sup>, and EpCAM-negative tumor cells in primary tumor.</p>
Full article ">Figure 7
<p>Manual annotation of spots by <span class="html-italic">EPCAM</span> level in BC tissue. Blue color indicates <span class="html-italic">EPCAM</span>-negative spots, green color—<span class="html-italic">EPCAM</span><sup>low</sup>, and orange color—<span class="html-italic">EPCAM</span><sup>high</sup> spots.</p>
Full article ">
21 pages, 11513 KiB  
Article
Expression Profiles of Fatty Acid Transporters and the Role of n-3 and n-6 Polyunsaturated Fatty Acids in the Porcine Endometrium
by Agnieszka Blitek and Magdalena Szymanska
Int. J. Mol. Sci. 2024, 25(20), 11102; https://doi.org/10.3390/ijms252011102 (registering DOI) - 16 Oct 2024
Abstract
Fatty acids (FAs) are important for cell membrane composition, eicosanoid synthesis, and metabolic processes. Membrane proteins that facilitate FA transport into cells include FA translocase (also known as CD36) and FA transporter proteins (encoded by SLC27A genes). The present study aimed to examine [...] Read more.
Fatty acids (FAs) are important for cell membrane composition, eicosanoid synthesis, and metabolic processes. Membrane proteins that facilitate FA transport into cells include FA translocase (also known as CD36) and FA transporter proteins (encoded by SLC27A genes). The present study aimed to examine expression profiles of FA transporters in the endometrium of cyclic and early pregnant gilts on days 3 to 20 after estrus and the possible regulation by conceptus signals and polyunsaturated FAs (PUFAs). The effect of PUFAs on prostaglandin (PG) synthesis and transcript abundance of genes related to FA action and metabolism, angiogenesis, and immune response was also determined. Day after estrus and reproductive status of animals affected FA transporter expression, with greater levels of CD36, SLC27A1, and SLC27A4 observed in pregnant than in cyclic gilts. Conceptus-conditioned medium and/or estradiol-17β stimulated SLC27A1 and CD36 expression. Among PUFAs, linoleic acid decreased SLC27A1 and SLC27A6 mRNA expression, while arachidonic, docosahexaenoic, and eicosapentaenoic acids increased SLC27A4 transcript abundance. Moreover, arachidonic acid stimulated ACOX1, CPT1A, and IL1B expression and increased PGE2 and PGI2 secretion. In turn, α-linolenic acid up-regulated VEGFA, FGF2, FABP4, and PPARG mRNA expression. These results indicate the presence of an active transport of FAs in the porcine endometrium and the role of PUFAs as modulators of the uterine activity during conceptus implantation. Full article
(This article belongs to the Section Molecular Endocrinology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Expression of <span class="html-italic">CD36</span>, <span class="html-italic">SLC27A1</span>, <span class="html-italic">SLC27A2</span>, <span class="html-italic">SLC27A3</span>, <span class="html-italic">SLC27A4</span>, and <span class="html-italic">SLC27A6</span> mRNA in the endometrium of cyclic (white bars) and pregnant (grey bars) gilts. Values from real-time PCR were normalized to geometric averaging of glyceraldehyde-3-phosphate dehydrogenase (<span class="html-italic">GAPDH</span>) and hypoxanthine phosphoribosyltransferase 1 (<span class="html-italic">HPRT1</span>) mRNA expression. Data are expressed as means ± SEM (<span class="html-italic">n</span> = 5–6). Bars marked with various letters differ among groups (a, b—cyclic; x, y, z—pregnant). Asterisks specify differences between cyclic and pregnant animals on particular days after estrus (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>Expression of CD36, SLC27A1, SLC27A4, and SLC27A6 proteins in the endometrium of cyclic (white bars) and pregnant (black bars) gilts. Representative blots are presented (panel <b>a</b>: C—cyclic, P—pregnant, M—marker, d.—day after estrus; full blots are included in <a href="#app1-ijms-25-11102" class="html-app">Supplementary Figures S1 and S2</a>). Values from densitometric analyses of bands were normalized to β-actin (ACTB) or glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and expressed as means ± SEM (<span class="html-italic">n</span> = 5; panel <b>b</b>). Bars marked with various letters differ among groups (a, b, c—cyclic; x, y, z—pregnant). Asterisks specify the differences between cyclic and pregnant animals on particular days after estrus (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>CD36 protein localization in the endometrium of days 11–12 cyclic and days 11–12 and 18–20 pregnant gilts using immunofluorescence technique. Tissue sections were incubated with Alexa–594-labeled secondary antibody (red) and counterstained with diamidino-2-phenylindole (DAPI; blue) to visualize nuclei. For negative control (NC), primary antibodies were replaced with rabbit IgG. Spleen was used as a positive control tissue for CD36 protein expression. dc: days of the estrous cycle; dp: days of pregnancy; LE: luminal epithelium; GE: glandular epithelium. Scale bars, 50 μm.</p>
Full article ">Figure 4
<p>SLC27A1, SLC27A4, and SLC27A6 protein localization in the endometrium of days 11–12 cyclic and days 11–12 and 18–20 pregnant gilts using immunohistochemical analysis. Tissue sections were counterstained with Mayer’s hematoxylin (blue staining) to visualize nuclei. For negative control (NC), primary antibodies were replaced with rabbit IgG. Kidney tissue was used as a positive control for SLC27A protein expression. Black arrows—luminal epithelium, yellow asterisks—glandular epithelium, red arrows—blood vessels. dc: days of the estrous cycle; dp: days of pregnancy. Scale bars, 50 μm.</p>
Full article ">Figure 5
<p>Effect of conceptus-conditioned medium (CCM), estradiol-17β (E2), prostaglandin E2 (PGE2), interleukin 1β (IL1β), and interferon γ (IFNγ) on the relative mRNA and protein expression of CD36 (<b>a</b>), SLC27A1 (<b>b</b>), SLC27A4 (<b>c</b>), and SLC27A6 (<b>d</b>) in the porcine endometrium. Values from real-time PCR were normalized to geometric averaging of glyceraldehyde-3-phosphate dehydrogenase (<span class="html-italic">GAPDH</span>) and hypoxanthine phosphoribosyltransferase 1 (<span class="html-italic">HPRT1</span>) mRNA expression. Values from densitometric analyses of bands were normalized to β-actin (ACTB). SLC27A1 protein was calculated as a sum of 73–75 and 60–62 kDa bands. Representative blots are presented (full blots are included in <a href="#app1-ijms-25-11102" class="html-app">Supplementary Figures S3 and S4</a>). All numerical data are expressed as means ± SEM (<span class="html-italic">n</span> = 6). Asterisks specify the differences compared with the control value (CTRL; *, <span class="html-italic">p</span> ≤ 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Effect of n-6 (linoleic acid [LA] and arachidonic acid [ARA]; grey bars) and n-3 (α-linolenic acid [ALA], docosahexaenoic acid [DHA], and eicosapentaenoic acid [EPA]; black bars) PUFAs on the relative mRNA and protein expression of CD36 (<b>a</b>), SLC27A1 (<b>b</b>), SLC27A4 (<b>c</b>), and SLC27A6 (<b>d</b>) in the endometrium. Values from real-time PCR were normalized to geometric averaging of glyceraldehyde-3-phosphate dehydrogenase (<span class="html-italic">GAPDH</span>) and hypoxanthine phosphoribosyltransferase 1 (<span class="html-italic">HPRT1</span>) mRNA expression. Values from densitometric analyses of bands were normalized to β-actin (ACTB). SLC27A1 protein was calculated as a sum of 73–75 and 60–62 kDa bands. Representative blots are presented (full blots are included in <a href="#app1-ijms-25-11102" class="html-app">Supplementary Figures S5 and S6</a>). All numerical data are expressed as means ± SEM (<span class="html-italic">n</span> = 5). Asterisks specify the differences compared with the control value (CTRL; *, <span class="html-italic">p</span> ≤ 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>Effect of n-6 (linoleic acid [LA] and arachidonic acid [ARA]; grey bars) and n-3 (α-linolenic acid [ALA], docosahexaenoic acid [DHA], and eicosapentaenoic acid [EPA]; black bars) PUFAs on the relative mRNA and protein expression of prostaglandin E synthase (PTGES; <b>a</b>) and prostaglandin I2 synthase (PTGIS; <b>c</b>) in the endometrium as well as prostaglandin E2 (PGE2; <b>b</b>) and 6-keto prostaglandin F1α (6-keto PGF1α, a stable metabolite of PGI2; <b>d</b>) concentrations in the incubation medium. Values from real-time PCR were normalized to geometric averaging of glyceraldehyde-3-phosphate dehydrogenase (<span class="html-italic">GAPDH</span>) and hypoxanthine phosphoribosyltransferase 1 (<span class="html-italic">HPRT1</span>) mRNA expression. Values from densitometric analyses of bands were normalized to β-actin (ACTB) or GAPDH. Representative blots are presented (full blots are included in <a href="#app1-ijms-25-11102" class="html-app">Supplementary Figure S7</a>). All numerical data are expressed as means ± SEM (<span class="html-italic">n</span> = 5). Asterisks specify the differences compared with the control value (CTRL; *, <span class="html-italic">p</span> ≤ 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 8
<p>Effect of n-6 (linoleic acid [LA] and arachidonic acid [ARA]; grey bars) and n-3 (α-linolenic acid [ALA], docosahexaenoic acid [DHA], and eicosapentaenoic acid [EPA]; black bars) PUFAs on the relative mRNA expression of genes involved in fatty acid binding (<span class="html-italic">FABP3</span>, <span class="html-italic">FABP4</span>, <span class="html-italic">FABP5</span>; <b>a</b>), action (<span class="html-italic">PPARA</span>, <span class="html-italic">PPARD</span>, <span class="html-italic">PPARG</span>; <b>b</b>), and metabolism (<span class="html-italic">ACOX1</span>, <span class="html-italic">CPT1A</span>; <b>c</b>) as well as genes related to immune response (<span class="html-italic">IL1B</span>, <span class="html-italic">IL6</span>, <span class="html-italic">TNF</span>; <b>d</b>) and angiogenesis (<span class="html-italic">VEGFA</span>, <span class="html-italic">FGF2</span>; <b>e</b>) in the endometrium. Values from real-time PCR were normalized to geometric averaging of glyceraldehyde-3-phosphate dehydrogenase (<span class="html-italic">GAPDH</span>) and hypoxanthine phosphoribosyltransferase 1 (<span class="html-italic">HPRT1</span>) mRNA expression. Data are expressed as means ± SEM (<span class="html-italic">n</span> = 5). Asterisks specify the differences compared with the control value (CTRL; *, <span class="html-italic">p</span> ≤ 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
19 pages, 850 KiB  
Review
Minimally Modified HIV-1 Infection of Macaques: Development, Utility, and Limitations of Current Models
by Manish Sharma, Mukta Nag and Gregory Q. Del Prete
Viruses 2024, 16(10), 1618; https://doi.org/10.3390/v16101618 (registering DOI) - 16 Oct 2024
Abstract
Nonhuman primate (NHP) studies that utilize simian immunodeficiency virus (SIV) to model human immunodeficiency virus (HIV-1) infection have proven to be powerful, highly informative research tools. However, there are substantial differences between SIV and HIV-1. Accordingly, there are numerous research questions for which [...] Read more.
Nonhuman primate (NHP) studies that utilize simian immunodeficiency virus (SIV) to model human immunodeficiency virus (HIV-1) infection have proven to be powerful, highly informative research tools. However, there are substantial differences between SIV and HIV-1. Accordingly, there are numerous research questions for which SIV-based models are not well suited, including studies of certain aspects of basic HIV-1 biology, and pre-clinical evaluations of many proposed HIV-1 treatment, prevention, and vaccination strategies. To overcome these limitations of NHP models of HIV-1 infection, several groups have pursued the derivation of a minimally modified HIV-1 (mmHIV-1) capable of establishing pathogenic infection in macaques that authentically recapitulates key features of HIV-1 in humans. These efforts have focused on three complementary objectives: (1) engineering HIV-1 to circumvent species-specific cellular restriction factors that otherwise potently inhibit HIV-1 in macaques, (2) introduction of a C chemokine receptor type 5 (CCR5)-tropic envelope, ideally that can efficiently engage macaque CD4, and (3) correction of gene expression defects inadvertently introduced during viral genome manipulations. While some progress has been made toward development of mmHIV-1 variants for use in each of the three macaque species (pigtail, cynomolgus, and rhesus), model development progress has been most promising in pigtail macaques (PTMs), which do not express an HIV-1-restricting tripartite motif-containing protein 5 α (TRIM5α). In our work, we have derived a CCR5-tropic mmHIV-1 clone designated stHIV-A19 that comprises 94% HIV-1 genome sequence and replicates to high acute-phase titers in PTMs. In animals treated with a cell-depleting CD8α antibody at the time of infection, stHIV-A19 maintains chronically elevated plasma viral loads with progressive CD4+ T-cell loss and the development of acquired immune-deficiency syndrome (AIDS)-defining clinical endpoints. However, in the absence of CD8α+ cell depletion, no mmHIV-1 model has yet displayed high levels of chronic viremia or AIDS-like pathogenesis. Here, we review mmHIV-1 development approaches, the phenotypes, features, limitations, and potential utility of currently available mmHIV-1s, and propose future directions to further advance these models. Full article
Show Figures

Figure 1

Figure 1
<p>Derivation pathway of minimally modified HIV-1 infectious molecular clones that have been evaluated in nonhuman primates (NHPs). Color coding corresponds to the NHP species in which each virus has been inoculated. X4-tropic, CXCR4-tropic; R5-tropic, CCR5 tropic; X4/R5-tropic, CXCR4/CCR5 dual tropic; Env, envelope; CA, capsid; Vif, viral infectivity factor; Vpu, virus protein u; PTM, pigtail macaque; CM, cynomolgus macaque; RM, rhesus macaque.</p>
Full article ">
Back to TopTop