Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,312)

Search Parameters:
Keywords = interferon response

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 768 KiB  
Review
Alternative Splicing as a Modulator of the Interferon-Gamma Pathway
by Parul Suri, Ariana Badalov and Matteo Ruggiu
Cancers 2025, 17(4), 594; https://doi.org/10.3390/cancers17040594 (registering DOI) - 10 Feb 2025
Viewed by 111
Abstract
Interferon-gamma (IFN-γ) is a critical cytokine that plays a pivotal role in immune system regulation. It is a key mediator of both cellular defense mechanisms and antitumor immunity. As the sole member of the type II interferon family, IFN-γ modulates immune responses by [...] Read more.
Interferon-gamma (IFN-γ) is a critical cytokine that plays a pivotal role in immune system regulation. It is a key mediator of both cellular defense mechanisms and antitumor immunity. As the sole member of the type II interferon family, IFN-γ modulates immune responses by activating macrophages, enhancing natural killer cell function, and regulating gene expression across multiple cellular processes. Alternative splicing is a post-transcriptional gene expression regulatory mechanism that generates multiple mature messenger RNAs from a single gene, dramatically increasing proteome diversity without the need of a proportional genome expansion. This process occurs in 90–95% of human genes, with alternative splicing events allowing for the production of diverse protein isoforms that can have distinct—or even opposing—functional properties. Alternative splicing plays a crucial role in cancer immunology, potentially generating tumor neoepitopes and modulating immune responses. However, how alternative splicing affects IFN-γ’s activity is still poorly understood. This review explores how alternative splicing regulates the expression and function of both upstream regulators and downstream effectors of IFN-γ, revealing complex mechanisms of gene expression and immune response modulation. Key transcription factors and signaling molecules of the IFN-γ pathway are alternatively spliced, and alternative splicing can dramatically alter IFN-γ signaling, immune cell function, and response to environmental cues. Specific splice variants can enhance or inhibit IFN-γ-mediated immune responses, potentially influencing cancer immunotherapy, autoimmune conditions, and infectious disease outcomes. The emerging understanding of these splicing events offers promising therapeutic strategies for manipulating immune responses through targeted molecular interventions. Full article
(This article belongs to the Special Issue IFN-Gamma Signaling in Cancer)
Show Figures

Figure 1

Figure 1
<p>Different types of alternative splicing. Exons are represented by boxes, and introns are represented by lines. Constitutive exons are shown in blue, while alternative exons are shown in red and green. Promoters are indicated with arrows, while different polyadenylation sites are indicated with AAA. There are seven major alternative splicing modes that represent binary choices: cassette exons or exon skipping (<b>A</b>), mutually exclusive exons (<b>B</b>), alternative 5′ splice sites (<b>C</b>), alternative 3′ splice sites (<b>D</b>), intron retention (<b>E</b>), alternative promoters (<b>F</b>), and alternative polyadenylation sites (<b>G</b>). Multiple alternative splicing events can also appear in different combinations on the same pre-mRNA (<b>H</b>), giving rise to a large repertoire of mRNAs originating from a single pre-mRNA.</p>
Full article ">Figure 2
<p>Alternative splicing modulates the activity of the IFN-γ pathway. Schematic representation of upstream regulators and downstream effectors of IFN-γ activity discussed in this review.</p>
Full article ">
24 pages, 11308 KiB  
Article
Microbiomic and Metabolomic Insights into the Mechanisms of Alfalfa Polysaccharides and Seaweed Polysaccharides in Alleviating Diarrhea in Pre-Weaning Holstein Calves
by Jianan Zhao, Haoliang Tian, Xiaohui Kong, Danqi Dang, Kaizhen Liu, Chuanyou Su, Hongxia Lian, Tengyun Gao, Tong Fu, Liyang Zhang, Wenqing Li and Wei Zhang
Animals 2025, 15(4), 485; https://doi.org/10.3390/ani15040485 - 8 Feb 2025
Viewed by 155
Abstract
Neonatal calves’ diarrhea, which can be severe enough to cause death, has a significant impact on the global cattle industry. In this study, alfalfa polysaccharides and seaweed polysaccharides were found to significantly improve the diarrhea condition in neonatal calves. To explore the underlying [...] Read more.
Neonatal calves’ diarrhea, which can be severe enough to cause death, has a significant impact on the global cattle industry. In this study, alfalfa polysaccharides and seaweed polysaccharides were found to significantly improve the diarrhea condition in neonatal calves. To explore the underlying mechanisms, further microbiomic and metabolomic analyses were conducted. This study investigated the impact of alfalfa polysaccharides and seaweed polysaccharides on growth performance, serum metabolites, gut microbiota, and metabolomics in neonatal Holstein calves. A total of 24 newborn calves were randomly assigned to three groups, with 8 calves per treatment group. The control (CON) group was fed a basal diet, the alfalfa polysaccharide (AP) group received a basal diet supplemented with alfalfa polysaccharides (4 g/calf/day), and the seaweed polysaccharide group (SP) received a basal diet supplemented with seaweed polysaccharides (4 g/calf/day). These polysaccharides were plant extracts. Compared to the CON group, the results indicated that SP significantly enhanced the body weight, height, chest circumference, and average daily gain of Holstein calves (p < 0.05), while also reducing the diarrhea rate and improving manure scoring (p < 0.05). Compared to the CON, AP also reduced the diarrhea rate (p < 0.05). In terms of serum biochemistry, supplementation with AP and SP increased serum alkaline phosphatase (ALP) and insulin-like growth factor 1 (IGF-1) levels compared to the CON group (p < 0.05). Both AP and SP elevated serum catalase (CAT) and Total Antioxidant Capacity (T-AOC) levels, indicating enhanced antioxidant status (p < 0.05). Regarding immune responses, supplementation with AP and SP significantly increased serum complement component 3 (C3) and immunoglobulin M (IgM) levels, while significantly reducing pro-inflammatory cytokines interleukin-18 (IL-18), tumor necrosis factor alpha (TNF-α), and interferon-gamma (IFN-γ) compared to the CON group (p < 0.05). Microbiota analysis revealed that AP modulated the abundance of Firmicutes, while SP influenced the abundance of Prevotella and Succiniclasticum. AP and SP differentially influenced intestinal metabolites compared to the CON group, leading to enrichment in pathways related to immunity, antibacterial, and anti-inflammatory functions. These pathways included the biosynthesis of alkaloids from ornithine, lysine, and nicotinic acid, glucocorticoid and mineralocorticoid receptor canothersis/antagonists, secondary metabolite biosynthesis, and alkaloid biosynthesis from histidine and purine, thus alleviating intestinal inflammation. Therefore, by supplementing with AP and SP, the diarrhea rate in calves was reduced, and the immune function of Holstein calves was enhanced, while simultaneously promoting a higher relative abundance of beneficial gut bacteria and suppressing the relative abundance of pathogenic bacteria. Additionally, gut pathways associated with immune response and inflammation were modulated by AP and SP. This study provided valuable insights and theoretical underpinnings for the use of AP and SP in preventing diarrhea in neonatal calves. Full article
(This article belongs to the Section Cattle)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Body weight, (<b>B</b>) body height, (<b>C</b>) body length, (<b>D</b>) chest size, (<b>E</b>) average daily intake, (<b>F</b>) average daily gain, (<b>G</b>) diarrhea rate, (<b>H</b>) manure scoring. * <span class="html-italic">p</span> &lt; 0.05 compared with the CON group.</p>
Full article ">Figure 2
<p>Effects of AP and SP on calf serum metabolites. (<b>A</b>) Superoxide dismutase, (<b>B</b>) Malondialdehyde, (<b>C</b>) Total antioxidant capacity, (<b>D</b>) Catalase, (<b>E</b>) Immunoglobulin A, (<b>F</b>) Immunoglobulin G, (<b>G</b>) Immunoglobulin M, (<b>H</b>) Complement component 3, (<b>I</b>) Complement component 4, (<b>J</b>) Total protein, (<b>K</b>) Alanine aminotransferase, (<b>L</b>) Aspartate aminotransferase, (<b>M</b>) Albumin, (<b>N</b>) Alkaline phosphatase, (<b>O</b>) Interleukin-4, (<b>P</b>) Interleukin-18, (<b>Q</b>) Interferon-gamma, (<b>R</b>) Tumor necrosis factor-alpha, (<b>S</b>) Growth hormone, (<b>T</b>) Insulin-like growth factor 1. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; compared with the CON group.</p>
Full article ">Figure 3
<p>Effects of MOP on hindgut microbiota diversity. (<b>A</b>) Venn diagram of OTUs, (<b>B</b>) alpha diversity indices in each group, (<b>C</b>) PCoA plot of the gut microflora in all groups, (<b>D</b>) species observation curve in each group.</p>
Full article ">Figure 4
<p>Effects of AP and SP on hindgut microbiota of calves. (<b>A</b>) Phylum level. (<b>B</b>) Genus level.</p>
Full article ">Figure 5
<p>LEfSe analysis and cladogram of gut microbiota in each group.</p>
Full article ">Figure 6
<p>(<b>A</b>) Overview of PICRUSt2 functional distribution in the CON group. (<b>B</b>) Functional classification at KEGG Level 2. (<b>C</b>) Differential functions between two comparison groups at KEGG Level 3.</p>
Full article ">Figure 7
<p>(<b>A</b>) OPLS-DA plot of CON vs. AP. (<b>B</b>) OPLS-DA plot of CON vs. SP. (<b>C</b>) Volcano plot of CON vs. AP. (<b>D</b>) Volcano plot of CON vs. SP. (<b>E</b>) Differential Venn diagram. (<b>F</b>) Heatmap of metabolite expression. (<b>G</b>) Significance bubble plot of CON vs. AP. (<b>H</b>) Significance bubble plot of CON vs. SP. Red indicates upregulation of metabolite expression; blue indicates downregulation of expression.</p>
Full article ">Figure 8
<p>(<b>A</b>) Correlation analysis between gut microbiota at the genus level and serum indices. (<b>B</b>) Module hierarchical clustering. (<b>C</b>) Heatmap of traits and module correlations. (<b>D</b>,<b>F</b>) KEGG enrichment analysis within modules. (<b>E</b>) Correlation analysis between gut microbiota and differential metabolites. * <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 ">
16 pages, 1447 KiB  
Article
Evaluation of the Immunogenicity of a Pool of Recombinant Lactococcus lactis Expressing Eight Antigens of African Swine Fever Virus in a Mouse Model
by Jingshan Huang, Tianqi Gao, Zhanhao Lu, Dailang Zhong, Mingzhi Li, Hua-Ji Qiu, Yongfeng Li, Hongxia Wu and Yuan Sun
Vet. Sci. 2025, 12(2), 140; https://doi.org/10.3390/vetsci12020140 - 7 Feb 2025
Viewed by 235
Abstract
African swine fever (ASF), caused by African swine fever virus (ASFV), poses a great threat to the global pig industry. There is an urgent demand for effective and safe vaccines to address this threat. This study reports the generation and evaluation of a [...] Read more.
African swine fever (ASF), caused by African swine fever virus (ASFV), poses a great threat to the global pig industry. There is an urgent demand for effective and safe vaccines to address this threat. This study reports the generation and evaluation of a recombinant Lactococcus lactis pool, each strain expressing one of eight ASFV antigens (F317L, H171R, D117L, E120R, B602L, CD2v, p54, and p72). We evaluated the immune responses in mice through oral gavage and intramuscular immunization to the recombinant L. lactis pool. The results show that the mice immunized via intramuscular injection induced high-level serum IgG antibodies within 7 d post-primary immunization, which was maintained over an extended period. Additionally, there was a marked increase in the interferon gamma (IFN-γ) and interleukin 10 (IL-10) levels in the sera. In contrast, the mice immunized via oral gavage did not induce obvious serum IgG antibodies. However, they experienced a transient peak of secretory IgA (sIgA) antibodies in fecal samples within 7 d post-primary immunization, which subsequently decreased to levels that were statistically similar to those in the control group. In addition, this study also found that the multi-antigen cocktail vaccination was safe for mice. This study provides a reference for the development and immunization of ASF vaccines with L. lactis as live carriers. Full article
16 pages, 3040 KiB  
Article
In Vitro and In Vivo Evaluation of Bacillus Strains as Prophylactic Agents Against Porcine Epidemic Diarrhea Virus
by You-Jia Chen, Chia-Fang Tsai, Chin-Wei Hsu, Hui-Wen Chang and Je-Ruei Liu
Animals 2025, 15(4), 470; https://doi.org/10.3390/ani15040470 - 7 Feb 2025
Viewed by 264
Abstract
Porcine epidemic diarrhea virus (PEDV), particularly the highly virulent G2b strains, has inflicted substantial economic losses on the global swine industry. This study evaluated the prophylactic effects of three Bacillus strains—B. amyloliquefaciens LN, B. licheniformis CK, and B. velezensis AC—against PEDV infection using in vitro [...] Read more.
Porcine epidemic diarrhea virus (PEDV), particularly the highly virulent G2b strains, has inflicted substantial economic losses on the global swine industry. This study evaluated the prophylactic effects of three Bacillus strains—B. amyloliquefaciens LN, B. licheniformis CK, and B. velezensis AC—against PEDV infection using in vitro and in vivo models. In vitro experiments with Vero cells demonstrated that B. amyloliquefaciens LN increased cell viability, reduced PEDV-N expression, and modulated proinflammatory cytokine responses. In vivo, piglets supplemented with B. amyloliquefaciens LN exhibited alleviated diarrhea symptoms, suppression of fecal viral RNA shedding to below the detection limit, and restoration of gut microbiota balance by increasing Bacteroidetes and reducing Proteobacteria abundance. Mechanistic studies indicated that the measured interferon (IFN)-related genes were not significantly influenced in this study, suggesting that the protective effects of B. amyloliquefaciens LN may involve the modulation of inflammatory responses and the inhibition of viral replication through reduced PEDV-N expression. This study illustrates the potential of using B. amyloliquefaciens LN as a feed additive to prevent PEDV infection. Full article
(This article belongs to the Special Issue Infections and Diarrhea of Enteric Coronaviruses in Pigs)
Show Figures

Figure 1

Figure 1
<p>Protective effects of <span class="html-italic">Bacillus</span> intracellular extracts and cell-wall fractions against PEDV infection in Vero cells. The figure shows the prophylactic effects of intracellular extracts (IE) and cell-wall fractions (CW) from <span class="html-italic">Bacillus amyloliquefaciens</span> LN, <span class="html-italic">B</span>. <span class="html-italic">licheniformis</span> CK, and <span class="html-italic">B</span>. <span class="html-italic">velezensis</span> AC on Vero cells infected with porcine epidemic diarrhea virus (PEDV). Experimental groups included: Mock (PBS-pretreated, mock-infected), PEDV (PBS-pretreated, PEDV-infected), PEDV+IFN (IFN-α2b-pretreated, PEDV-infected), and PEDV+LN/IE, PEDV+LN/CW, PEDV+CK/IE, PEDV+CK/CW, PEDV+AC/IE, PEDV+AC/CW (pretreated with intracellular extracts or cell-wall fractions from LN, CK, or AC strains, followed by PEDV infection). Vero cells were pretreated with 100 μL of <span class="html-italic">Bacillus</span> preparations or IFN-α2b (6500 IU) for 24 h before infection with PEDV at a dose of 40 TCID<sub>50</sub> per well. Data are expressed as a mean ± SD (<span class="html-italic">n</span> = 3). Bars sharing the same letter indicate no significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2
<p>Impact of <span class="html-italic">Bacillus</span> intracellular extracts and cell-wall fractions on PEDV-N expression in Vero cells. The figure illustrates the effects of intracellular extracts (IE) and cell-wall fractions (CW) from <span class="html-italic">Bacillus amyloliquefaciens</span> LN and <span class="html-italic">B</span>. <span class="html-italic">velezensis</span> AC on the expression of the porcine epidemic diarrhea virus nucleocapsid (PEDV-N) protein in Vero cells. (<b>A</b>) Intracellular PEDV-N expression levels. (<b>B</b>) Extracellular PEDV-N expression levels. The experimental groups are consistent with those described in <a href="#animals-15-00470-f001" class="html-fig">Figure 1</a>. Data are presented as a mean ± SD (<span class="html-italic">n</span> = 3). Bars sharing the same letter indicate no significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Effects of <span class="html-italic">Bacillus</span> intracellular extracts and cell-wall fractions on ISG expression in PEDV-infected Vero cells. The figure depicts the impact of intracellular extracts (IE) and cell-wall fractions (CW) from <span class="html-italic">Bacillus amyloliquefaciens</span> LN and <span class="html-italic">B</span>. <span class="html-italic">velezensis</span> AC on the expression levels of interferon-stimulated genes (ISGs) in Vero cells infected with porcine epidemic diarrhea virus (PEDV). (<b>A</b>) Interferon-stimulated gene 15 (ISG15). (<b>B</b>) Myxovirus resistance 1 (Mx1). (<b>C</b>) 2′-5′-Oligoadenylate synthetase 1 (OAS1). Groups: NC (PBS-pretreated), IFN (IFN-α2b-pretreated), LN/IE (intracellular extract of LN), LN/CW (cell-wall fraction of LN), AC/IE (intracellular extract of AC), and AC/CW (cell-wall fraction of AC). Data are shown as a mean ± SD (<span class="html-italic">n</span> = 3). Bars sharing the same letter indicate no significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">Bacillus</span> intracellular extracts and cell-wall fractions on proinflammatory cytokine expression in PEDV-infected Vero cells. The figure illustrates the effects of intracellular extracts (IE) and cell-wall fractions (CW) from <span class="html-italic">Bacillus amyloliquefaciens</span> LN and <span class="html-italic">B</span>. <span class="html-italic">velezensis</span> AC on the expression levels of proinflammatory cytokines in Vero cells infected with porcine epidemic diarrhea virus (PEDV). (<b>A</b>) Interleukin (IL)-1β. (<b>B</b>) IL-6. (<b>C</b>) IL-8. (<b>D</b>) Tumor necrosis factor-α (TNF-α). Groups: NC (negative control, PBS-pretreated), IFN (IFN-α2b-pretreated), LN/IE (intracellular extract of LN), LN/CW (cell-wall fraction of LN), AC/IE (intracellular extract of AC), and AC/CW (cell-wall fraction of AC), as described in <a href="#animals-15-00470-f003" class="html-fig">Figure 3</a>. Data are presented as a mean ± SD (<span class="html-italic">n</span> = 3). Bars sharing the same letter indicate no significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 5
<p>Effects of dietary supplementation with <span class="html-italic">Bacillus amyloliquefaciens</span> LN or <span class="html-italic">B. velezensis</span> AC on fecal consistency in porcine epidemic diarrhea virus (PEDV)-infected piglets. (<b>A</b>) Mock: mock-infected piglets. (<b>B</b>) PEDV: PEDV-infected piglets. (<b>C</b>) PEDV+LN: PEDV-infected piglets supplemented with <span class="html-italic">B. amyloliquefaciens</span> LN. (<b>D</b>) PEDV+AC: PEDV-infected piglets supplemented with <span class="html-italic">B. velezensis</span> AC. Fecal consistency was scored daily using a 4-point scale: 0 (normal), 1 (pasty feces), 2 (semi-fluid feces), 3 (watery diarrhea).</p>
Full article ">Figure 6
<p>Effects of dietary supplementation with <span class="html-italic">Bacillus amyloliquefaciens</span> LN or <span class="html-italic">B. velezensis</span> AC on fecal viral shedding in porcine epidemic diarrhea virus (PEDV)-infected piglets. Groups: Mock (mock-infected piglets), PEDV (PEDV-infected piglets), PEDV+LN (PEDV-infected piglets supplemented with <span class="html-italic">B. amyloliquefaciens</span> LN), and PEDV+AC (PEDV-infected piglets supplemented with <span class="html-italic">B. velezensis</span> AC).</p>
Full article ">Figure 7
<p>Effects of dietary supplementation with <span class="html-italic">Bacillus amyloliquefaciens</span> LN or <span class="html-italic">B. velezensis</span> AC on fecal microbiota in porcine epidemic diarrhea virus (PEDV)-infected piglets. (<b>A</b>) The Shannon diversity index showing alpha diversity. (<b>B</b>) Principal coordinates analysis (PCoA) of beta diversity based on the Bray–Curtis dissimilarity. (<b>C</b>) Relative abundance at the phylum level. (<b>D</b>) Relative abundance of dominant bacterial families. Groups: Mock (mock-infected piglets), PEDV (PEDV-infected piglets), PEDV+LN (PEDV-infected piglets supplemented with <span class="html-italic">B. amyloliquefaciens</span> LN), and PEDV+AC (PEDV-infected piglets supplemented with <span class="html-italic">B. velezensis</span> AC).</p>
Full article ">
13 pages, 2417 KiB  
Article
Neutralizing IL-15 Inhibits Tissue-Damaging Immune Response in Ex Vivo Cultured Untreated Celiac Intestinal Mucosa
by Vera Rotondi Aufiero, Giuseppe Iacomino, Giovanni De Chiara, Errico Picariello, Gaetano Iaquinto, Riccardo Troncone and Giuseppe Mazzarella
Cells 2025, 14(3), 234; https://doi.org/10.3390/cells14030234 - 6 Feb 2025
Viewed by 323
Abstract
In celiac disease (CeD), interleukin 15 (IL-15) affects the epithelial barrier by acting on intraepithelial lymphocytes, promoting interferon γ (IFN-γ) production and inducing strong cytotoxic activity as well as eliciting apoptotic death of enterocytes by the Fas/Fas ligand system. This study investigates the [...] Read more.
In celiac disease (CeD), interleukin 15 (IL-15) affects the epithelial barrier by acting on intraepithelial lymphocytes, promoting interferon γ (IFN-γ) production and inducing strong cytotoxic activity as well as eliciting apoptotic death of enterocytes by the Fas/Fas ligand system. This study investigates the effects of a monoclonal antibody neutralizing the effects of IL-15 (aIL-15) on tissue-damaging immune response in untreated CeD patients by using an organ culture system. Jejunal biopsies from 10 untreated CeD patients were cultured ex vivo with or without aIL-15. Epithelial expressions of CD95/Fas, HLA-E and perforin were analyzed by immunohistochemistry. Apoptosis was detected in the epithelium by using the terminal deoxynucleotidyl transferase-mediated dUTP nick-end labeling (TUNEL) assay. Additionally, the surface epithelium compartment of ex vivo cultured biopsy samples was isolated by laser capture microdissection (LCM). RNA from each LCM sample was extracted and the relative expression of IFN-γ was evaluated by quantitative reverse transcriptase-PCR (qRT-PCR). Biopsies cultured with the aIL-15 antibody showed a reduction in Fas, HLA-E and perforin epithelial expression, as well as a decrease in epithelial TUNEL+ cells compared to biopsies cultured without the aIL-15 antibody. Moreover, downregulation of epithelial IFN-γ expression was recorded in biopsies incubated with aIL-15, compared to those cultured without aIL-15. Our findings suggest that neutralizing the effects of IL-15 in ex vivo cultured untreated CeD intestinal mucosa could block apoptosis by downregulating Fas and HLA-E expression and the release of cytotoxic proteins, such as perforin. Furthermore, it can dampen the hyperactive immune response by reducing IFN-γ expression. More generally, our study provides new evidence for the effects of anti-IL-15 neutralizing monoclonal antibodies in preventing or repairing epithelial damage and further supports the concept that IL-15 is a meaningful therapeutic target in CeD, or inflammatory diseases associated with the upregulation of IL-15. Full article
Show Figures

Figure 1

Figure 1
<p><b>Upper left</b> panel: number of perforin cytotoxic granules analyzed by immunohistochemistry, in mucosal explants from untreated CeD cultured ex vivo without (w/o) or with aIL-15 antibody. Perforin cytotoxic granules were counted in 1 mm epithelium from at least four different fields. The mean value is reported for each subject, and dashes indicate the mean values. Statistical significance was evaluated by comparing responses without or with aIL-15 antibody (* <span class="html-italic">p</span> &lt; 0.01). <b>Upper right</b> panel: perforin cytotoxic granules (brown) in the epithelium of jejunal mucosa from untreated CeD patient cultured ex vivo without or with aIL-15 antibody. In the latter, a decrease in perforin cytotoxic granules (brown) in the epithelial compartment is evident. The image is representative of ten separate experiments in which biopsies taken from ten patients with untreated CeD cultured with or without the antibody aIL-15 were analyzed. Original magnification ×63; scale bar 10 µm.</p>
Full article ">Figure 2
<p><b>Upper left</b> panel: Fas epithelial expression is decreased in the surface epithelium of duodenal mucosa of untreated CeD cultured with aIL-15 antibody. Fas expression in intestinal surface epithelium was evaluated in terms of staining intensity and graded on an arbitrary scale of staining from 1 to 3. The criteria for epithelium staining were as follows: weak staining (+) = 1, moderate staining (++) = 2, and strong staining (+++) = 3. Circles represent the response from individual patients. * <span class="html-italic">p</span> &lt; 0.01, ** <span class="html-italic">p</span> &lt; 0.001. <b>Upper right</b> panel: Fas expression in the epithelium of jejunal mucosa from n = 10 untreated CeD cultured ex vivo without (w/o) or with aIL-15 antibody. In the latter, lower staining is detected, particularly in almost all the epithelial cells. Original magnification ×63, scale bar 10 µm.</p>
Full article ">Figure 3
<p>Relative levels of IFN-γ in the surface epithelium (Ep) compartment isolated by LCM from jejunal biopsies cultured ex vivo without (w/o) or with aIL-15 antibody. The biopsies were analyzed by RT-qPCR from the untreated CeD. The fold change represents the relative expression of IFN-γ mRNA normalized to GAPDH. Each point on the plot is representative of a distinct patient (<span class="html-italic">n</span> = 5). The lines link the IFN-γ gene expression of each subject in the two different experimental conditions.</p>
Full article ">Figure 4
<p>A possible model whereby anti-IL-15 may block apoptosis mediated by cytolytic mechanisms in celiac disease. (1) CD8+IELs, activated by IL-15, synthesize cytotoxic molecules such as perforin that cause the cytolysis of enterocytes. (2) Moreover, activated CD8+IELs synthesize IFN-γ, which induces enterocyte HLA-E and Fas expression. HLA-E and Fas can combine with the CD8+ IELs cell-surface receptors CD94/NKG2C and FasL, respectively, contributing to enterocyte apoptosis. (3) Anti-IL-15 may limit cytotoxic T-cell function by inhibiting the expression of perforin, HLA-E, Fas and IFN-γ.</p>
Full article ">
11 pages, 4355 KiB  
Case Report
Peripheral Blood Mononuclear Cells Cytokine Profile in a Patient with Toxic Epidermal Necrolysis Triggered by Lamotrigine and COVID-19: A Case Study
by Margarita L. Martinez-Fierro, Idalia Garza-Veloz, Sidere Monserrath Zorrilla-Alfaro, Andrés Eduardo Campuzano-Garcia and Monica Rodriguez-Borroel
Int. J. Mol. Sci. 2025, 26(3), 1374; https://doi.org/10.3390/ijms26031374 - 6 Feb 2025
Viewed by 264
Abstract
Stevens–Johnson Syndrome (SJS)/toxic epidermal necrolysis (TEN) is a severe mucocutaneous reaction often induced by medications. The co-occurrence of SJS/TEN and COVID-19 presents a unique challenge due to overlapping inflammatory pathways. This case study examined the cytokine profile of a patient with both TEN [...] Read more.
Stevens–Johnson Syndrome (SJS)/toxic epidermal necrolysis (TEN) is a severe mucocutaneous reaction often induced by medications. The co-occurrence of SJS/TEN and COVID-19 presents a unique challenge due to overlapping inflammatory pathways. This case study examined the cytokine profile of a patient with both TEN (triggered by lamotrigine) and COVID-19. The clinical history of the patient, including lamotrigine exposure and COVID-19 diagnosis, was documented. A 13-cytokine profile assessment was performed in peripheral blood mononuclear cells from the patient and their parents by using quantitative Real Time-Polymerase Chain Reaction (qRT-PCR). A 6-year-old male patient developed lamotrigine-induced TEN with concomitant COVID-19 affecting 90% of the body surface area. Compared with their parents, who were positive for COVID-19, IL-6, IL-4, and IL-12 were modulated (downregulated) by TEN. The cytokine profile showed elevated levels of IL-1α, IL-1β, IL-5, IL-8, NF-κβ, and interferons (IFN; α, β, and γ), indicating a robust antiviral response. The immune profile suggested a hyperactivated immune state that contributed to the severity of the patient’s clinical manifestations, leading to death 18 days after hospitalization. Understanding the immune response is important for developing future targeted therapeutic strategies and improving patient outcomes. Further research is needed to explore the interaction between drug-induced SJS/TEN and infections. Full article
(This article belongs to the Special Issue Targeted Therapy for Immune Diseases)
Show Figures

Figure 1

Figure 1
<p>Toxic epidermal necrolysis clinical spectrum. (<b>A</b>–<b>D</b>): exanthematous rash. Lesions start on the face and thorax before spreading to other areas and are symmetrically distributed. Early lesions typically begin with ill-defined, coalescing, erythematous macules; (<b>E</b>–<b>H</b>): extensive, sheet-like detachment and erosions, and Nikolsky sign is present.</p>
Full article ">Figure 2
<p>Expression profile of evaluated cytokines. Figure shows the expression level of a 13-citokine panel for the patient with toxic epidermal necrolysis (TEN) and concurrent COVID-19, and for his parents, grouped by immune response in which they participate (<b>A</b>); and according with their status of overexpression and underexpression profile (<b>B</b>). Expression levels were calculated by quantitative real-time polymerase chain reaction by using GAPDH as endogenous control and RNA of peripheral blood mononuclear cell obtained from healthy controls (with negative qRT-PCR for SARS-CoV-2) as calibrator.</p>
Full article ">Figure 3
<p>Cell and immune pathophysiology of TEN. Infiltration of the epidermis by activated T lymphocytes (CD8+ epidermis; CD4+ dermis) and natural killer cells induce an immune response against the drug-reactive metabolites. TCRs recognize the molecules and produce interleukins (mainly TNFα) which cause epidermal detachment secondary to keratinocyte apoptosis induced by granzymes, perforins, and Fas/Fas ligand. MHC-II: major histocompatibility complex class II; TCR: T lymphocyte receptor; IFNγ: interferon gamma; TNFα: tumor necrosis factor alpha; IL: interleukin.</p>
Full article ">
31 pages, 8851 KiB  
Article
Autologous Human Dendritic Cells from XDR-TB Patients Polarize a Th1 Response Which Is Bactericidal to Mycobacterium tuberculosis
by Rolanda Londt, Lynn Semple, Aliasgar Esmail, Anil Pooran, Richard Meldau, Malika Davids, Keertan Dheda and Michele Tomasicchio
Microorganisms 2025, 13(2), 345; https://doi.org/10.3390/microorganisms13020345 - 5 Feb 2025
Viewed by 402
Abstract
Extensively drug-resistant tuberculosis (XDR-TB) is a public health concern as drug resistance is outpacing the drug development pipeline. Alternative immunotherapeutic approaches are needed. Peripheral blood mononuclear cells (PBMCs) were isolated from pre-XDR/XDR-TB (n = 25) patients and LTBI (n = 18) [...] Read more.
Extensively drug-resistant tuberculosis (XDR-TB) is a public health concern as drug resistance is outpacing the drug development pipeline. Alternative immunotherapeutic approaches are needed. Peripheral blood mononuclear cells (PBMCs) were isolated from pre-XDR/XDR-TB (n = 25) patients and LTBI (n = 18) participants. Thereafter, monocytic-derived dendritic cells (mo-DCs) were co-cultured with M. tb antigens, with/without a maturation cocktail (interferon-γ, interferon-α, CD40L, IL-1β, and TLR3 and TLR7/8 agonists). Two peptide pools were evaluated: (i) an ECAT peptide pool (ESAT6, CFP10, Ag85B, and TB10.4 peptides) and (ii) a PE/PPE peptide pool. Sonicated lysate of the M. tb HN878 strain served as a control. Mo-DCs were assessed for DC maturation markers, Th1 cytokines, and the ability of the DC-primed PBMCs to restrict the growth of M. tb-infected monocyte-derived macrophages. In pre-XDR/XDR-TB, mo-DCs matured with M. tb antigens (ECAT or PE/PPE peptide pool, or HN878 lysate) + cocktail, compared to mo-DCs matured with M. tb antigens only, showed higher upregulation of co-stimulatory molecules and IL-12p70 (p < 0.001 for both comparisons). The matured mo-DCs had enhanced antigen-specific CD8+ T-cell responses to ESAT-6 (p = 0.05) and Ag85B (p = 0.03). Containment was higher with mo-DCs matured with the PE/PPE peptide pool cocktail versus mo-DCs matured with the PE/PPE peptide pool (p = 0.0002). Mo-DCs matured with the PE/PPE peptide pool + cocktail achieved better containment than the ECAT peptide pool + cocktail [50%, (IQR:39–75) versus 46%, (IQR:15–62); p = 0.02]. In patients with pre-XDR/XDR-TB, an effector response primed by mo-DCs matured with an ECAT or PE/PPE peptide pool + cocktail was capable of restricting the growth of M. tb in vitro Full article
(This article belongs to the Special Issue Mycobacterial Tuberculosis Pathogenesis and Vaccine Development)
Show Figures

Figure 1

Figure 1
<p>Schematic of the experimental procedures used to determine the efficacy of mo-DCs against <span class="html-italic">M. tb.</span> (<b>A</b>) Monocytes were differentiated into immature DCs with GM-CSF and IL-4 for five days. The immature DCs were loaded with <span class="html-italic">M. tb</span> antigens for six hours, then incubation with/without a maturation cocktail for two days. Immature and mature mo-DCs were stained for flow cytometry analysis and culture supernatants assessed for soluble cytokine secretion using ELISA. (<b>B</b>) For the generation of effector cells, PBMCs were co-cultured with mature mo-DCs for seven days. DC-primed effector cells were stained for flow cytometry, the culture supernatants assessed for soluble cytokine secretion using Luminex, and the DC-primed effector cells assessed for antigen specificity using a tetramer assay. After seven days, the mo-DC-primed effector T-cells were co-cultured with <span class="html-italic">M. tb</span> H37<span class="html-italic">Rv</span>-infected macrophages and <span class="html-italic">M. tb</span> containment (CFU/mL) determined using a stasis assay by counting colonies on agar plates. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 2
<p>The pre-XDR/XDR-TB patient recruitment and sample schematic. A total of 71 individuals presenting with resistance beyond MDR were asked to consent to the study. After screening, only 43 were eligible for consent and sample collection. In total, 18 patients were further excluded due to failure to provide adequate blood volumes for the follow-up visit or amendments to the experimental protocol. A total of 25 pre-XDR/XDR-TB patient samples were included in the final results. Samples from 14 pre-XDR/XDR-TB patients were used for the phenotypic assessment of mature mo-DCs using flow cytometry, the determination of soluble cytokine secretion of mature mo-DCs using ELISA, the CD8<sup>+</sup> specific tetramer assay for the determination of antigen specificity using flow cytometry, and the <span class="html-italic">M. tb</span> containment assay. Exceptions for the aforementioned experiments are reported. Samples from 11 pre-XDR/XDR-TB patients were used for the characterisation of DC-primed CD4<sup>+</sup> and CD8<sup>+</sup> T-cell responses and the determination of soluble cytokine secretion from the DC-primed effector cell responses. Exclusions due to experimental errors are reported.</p>
Full article ">Figure 3
<p>Matured mo-DCs from pre-XDR/XDR-TB patients expressed high levels of co-stimulatory molecules. Immature mo-DCs from pre-XDR/XDR-TB patients (<span class="html-italic">n</span> = 14) were differentiated from monocytes with IL-4 and GM-CSF for five days. After five days, the immature mo-DCs were matured with <span class="html-italic">M. tb</span> antigens, with/without the maturation cocktail for 48 h. The mo-DCs were analysed for CCR7 (<b>A</b>), CD80 (<b>B</b>), CD83 (<b>C</b>), and CD86 (<b>D</b>) by flow cytometry. The experimental conditions are immature DC, negative control (DCs matured without <span class="html-italic">M. tb</span> antigens or cocktail), mo-DCs matured with ECAT only, mo-DCs matured with PE/PPE only, mo-DCs matured with HN878 only, mo-DCs matured with limited cocktail only, mo-DCs matured with ECAT + C, mo-DCs matured with PE/PPE + C, and mo-DCs matured with HN878 + C. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as *, **, ***, **** for <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 and <span class="html-italic">p</span> &lt; 0.0001 respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 4
<p>Matured mo-DCs from pre-XDR/XDR-TB patients polarized a Th1 response. The levels of IL-12p70 (<b>A</b>) and IL-10 (<b>B</b>) from the culture supernatants (<span class="html-italic">n</span> = 9) of the immature and matured mo-DCs were determined using the ELISAPRO IL-12p70 and IL-10 detection kit (Mabtech). In order to determine the Th-polarizing response, the IL-12p70/IL-10 (<b>C</b>) ratio was assessed. The experimental conditions are immature DC, negative control (DCs matured without <span class="html-italic">M. tb</span> antigens or cocktail), mo-DCs matured with ECAT only, mo-DCs matured with PE/PPE only, mo-DCs matured with HN878 only, mo-DCs matured with limited cocktail only, mo-DCs matured with ECAT + C, mo-DCs matured with PE/PPE + C, and mo-DCs matured with HN878 + C. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as * and ** for <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 5
<p>Matured mo-DCs from pre-XDR/XDR-TB patients primed CD4<sup>+</sup> that expressed high levels of Th1 effector cytokines. PBMCs were co-cultured with XDR-TB patient-derived, matured mo-DCs (<span class="html-italic">n</span> = 9) for seven days. The single (IFN-γ, TNF-α, and IL-2) and polyfunctional (IFN-γ/IL-2; IFN-γ/TNF-α; IL-2 /TNF-α and IFN-γ/IL-2/TNF-α) cytokine secretion from the DC primed CD4<sup>+</sup> T-cells, matured with either the ECAT peptide pool (<b>A</b>), the PE/PPE peptide pool (<b>B</b>) or the <span class="html-italic">M. tb</span> HN878 lysate (<b>C</b>) was determined by flow cytometry. The flow cytometry data were analysed using FlowJo. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as * and ** for <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 6
<p>Matured mo-DCs from pre-XDR/XDR-TB patients primed CD8<sup>+</sup> T-cells that expressed high levels of cytolytic markers. PBMCs were co-cultured with XDR-TB patient-derived, matured mo-DCs (<span class="html-italic">n</span> = 9) for seven days. The expression levels of granulysin and perforin from the DC-primed CD8<sup>+</sup> T-cells, matured with either the ECAT peptide pool (<b>A</b>), the PE/PPE peptide pool (<b>B</b>) or the <span class="html-italic">M. tb</span> HN878 lysate (<b>C</b>) was determined by flow cytometry. The flow cytometry data were analysed using FlowJo. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as * and ** for <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 7
<p>Matured mo-DCs from pre-XDR/XDR-TB patients primed effector cells that secreted high levels of soluble cytokines. PBMCs were co-cultured with XDR-TB patient-derived, matured mo-DCs (<span class="html-italic">n</span> = 9) for seven days. The experimental conditions are immature DC, negative control (DCs matured without <span class="html-italic">M. tb</span> antigens or cocktail), mo-DCs matured with ECAT only, mo-DCs matured with PE/PPE only, mo-DCs matured with HN878 only, mo-DCs matured with limited cocktail only, mo-DCs matured with ECAT + C, mo-DCs matured with PE/PPE + C, and mo-DCs matured with HN878 + C. The cell culture supernatants were analysed for the expression of the TNF-α (<b>A</b>)<b>,</b> IL-6 (<b>B</b>), IL-13 (<b>C</b>), and RANTES (<b>D</b>) using a Miliplex Luminex assay (Merck, Germany). Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as *, **, and *** for <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.005 respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 8
<p>The TCRs of CD8<sup>+</sup> T-cells primed with matured mo-DCs can recognize ESAT-6 and Ag85B tetramers. Autologous PBMCs were co-cultured with pre-XDR/XDR-TB patient-derived, matured mo-DCs (<span class="html-italic">n</span> = 4) for seven days to generate effector cells. The experimental conditions are immature DC, negative control (DCs matured without <span class="html-italic">M. tb</span> antigens or cocktail), mo-DCs matured with ECAT only, mo-DCs matured with HN878 only, mo-DCs matured with ECAT + C, and mo-DCs matured with HN878 + C. The CD8<sup>+</sup> T-cells were stained for ESAT-6- (<b>A</b>) and Ag85B (<b>B</b>)-specific tetramers on day seven. Flow cytometry data were analysed using FlowJo. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as * and ** for <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure 9
<p>In pre-XDR/XDR-TB patients, the effector cells primed by matured mo-DCs were bactericidal to <span class="html-italic">M. tb</span> in vitro. <span class="html-italic">M. tb</span>-infected MDMs were incubated with/without DC-primed effector cells for 24 h (<span class="html-italic">n</span> = 14). Colony-forming units/mL were determined on Middlebrook H9/OADC agar. Mycobacterial containment is shown as a percentage and was determined relative to reference control. The dotted line represents a relative level of containment by reference control (<span class="html-italic">M. tb</span>-infected MDMs only, i.e., no <span class="html-italic">M. tb</span> containment). The grouped containment response from DC matured with <span class="html-italic">M. tb</span> antigen is shown in (<b>A</b>) and the containment response for all experimental conditions is shown in (<b>B</b>). The experimental conditions are immature DC, negative control (DCs matured without <span class="html-italic">M. tb</span> antigens or cocktail), mo-DCs matured with ECAT only, mo-DCs matured with PE/PPE only, mo-DCs matured with HN878 only, mo-DCs matured with limited cocktail only, mo-DCs matured with ECAT + C, mo-DCs matured with PE/PPE + C, and mo-DCs matured with HN878 + C. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as *, **, and *** for <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.005 respectively. Outliers are represented by symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure A1
<p>The recruitment and sample schematic for participants with latent TB infection (LTBI). Eighteen individuals with latent TB infection were asked to consent to the study. All participants met the inclusion criteria, were healthy upon consent, and samples for 18 participants were collected. Twelve LTBI participants were used for the phenotypic assessment of mature mo-DCs using flow cytometry, the determination of soluble cytokine secretion by mature mo-DCs using ELISA, and the <span class="html-italic">M. tb</span> containment assay. Six LTBI participants were used for the characterization of the DC-primed CD4<sup>+</sup> and CD8<sup>+</sup> T-cell responses.</p>
Full article ">Figure A2
<p>The gating strategy used to identify the monocytic-derived DCs for assessment of their maturation phenotype. The lymphocyte population was gated according to SSC-A and FSC-A (<b>A</b>). Only single cells were included (<b>B</b>); monocytic-derived DCs were identified by HLA-DR+.</p>
Full article ">Figure A3
<p>Gating strategy used for identifying the DC-primed CD4<sup>+</sup> and CD8<sup>+</sup> T-cells for cytokine and cytotoxic assessment. The lymphocyte population was gated on SSC-A and FSC-A (<b>A</b>). The doublets (<b>B</b>) and dead cells (<b>C</b>) were removed from selection, and single cell T-cells were identified according to CD4<sup>+</sup> and CD8<sup>+</sup> surface marker (<b>D</b>). Boolean combination gating was applied for identifying polyfunctional effector cells.</p>
Full article ">Figure A4
<p>Matured mo-DCs from LTBI participants expressed high levels of co-stimulatory molecules (<b>A</b>–<b>D</b>). Immature DCs (<span class="html-italic">n</span> = 12) were differentiated from monocytes in the presence of IL-4 and GM-CSF for five days. After five days, the immature DCs were matured with <span class="html-italic">M. tb</span> antigens and with/without the maturation cocktail for 48 h. The mo-DCs were characterised for the expression of the co-stimulatory markers by flow cytometry. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Outliers are represented symbols, which indicate data points that fall significantly outside the interquartile range. Significance levels are indicated as *, *** and **** for <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.005 and p&lt;0.0001 respectively. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure A5
<p>Matured mo-DCs from LBTI participants expressed high levels of IL-12p70 and polarized a dominant Th1 response (<b>A</b>–<b>C</b>). Immature DCs (<span class="html-italic">n</span> = 6) were differentiated from monocytes in the presence of IL-4 and GM-CSF for five days. After five days, the immature DCs were <span class="html-italic">M. tb</span> antigens, with/without the maturation cocktail for 48 h. IL-12p70 and IL-10 from the culture supernatants were determined using the ELISAPRO IL-12p70 and IL-10 detection kit from Mabtech. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as *, *** and **** for <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.005, and <span class="html-italic">p</span> &lt; 0.001, respectively. Outliers are represented symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure A6
<p>In LTBI, effector cells primed by matured mo-DC did not mediate a containment against <span class="html-italic">M. tb</span> in vitro. Effector cells, primed by mo-DCs matured with ECAT + C were shown to mediate a better containment response compared to effector cells primed by mo-DCs matured with PE/PPE + C. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as * for <span class="html-italic">p</span> &lt; 0.05, respectively. Outliers are represented symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure A7
<p>Matured mo-DCs from pre-XDR/XDR-TB patients expressed high levels of PD-L1 compared to peptide pool and lysate-only matured DCs. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as *** and **** for <span class="html-italic">p</span> &lt; 0.005 and <span class="html-italic">p</span> &lt; 0.0001 respectively. Outliers are represented symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L, Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">Figure A8
<p>Matured mo-DCs from XDR-TB patients primed effector T-cells that expressed high levels of CD69 and PD-1. PBMCs were co-cultured with pre-XDR/XDR-TB patient-derived DCs (<span class="html-italic">n</span> = 9), matured with <span class="html-italic">M. tb</span> antigens, with/without the maturation cocktail for 48 h. The T-cells were stained for immunofluorescent characterization on day seven. Flow cytometry data were analysed using FlowJo. Statistical analysis was performed using GraphPad Prism. One-way ANOVA with Dunnett’s post-test was used to compare experimental groups to the control group (immature mo-DCs). For paired comparisons between mo-DCs matured with <span class="html-italic">M. tb</span> antigens only and mo-DCs with <span class="html-italic">M. tb</span> antigens + C, the Wilcoxon signed-rank test was applied. Significance levels are indicated as * and ** for <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively. Outliers are represented symbols, which indicate data points that fall significantly outside the interquartile range. Abbreviations: ECAT = peptide pool consisting of ESAT-6, CFP-10, AG85b, and TB10.4 peptides; PE/PPE = peptide pool consisting of PE and PPE peptides; HN878 = sonicated lysate of <span class="html-italic">M. tb</span> HN878; LC = limited maturation cocktail containing IFN-γ, IFN-α, IL1-β, and CD40L; C = full maturation cocktail containing IFN-γ, IFN-α, IL1-β, CD40L and Ampligen (TLR3 agonist), and R848 (TLR7/8 agonist).</p>
Full article ">
18 pages, 4044 KiB  
Article
The Effects of Fisetin on Gene Expression Profile and Cellular Metabolism in IFN-γ-Stimulated Macrophage Inflammation
by Ziyu He, Xuchi Pan, Kun Xie, Kozue Sakao, Jihua Chen, Masaharu Komatsu and De-Xing Hou
Antioxidants 2025, 14(2), 182; https://doi.org/10.3390/antiox14020182 - 4 Feb 2025
Viewed by 492
Abstract
Although interferon-gamma (IFN-γ) is known as a critical factor in polarizing macrophages into the pro-inflammatory state for immune response, how dietary flavonoids regulate IFN-γ response for anti-inflammation is incompletely elucidated. This study aims to investigate the effect of fisetin, a typical flavonol, on [...] Read more.
Although interferon-gamma (IFN-γ) is known as a critical factor in polarizing macrophages into the pro-inflammatory state for immune response, how dietary flavonoids regulate IFN-γ response for anti-inflammation is incompletely elucidated. This study aims to investigate the effect of fisetin, a typical flavonol, on the inhibition of IFN-γ-induced inflammation by RNA sequencing (RNA-Seq) and cellular metabolism analysis. RAW264 macrophages pretreated with fisetin following IFN-γ stimulation were subjected to RNA-Seq to analyze alterations in gene expression. Cellular signaling and transcription were investigated using enrichment analysis, motif analysis, and transcription factor prediction. Cellular metabolic state was assessed by measuring the oxygen consumption rate (OCR) and lactate level to reflect mitochondrial respiration and glycolysis. Alterations in signaling proteins were confirmed by Western blot. The results revealed that fisetin downregulated the IFN-γ-induced expression of pro-inflammatory genes and M1 marker genes such as Cxcl9, Il6, Cd80, Cd86, and Nos2. In cellular metabolism, fisetin upregulated the oxidative phosphorylation (OXPHOS) pathway, restored impaired OCR, and reduced lactate production induced by IFN-γ. Motif analysis suggested that fisetin suppressed the activation of IFN-regulatory factor 1 (IRF1). Western blot data further confirmed that fisetin inhibited the phosphorylation of Jak1, Jak2, and STAT1, and decreased the nuclear accumulation of phosphorylated STAT1 and IRF1 induced by IFN-γ. Taken together, our data revealed that fisetin is a potent flavonoid that attenuates IFN-γ-stimulated murine macrophage inflammation and ameliorates disrupted cellular metabolism with a possible Jak1/2-STAT1-IRF1 pathway. Full article
Show Figures

Figure 1

Figure 1
<p>Fisetin altered the gene expression profile in interferon-gamma (IFN)-γ-stimulated macrophages. (<b>A</b>) Principal component analysis (PCA) of RNA sequencing (RNA-Seq) data showed that the gene expression profile with or without fisetin pretreatment in IFN-γ-stimulated macrophages formed distinct clusters. (<b>B</b>) Volcano plot of the differentially expressed genes (DEGs) in IFN-γ-stimulated cells compared to non-treated cells (IFN-γ/CON). (<b>C</b>) Volcano plot of the DEGs in fisetin-pretreated IFN-γ-stimulated cells compared to IFN-γ-stimulated cells (FisIFN-γ/IFN-γ). RAW264 cells were pre-cultured for 21 h and starved in serum-free medium for 2.5 h. The cells were then treated with or without 5 μM fisetin for 30 min and subsequently exposed to 10 ng/mL IFN-γ for 12 h. Total RNA was isolated for RNA-Seq. Fold change (FC) &gt; 1.3 or &lt; 0.77 were identified as DEGs with <span class="html-italic">q</span>-value &lt; 0.05. Three biological replicates were used for each group.</p>
Full article ">Figure 2
<p>Fisetin inhibited IFN-γ-induced inflammation. (<b>A</b>,<b>B</b>) Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis of upregulated DEGs in IFN-γ-stimulated cells compared to non-treated cells (IFN-γ/CON) (<b>A</b>) and downregulated DEGs in fisetin-pretreated IFN-γ-stimulated cells compared to IFN-γ-stimulated cells (FisIFN-γ/IFN-γ) (<b>B</b>). Red and blue terms represent that the pathway was upregulated by IFN-γ, but downregulated by fisetin pretreatment. (<b>C</b>,<b>D</b>) Gene Set Enrichment Analysis (GSEA) plots of the total altered genes in interferon gamma response gene set in IFN-γ-stimulated cells compared to control without treatment (<b>C</b>) and in fisetin-pretreated IFN-γ-stimulated cells compared to IFN-γ-stimulated cells (<b>D</b>). (<b>E</b>,<b>F</b>) GSEA plots of the total altered genes in inflammatory response gene set in IFN-γ-stimulated cells compared to control without treatment (<b>E</b>) and in fisetin-pretreated IFN-γ-stimulated cells compared to IFN-γ-stimulated cells (<b>F</b>). The gene sets were from mouse-ortholog hallmark gene sets of the molecular signatures database. (<b>G</b>–<b>I</b>) Expression heatmaps of representative genes from interferon gamma response gene set (<b>G</b>), representative pro-inflammatory and anti-inflammatory genes (<b>H</b>), as well as representative genes from macrophages M1 and M2 markers (<b>I</b>). Color gradient reflects row Z-score.</p>
Full article ">Figure 3
<p>Fisetin modulated metabolism in IFN-γ-stimulated macrophages. (<b>A</b>,<b>B</b>) GSEA plot of oxidative phosphorylation (OXPHOS) gene set in IFN-γ-stimulated cells compared to non-treated cells (<b>A</b>) and in fisetin-pretreated IFN-γ-stimulated cells compared to IFN-γ-stimulated cells (<b>B</b>). The gene set was from mouse-ortholog hallmark gene sets of the molecular signatures database. (<b>C</b>) Expression heatmap of representative genes from OXPHOS gene set. Color gradient reflects row Z-score. (<b>D</b>) Fisetin rescued impaired basal oxygen consumption rate (OCR) in IFN-γ-induced macrophages (<span class="html-italic">n</span> = 6). (<b>E</b>) Fisetin decreased IFN-γ-enhanced lactate production (<span class="html-italic">n</span> = 3). For (<b>D</b>,<b>E</b>), RAW264 cells were pre-cultured for 21 h and starved in serum-free medium for 2.5 h. The cells were then treated with 0–5 μM fisetin for 30 min and subsequently exposed to 10 ng/mL IFN-γ for 12 h. The cells were used for OCR measurement, and the culture medium was collected for lactate determination. Each value represents the mean ± SD; different letters between groups indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). For OCR and lactate determination, at least two separate experiments were performed.</p>
Full article ">Figure 4
<p>Motif ISRE and IRF1 presented in the promoters of fisetin-downregulated DEGs. (<b>A</b>) Homer known motif enrichment results of IFN-γ-upregulated DEGs compared to non-treated cells (IFN-γ/CON, UP). Top 11 motifs were shown. (<b>B</b>) Homer known motif enrichment results of fisetin-downregulated DEGs compared to IFN-γ-stimulated cells (FisIFN-γ/IFN-γ, DOWN). Top 11 motifs were shown. (<b>C</b>) Homer <span class="html-italic">de novo</span> motif enrichment results of fisetin-downregulated DEGs compared to IFN-γ-stimulated cells (FisIFN-γ/IFN-γ, DOWN). The generated 5 motifs were shown. The motifs colored in red are enriched in both IFN-γ-upregulated and fisetin-downregulated DEGs.</p>
Full article ">Figure 5
<p>Fisetin-downregulated genes were associated with IFN-regulatory factor1 (IRF1). (<b>A</b>,<b>B</b>) Transcription factor inference by epigenetic landscape in silico deletion analysis (LISA) on the top 500 DEGs that were upregulated (Vertical axis) or downregulated (Horizontal axis) by IFN-γ, compared to non-treated cells (<b>A</b>), and by fisetin pretreatment, compared to IFN-γ-stimulated cells (<b>B</b>). (<b>C</b>–<b>E</b>) Expression heatmaps of the representative genes targeted by STAT1 (<b>C</b>), IRF1 (<b>D</b>), or both STAT1 and IRF1 (<b>E</b>). Color gradient reflects row Z-score.</p>
Full article ">Figure 6
<p>Fisetin suppressed IFN-γ-induced pro-inflammatory mediators at protein levels. (<b>A</b>) Fisetin repressed IFN-γ-induced inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) expression in a dose-dependent manner. (<b>B</b>) Fisetin dose-dependently inhibited IFN-γ-induced nitric oxide (NO) production (<span class="html-italic">n</span> = 3). RAW264 cells were pre-cultured for 21 h and starved in serum-free medium for 2.5 h. The cells were treated with 0–20 μM fisetin for 30 min and then exposed to 10 ng/mL IFN-γ for 12 h. Whole-cell lysates were harvested for Western blot assay, and the culture medium was collected for NO determination. The relative density was calculated as the intensity of the treatment relative to that of the control normalized to α-tubulin by densitometry. Each value represents the mean ± SD; different letters between groups indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). The blots presented are representatives from at least three independent experiments, using cells derived from at least two separate preparations. For NO determination, at least two separate experiments were performed.</p>
Full article ">Figure 7
<p>IFN-γ did not activate MAPK pathway. (<b>A</b>,<b>B</b>) Time-course experiment showed that LPS activated MKK4-JNK-AP-1 cascade with no effect on Jak2-STAT1-IRF1 pathway (<b>A</b>), while IFN-γ activated Jak2-STAT1-IRF1 pathway with no effect on MKK4-JNK-AP-1 cascade (<b>B</b>). (<b>C</b>) IFN-γ did not activate AP-1 (p-c-Jun). RAW264 cells were pre-cultured as described above. The cells were then treated with or without 20 μM fisetin for 30 min and subsequently exposed to 40 ng/mL LPS or 10 ng/mL IFN-γ. Whole-cell lysates were harvested after a defined stimulation time (<b>A</b>,<b>B</b>) or 30 min (for c-Jun and p-c-Jun in (<b>C</b>)), and analyzed by Western blot assay. The relative density was calculated as the intensity of the treatment relative to that of the control normalized to α-tubulin or respective total proteins by densitometry. The bands of total proteins for (<b>A</b>,<b>B</b>) and semi-quantitative graphs are given in <a href="#app1-antioxidants-14-00182" class="html-app">Supplementary Figure S1</a>. Each value represents the mean ± SD; different letters between groups indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). The blots presented are representatives from at least three independent experiments, using cells derived from at least two separate preparations.</p>
Full article ">Figure 8
<p>Fisetin inhibited Jak1/2-STAT1-IRF1 pathway. (<b>A</b>) Fisetin inhibited phosphorylation of Jak1, Jak2, and STAT1, and the expression of IRF1 induced by IFN-γ. (<b>B</b>) Fisetin reduced nuclear accumulation of p-STAT1 and IRF1 in a dose-dependent manner. RAW264 cells were pre-cultured as described above. The cells were then treated with 0–20 μM fisetin for 30 min and subsequently exposed to 10 ng/mL IFN-γ. Whole-cell lysates were harvested after 30 min (for p-Jak1, Jak1, p-Jak2, Jak2, p-STAT1, and STAT1 in (<b>A</b>)) or 2 h (for IRF1 in (<b>A</b>,<b>B</b>)), and then analyzed by Western blot assay. Nuclear and cytoplasmic fractionation was performed as described in Materials and Methods. The relative density was calculated as the intensity of the treatment relative to that of the control normalized to respective total proteins, α-tubulin (for cytoplasmic fraction), or TBP (for nuclear fraction) by densitometry. Each value represents the mean ± SD; different letters between groups indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). The blots presented are representatives from at least three independent experiments, using cells derived from at least two separate preparations.</p>
Full article ">
19 pages, 11686 KiB  
Article
Cross-Talk Between Tumor Cells and Stellate Cells Promotes Oncolytic VSV Activity in Intrahepatic Cholangiocarcinoma
by Victoria Neumeyer, Purva Chavan, Katja Steiger, Oliver Ebert and Jennifer Altomonte
Cancers 2025, 17(3), 514; https://doi.org/10.3390/cancers17030514 - 4 Feb 2025
Viewed by 301
Abstract
As the mechanisms underlying tumorigenesis become better understood, the dynamic roles of cellular components of the tumor microenvironment, and their cross-talk with tumor cells, have come to light as key drivers of disease progression and have emerged as important targets of new cancer [...] Read more.
As the mechanisms underlying tumorigenesis become better understood, the dynamic roles of cellular components of the tumor microenvironment, and their cross-talk with tumor cells, have come to light as key drivers of disease progression and have emerged as important targets of new cancer therapies. In the field of oncolytic virus (OV) therapy, stromal cells have been considered as potential barriers to viral spread, thus limiting virus replication and therapeutic outcome. However, new evidence indicates that intratumoral fibroblasts could support virus replication. We have demonstrated in a rat model of stromal-rich intrahepatic cholangiocarcinoma (CCA) that vesicular stomatitis virus (VSV) can be localized within intratumoral hepatic stellate cells (HSCs), in addition to tumor cells, when the virus was applied via hepatic arterial infusion. Furthermore, VSV was shown to efficiently kill CCA cells and activated HSCs, and co-culture of CCA and HSCs increased viral titers. Interestingly, this effect is also observed when each cell type is cultured alone in a conditioned medium of the other cell type, indicating that secreted cell factors are at least partially responsible for this phenomenon. Partial reduction in sensitivity to type I interferons was observed in co-culture systems, providing a possible mechanism for the increased viral titers. Together, the results indicate that targeting activated HSCs with VSV could provide an additional mechanism of OV therapy, which, until now has not been considered. Furthermore, these findings suggest that VSV is a potentially powerful therapeutic agent for stromal-rich tumors, such as CCA and pancreatic cancer, both of which are very difficult to treat with conventional therapy and have a very poor prognosis. Full article
(This article belongs to the Special Issue The Role of Viruses in the Development of Cancer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Crosstalk between CCA cells and HSCs leads to differential gene expression and HSC activation. (<b>A</b>) Human CCA cell lines, RBE and HuCCT1, and primary human hepatic stellate cells (HSCs) were either cultured alone or as co-culture at a ratio of 1:1. Expression of TGF-β, αSMA, and TIMP-1 were analyzed by quantitative real-time RT-PCR and normalized to GAPDH. Mean values from three independent experiments are shown, and error bars indicate SEM. Statistical significance was determined by Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001). (<b>B</b>) Primary human HSCs were cultured in their own medium or medium conditioned by RBE or HuCCT1 cells for 48 h. Immunofluorescence staining for αSMA (red) and nuclei (DAPI, blue) was performed. Pictures were taken with a fluorescence microscope at a magnification of 200×.</p>
Full article ">Figure 2
<p>Human CCA cells are susceptible to VSV-GFP expression. Virus replication was monitored after infection of (<b>A</b>) HuCCT1 and (<b>D</b>) RBE cells at MOI 10 and 0.01 at several time points (0, 6, 12, 24, and 48 h post-infection). Viral titers were determined by TCID<sub>50</sub> assay. Cell viability of (<b>B</b>) HuCCT1 was measured by MTS assay at 24, 48, and 72 h post-infection and (<b>E</b>) RBE cells at 24 and 48 h. Representative pictures of uninfected and VSV-GFP-infected (<b>C</b>) HuCCT1 and (<b>F</b>) RBE cells at 48 h post-infection are shown in bright fields (top) and fluorescence for GFP visualization (bottom). The scale bar indicates 100 µm. Mean values from three independent experiments are shown, and error bars indicate SEM.</p>
Full article ">Figure 3
<p>Crosstalk between CCA cells and HSCs enhances viral replication and cytotoxicity. (<b>A</b>) Viral titers were determined 24 h post-infection by TCID<sub>50</sub> assay after co-culture of HuCCT1 and RBE cells with HSCs at a ratio of 1:1. (<b>B</b>) Cytotoxicity was measured by LDH release upon infection of HuCCT1 and RBE cells with VSV-GFP at MOI 0.01 for 24 h in co-culture with LX2 cells. Values were normalized to a maximum release control. TCID<sub>50</sub> assay of HuCCT1 and RBE cells cultured in (<b>C</b>) primary HSC or (<b>D</b>) LX2 conditioned media were measured at 48 h post-infection. (<b>E</b>) Cytotoxicity was measured as a function of LDH release from infected HuCCT1 and RBE cells cultured in own or LX2-conditioned medium at 48 h post-infection at MOI 0.01. Values were normalized to a maximum release control. (<b>F</b>) Representative images of HuCCT1 cells infected with VSV-GFP at MOI 0.01 48 h post-infection. The scale bar indicates 100 µm. (<b>G</b>) Viral titers measured by TCID<sub>50</sub> assay from VSV-GFP-infected primary HSCs and LX2 cells 48 h post-infection. (<b>H</b>) Cytotoxicity was measured by LDH release assay upon infection of LX2 cells with VSV-GFP after 48 h. Values were normalized to a maximum release control. (<b>I</b>) Representative photomicrographs of LX2 cells infected in own, HuCCT1, or RBE conditioned media with VSV-GFP at MOI 0.01 at 48 h post-infection are shown. The scale bar indicates 100 µm. Mean values from three independent experiments are shown, and error bars indicate SEM. Statistical significance was determined by Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>Crosstalk between CCA cells and HSCs dampens IFN signaling and response pathways. (<b>A</b>) IFNβ and (<b>B</b>) Interferon-stimulated response element (ISRE) promoter activation were measured in HuCCT1, RBE, and primary human stellate cells (HSCs) after overnight infection with VSV or VSV(MΔ51) or stimulation with poly I:C or universal type-I IFN, respectively, using luciferase reporter plasmids and the Dual-Luciferase Reporter assay. Values were normalized to control the luciferase signal and are shown as fold-induction compared to uninfected controls. (<b>C</b>) HuCCT1, RBE, and HSC cells were treated with type-I IFN overnight prior to infection with VSV at MOI 1. Viral titers were measured 18 h post-infection using TCID<sub>50</sub> assay. Mean values from three independent experiments are shown, and error bars indicate SEM. Statistical significance was determined by Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>VSV replicates in tumor cells and HSCs in a rat model of CCA. Intrahepatic CCA was induced in male rats by long-term treatment with thioacetamide. Representative photomicrographs of (<b>A</b>) hematoxylin-eosin for histology analysis and (<b>B</b>) Elastica van Gieson staining for collagen (stained pink) of intrahepatic CCA sampled 24 h after treatment with PBS by hepatic arterial infusion. The scale bar indicates 100 µm. (<b>C</b>) Virus titers were measured by TCID<sub>50</sub> assay from lysates of the tumor and healthy liver tissue was isolated and snap-frozen 24 h after treatment. Mean values from four individual animals are shown; error bars indicate SEM. (<b>D</b>) Representative images of immunohistochemical staining of intrahepatic CCA tumors stained for VSV-M (red) or (<b>E</b>) immunofluorescent staining of α-SMA (green) and VSV-M (red) 1 day after treatment with PBS or VSV.</p>
Full article ">Figure 6
<p>VSV reduces fibrosis in a rat model of CCA. CCA tumors were induced in rats by long-term thioacetamide treatment in drinking water, and animals were treated with PBS or VSV by intrahepatic arterial infusion. Expression of (<b>A</b>) α-SMA, (<b>B</b>) TGF-β, and (<b>C</b>) collagen was analyzed by RT-qPCR 1-day post-treatment. mRNA levels were normalized to GAPDH and are depicted as fold-change compared to PBS-treated controls. (<b>D</b>) Intratumoral fibrotic content was quantified by analysis of pink-stained collagen fibers from the Elastica van Gieson staining of VSV- or PBS-treated CCA tumors. Mean values from four animals and representative photomicrographs are shown; error bars indicate SEM. Statistical significance was determined by Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
44 pages, 5541 KiB  
Review
Harnessing Epigenetics: Innovative Approaches in Diagnosing and Combating Viral Acute Respiratory Infections
by Ankita Saha, Anirban Ganguly, Anoop Kumar, Nityanand Srivastava and Rajiv Pathak
Pathogens 2025, 14(2), 129; https://doi.org/10.3390/pathogens14020129 - 1 Feb 2025
Viewed by 521
Abstract
Acute respiratory infections (ARIs) caused by viruses such as SARS-CoV-2, influenza viruses, and respiratory syncytial virus (RSV), pose significant global health challenges, particularly for the elderly and immunocompromised individuals. Substantial evidence indicates that acute viral infections can manipulate the host’s epigenome through mechanisms [...] Read more.
Acute respiratory infections (ARIs) caused by viruses such as SARS-CoV-2, influenza viruses, and respiratory syncytial virus (RSV), pose significant global health challenges, particularly for the elderly and immunocompromised individuals. Substantial evidence indicates that acute viral infections can manipulate the host’s epigenome through mechanisms like DNA methylation and histone modifications as part of the immune response. These epigenetic alterations can persist beyond the acute phase, influencing long-term immunity and susceptibility to subsequent infections. Post-infection modulation of the host epigenome may help distinguish infected from uninfected individuals and predict disease severity. Understanding these interactions is crucial for developing effective treatments and preventive strategies for viral ARIs. This review highlights the critical role of epigenetic modifications following viral ARIs in regulating the host’s innate immune defense mechanisms. We discuss the implications of these modifications for diagnosing, preventing, and treating viral infections, contributing to the advancement of precision medicine. Recent studies have identified specific epigenetic changes, such as hypermethylation of interferon-stimulated genes in severe COVID-19 cases, which could serve as biomarkers for early detection and disease progression. Additionally, epigenetic therapies, including inhibitors of DNA methyltransferases and histone deacetylases, show promise in modulating the immune response and improving patient outcomes. Overall, this review provides valuable insights into the epigenetic landscape of viral ARIs, extending beyond traditional genetic perspectives. These insights are essential for advancing diagnostic techniques and developing innovative treatments to address the growing threat of emerging viruses causing ARIs globally. Full article
(This article belongs to the Special Issue The Epidemiology and Diagnosis of Acute Respiratory Infections)
Show Figures

Figure 1

Figure 1
<p><b>Epigenetic regulation of innate immune defense in response to respiratory viral infections</b>: This schematic illustrates the mechanisms of innate immune defense pathways and their regulation by epigenetic modifications during acute respiratory viral infections. Toll-like receptors (TLRs) at the cell membrane and within endosomes detect pathogen-associated molecular patterns (PAMPs), with TLR2/4 responding to viral proteins and endosomal TLR3, TLR7/8, and TLR9 recognizing viral dsRNA, ssRNA, and CpG DNA, respectively. These receptors activate adaptor proteins such as MyD88 and TRIF, triggering NF-κB and IRF3/7 signaling pathways that drive the production of pro-inflammatory cytokines and type I interferons (IFN-α/β). In the cytosol, RIG-I-like receptors (RLRs), including RIG-I and MDA5, detect viral RNA and signal through mitochondrial antiviral signaling protein (MAVS) to activate NF-κB and IRFs. NOD-like receptors (NLRs) such as NOD1 and NOD2 sense bacterial and viral components, promoting NF-κB activation and RNA degradation via OAS2 and RNase L, while absent in melanoma 2-like receptors (ALRs) like AIM2 detect viral dsDNA, forming inflammasomes that activate IL-1β and IL-18. Epigenetic regulation in the nucleus modulates these immune responses, with DNA methylation suppressing PRR gene expression, histone acetylation (e.g., H3K27ac) enhancing gene expression, and miRNAs interfering with pathways like NF-κB (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>).</p>
Full article ">Figure 2
<p><b>Impact of respiratory viral infections on epigenetic modifications and therapeutic interventions:</b> This figure illustrates how respiratory viral infections influence key epigenetic modifications, including histone acetylation, histone methylation, DNA methylation, and RNA-based transcriptional regulation, alongside potential therapeutic strategies targeting these pathways. Viral infections modulate histone acetylation within the nucleus (via histone acetyltransferases [HATs] and histone deacetylases [HDACs]), with acetylation promoting transcriptional activation and deacetylation suppressing antiviral responses. They also alter histone methylation through enzymes like PRC1/PRC2 and SET/SuVAR, where repressive marks (e.g., H3K27me3) suppress antiviral genes, and active marks (e.g., H3K4me3) are dysregulated during viral infections. DNA methyltransferases (DNMTs) silence promoter regions of antiviral genes, while TET enzymes regulate DNA demethylation to restore immune gene expression. RNA-based regulation, including miRNA disruption and lncRNA dysregulation, interferes with transcriptional responses, while excessive NETosis exacerbates inflammation. Therapeutic interventions include epigenetic modulators such as HAT inhibitors (e.g., curcumin), HDAC inhibitors (e.g., metformin, statins), DNMT inhibitors (e.g., azacitidine), and direct antiviral treatments (e.g., monoclonal antibodies, CRISPR-based therapies, remdesivir) (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>).</p>
Full article ">
20 pages, 5926 KiB  
Article
Crosstalk Between nNOS/NO and COX-2 Enhances Interferon-Gamma-Stimulated Melanoma Progression
by Anika Patel, Shirley Tong, Moom R. Roosan, Basir Syed, Amardeep Awasthi, Richard B. Silverman and Sun Yang
Cancers 2025, 17(3), 477; https://doi.org/10.3390/cancers17030477 - 31 Jan 2025
Viewed by 443
Abstract
Background/Objectives: Interferon gamma (IFN-γ) in the melanoma tumor microenvironment plays opposing roles, orchestrating both pro-tumorigenic activity and anticancer immune responses. Our previous studies demonstrated the role of neuronal nitric oxide synthase (nNOS) in IFN-γ-stimulated melanoma progression. However, the underlying mechanism has not been [...] Read more.
Background/Objectives: Interferon gamma (IFN-γ) in the melanoma tumor microenvironment plays opposing roles, orchestrating both pro-tumorigenic activity and anticancer immune responses. Our previous studies demonstrated the role of neuronal nitric oxide synthase (nNOS) in IFN-γ-stimulated melanoma progression. However, the underlying mechanism has not been well defined. This study determined whether the nNOS/NO and COX-2/PGE2 signaling pathways crosstalk and augment the pro-tumorigenic effects of IFN-γ in melanoma. Methods: Bioinformatic analysis of patient and cellular proteomic data was conducted to identify proteins of interest associated with IFN-γ treatment in melanoma. Changes in protein expression were determined using various analytical techniques including western blot, flow cytometry, and confocal microscopy. The levels of PGE2 and nitric oxide (NO) were analyzed by HPLC chromatography and flow cytometry. In vivo antitumor efficacy was determined utilizing a human melanoma xenograft mouse model. Results: Our omics analyses revealed that the induction of COX-2 was significantly predictive of IFN-γ treatment in melanoma cells. In the presence of IFN-γ, PGE2 further enhanced PD-L1 expression and amplified the induction of nNOS, which increased intracellular NO levels. Cotreatment with celecoxib effectively diminished these changes induced by PGE2. In addition, nNOS blockade using a selective small molecule inhibitor (HH044), efficiently inhibited IFN-γ-induced PGE2 and COX-2 expression levels in melanoma cells. STAT3 inhibitor napabucasin also inhibited COX-2 expression both in the presence and absence of IFN-γ. Furthermore, celecoxib was shown to enhance HH044 cytotoxicity in vitro and effectively inhibit human melanoma tumor growth in vivo. HH044 treatment also significantly reduced tumor PGE2 levels in vivo. Conclusions: Our study demonstrates the positive feedback loop linking nNOS-mediated NO signaling to the COX-2/PGE2 signaling axis in melanoma, which further potentiates the pro-tumorigenic activity of IFN-γ. Full article
Show Figures

Figure 1

Figure 1
<p>Differentially expressed proteins that can predict IFN-γ treatment in melanoma. (<b>a</b>) Reverse Phase Protein Array data were used to identify proteins with significant changes in expression after IFN-γ (250 units/mL) treatment compared to untreated cells across three human melanoma cell lines. The color scale indicates log2 median-centered intensity in protein expression for each cell line. The top 35 genes used to predict IFN-γ treatment are shown. (<b>b</b>) The Cancer Proteome Atlas Level 4 PD-L1 protein expression data for a total of 110 patients who had skin cutaneous melanoma (SKCM) were analyzed to compare patients with advanced disease (stage III/IV, <span class="html-italic">n</span> = 45) to stage I/II melanoma patients (<span class="html-italic">n</span> = 65, *, <span class="html-italic">p</span> = 0.03).</p>
Full article ">Figure 2
<p>Effects of PGE<sub>2</sub> and COX-2 inhibition on PD-L1 expression in human melanoma. (<b>a</b>) Immunoblots detected the effects of PGE<sub>2</sub> on PD-L1 expression in human melanoma A375 cells in the presence and absence of IFN-γ. β-actin was utilized to normalize loaded protein. The bar graph represents the 24 h results (<span class="html-italic">n</span> = 3). Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S3</a>. (<b>b</b>,<b>c</b>) Representative flow cytometry and confocal microscopy images of A375 cells treated with PGE<sub>2</sub> with and without IFN-γ for 24 and 72 h, respectively. Cells were stained for expression of PD-L1 (Alexa Fluor 488, green) and nuclei (DAPI, blue). Confocal imaging was taken at 20× magnification; scale bar: 50 µm. (<b>d</b>) Celecoxib cotreatment reduced PD-L1 levels compared to IFN-γ treatment alone. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S4</a>. (<b>e</b>) Representative flow cytometry images of PD-L1 mean fluorescence and (<b>f</b>) images of immunofluorescent staining of PD-L1 by confocal microscopy (60× magnification, scale bar: 50 µm) after celecoxib treatment with and without IFN-γ. Cells were stained for expression of PD-L1 (Alexa Fluor 488, green) and nuclei (DAPI, blue). * <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, ns—not significant.</p>
Full article ">Figure 2 Cont.
<p>Effects of PGE<sub>2</sub> and COX-2 inhibition on PD-L1 expression in human melanoma. (<b>a</b>) Immunoblots detected the effects of PGE<sub>2</sub> on PD-L1 expression in human melanoma A375 cells in the presence and absence of IFN-γ. β-actin was utilized to normalize loaded protein. The bar graph represents the 24 h results (<span class="html-italic">n</span> = 3). Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S3</a>. (<b>b</b>,<b>c</b>) Representative flow cytometry and confocal microscopy images of A375 cells treated with PGE<sub>2</sub> with and without IFN-γ for 24 and 72 h, respectively. Cells were stained for expression of PD-L1 (Alexa Fluor 488, green) and nuclei (DAPI, blue). Confocal imaging was taken at 20× magnification; scale bar: 50 µm. (<b>d</b>) Celecoxib cotreatment reduced PD-L1 levels compared to IFN-γ treatment alone. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S4</a>. (<b>e</b>) Representative flow cytometry images of PD-L1 mean fluorescence and (<b>f</b>) images of immunofluorescent staining of PD-L1 by confocal microscopy (60× magnification, scale bar: 50 µm) after celecoxib treatment with and without IFN-γ. Cells were stained for expression of PD-L1 (Alexa Fluor 488, green) and nuclei (DAPI, blue). * <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, ns—not significant.</p>
Full article ">Figure 3
<p>Effects of PGE<sub>2</sub> and COX-2 inhibition on nNOS expression and NO levels in human melanoma cells. (<b>a</b>) A375 cells were treated with PGE<sub>2</sub> (25 μM) with and without IFN-γ (250 units/mL) for 24 h, followed by immunoblot analysis of nNOS expression. β-actin was used to normalize loaded protein. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S5</a>. (<b>b</b>) Representative flow cytometry images of intracellular NO levels increased by PGE<sub>2</sub> in the presence and absence of IFN-γ as detected using a DAF fluorescence probe. (<b>c</b>) Immunoblot analysis of nNOS expression levels after celecoxib treatment (50 μM) for 48 h. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S6</a>. (<b>d</b>) Celecoxib cotreatment reduced intracellular NO levels in the presence of IFN-γ in A375 cells. The impact of different treatments on NO levels in human melanoma SK-MEL-28 cells is shown in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Figure S10</a>. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns—not significant.</p>
Full article ">Figure 3 Cont.
<p>Effects of PGE<sub>2</sub> and COX-2 inhibition on nNOS expression and NO levels in human melanoma cells. (<b>a</b>) A375 cells were treated with PGE<sub>2</sub> (25 μM) with and without IFN-γ (250 units/mL) for 24 h, followed by immunoblot analysis of nNOS expression. β-actin was used to normalize loaded protein. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S5</a>. (<b>b</b>) Representative flow cytometry images of intracellular NO levels increased by PGE<sub>2</sub> in the presence and absence of IFN-γ as detected using a DAF fluorescence probe. (<b>c</b>) Immunoblot analysis of nNOS expression levels after celecoxib treatment (50 μM) for 48 h. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S6</a>. (<b>d</b>) Celecoxib cotreatment reduced intracellular NO levels in the presence of IFN-γ in A375 cells. The impact of different treatments on NO levels in human melanoma SK-MEL-28 cells is shown in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Figure S10</a>. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns—not significant.</p>
Full article ">Figure 4
<p>Effects of NO stress on COX-2 expression and PGE<sub>2</sub> levels in human melanoma. NO stress significantly induced COX-2 expression (<b>a</b>) and PGE<sub>2</sub> production (<b>b</b>) in melanoma cells. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figures S7 and S8</a>. A375 cells were cultured in serum-free DMEM medium with DetaNONOate 100 μM for 24 and 48 h. After 48 h, the media was collected for analysis of PGE<sub>2</sub> levels using LC-MS/MS. The internal standard PGE<sub>2</sub>-d4 chromatogram is shown in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Figure S1</a>. Cotreatment with nNOS inhibitor HH044 effectively diminished the induction of COX-2 (<b>c</b>) and PGE<sub>2</sub> production (<b>d</b>) by IFN-γ in SK-MEL-28 and A375 cells, respectively. DMEM media was collected after treatment with HH044 20 μM and IFN-γ 250 units/mL for 48 h. The corresponding PGE<sub>2</sub>-d4 chromatogram is shown in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Figure S2</a>. The impact of different treatments on PGE<sub>2</sub> levels in human melanoma SK-MEL-28 cells is shown in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Figure S11</a>. * <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, ns—not significant.</p>
Full article ">Figure 5
<p>Effects of STAT3 inhibitor napabucasin on COX-2 expression in melanoma cells. SK-MEL-28 cells were treated with napabucasin 1 μM with and without IFN-γ 250 units/mL for 72 h and analyzed using immunoblot. Expression levels of COX-2 were normalized via β-actin. ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001. Full-length blots of manuscript is shown in <a href="#app1-cancers-17-00477" class="html-app">Figure S9</a>.</p>
Full article ">Figure 6
<p>Celecoxib enhances the cytotoxicity of nNOS inhibitor HH044 in melanoma cells. IC<sub>50</sub> was determined using GraphPad Prism, detected by MTT colorimetric analysis. A375 cells were treated with various concentrations of HH044 and celecoxib for 72 h, and viable cells were measured by absorbance at 595 nm. All IC<sub>50</sub> values are included in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Table S3</a>.</p>
Full article ">Figure 7
<p>The nNOS inhibitor HH044 decreased tumor PGE<sub>2</sub> levels in vivo. DBA/2 male mice were injected with Cloudman S91 cells to induce tumor growth. Mice were then randomized into different groups (vehicle control, HH044 10 mg/kg i.p., and celecoxib 50 mg/kg p.o. daily for 24 days). At the end of the study, tumors were collected and processed for PGE<sub>2</sub> analysis using LC/MS-MS. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to control.</p>
Full article ">Figure 8
<p>Celecoxib showed potent anti-melanoma activity in vivo. Nude mice were injected with human melanoma A375 cells to induce tumor growth and were treated with celecoxib 50 mg/kg/day p.o. for 23 days. Tumors were measured biweekly. (<b>a</b>) Tumor growth is presented as the fold of tumor volume on Day 0 (5 days post inoculation, and the day before starting treatment) for each individual mouse. The average fold of tumor growth is shown in red for the control (<span class="html-italic">n</span> = 7) and in blue for celecoxib (<span class="html-italic">n</span> = 7). The tumor growth curves of individual mice treated with celecoxib are represented in gray. The tumor growth curve in volume is presented in <a href="#app1-cancers-17-00477" class="html-app">Supplementary Figure S12</a>. (<b>b</b>) Kaplan–Meier curve for overall survival in nude mice treated with control and celecoxib. * <span class="html-italic">p</span> &lt; 0.05 compared to control.</p>
Full article ">Figure 9
<p>A schematic representation of the crosstalk between the nNOS/NO and COX-2/PGE<sub>2</sub> signaling pathways, which enhanced IFN-γ-induced PD-L1 expression in melanoma. IFN-γ has been implicated as a pro-tumorigenic cytokine attributed to melanoma progression. Our study has demonstrated that the nNOS/NO and COX-2/PGE<sub>2</sub> pathways are activated in the presence of IFN-γ, both of which generate proinflammatory molecules. The enzymatic product of COX-2 activity, PGE<sub>2</sub>, induced the expression of nNOS in melanoma cells, while NO produced by nNOS further increased the expression of COX-2. As a result, this feedforward loop amplifies the pro-tumorigenic effects of IFN-γ in melanoma by inducing PD-L1, leading to immune suppression within the tumor microenvironment. Blockade of COX-2 and/or nNOS using selective inhibitors may be a promising approach for melanoma therapy, which effectively alleviates the pro-tumorigenic effects of IFN-γ in melanoma cells without inhibiting the essential immune cell intrinsic function of IFN-γ in tumor immunosurveillance. The blue downward arrows indicate the effect of COX-2 inhibitor celecoxib and green downward arrows indicate the effect of the nNOS inhibitor HH044.</p>
Full article ">
15 pages, 1054 KiB  
Review
Targeted Cellular Treatment of Systemic Lupus Erythematosus
by Panagiotis Athanassiou, Lambros Athanassiou, Ifigenia Kostoglou-Athanassiou and Yehuda Shoenfeld
Cells 2025, 14(3), 210; https://doi.org/10.3390/cells14030210 - 31 Jan 2025
Viewed by 393
Abstract
Systemic lupus erythematosus (SLE) is a systemic autoimmune disease affecting all organ systems. The disease preferentially affects females of childbearing age. It runs a variable course. It may run a mild course that may never lead to severe disease and manifestations from critical [...] Read more.
Systemic lupus erythematosus (SLE) is a systemic autoimmune disease affecting all organ systems. The disease preferentially affects females of childbearing age. It runs a variable course. It may run a mild course that may never lead to severe disease and manifestations from critical organ systems. However, it may also run an undulating course with periods of mild and severe disease. It may run as a mild disease, quickly deteriorating to severe disease and affecting multiple organ systems. Various immune pathways related both to the innate and adaptive immune response are involved in the pathogenesis of SLE. Various drugs have been developed targeting cellular and molecular targets in these pathways. Interferons are involved in the pathogenesis of SLE, and various drugs have been developed to target this pathway. T and B lymphocytes are involved in the pathophysiology of SLE. Various treatment modalities targeting cellular targets are available for the treatment of SLE. These include biologic agents targeting B lymphocytes. However, some patients have disease refractory to these treatment modalities. For these patients, cell-based therapies may be used. Hematopoietic stem cell transplantation involving autologous cells is an option in the treatment of refractory SLE. Mesenchymal stem cells are also applied in the treatment of SLE. Chimeric antigen receptor (CAR)-T cell therapy is a novel treatment also used in SLE management. This novel treatment method holds major promise for the management of autoimmune diseases and, in particular, SLE. Major hurdles to be overcome are the logistics involved, as well as the need for specialized facilities. This review focuses on novel treatment modalities in SLE targeting cellular and molecular targets in the immune system. Full article
(This article belongs to the Special Issue Advances in Cellular and Molecular Treatment of Autoimmune Diseases)
Show Figures

Figure 1

Figure 1
<p>Cell-based treatment modalities for treatment of refractory systemic lupus erythematosus.</p>
Full article ">Figure 2
<p>B cell-targeted treatment modalities in systemic lupus erythematosus.</p>
Full article ">Figure 3
<p>Molecular targets on B lymphocytes.</p>
Full article ">
26 pages, 4444 KiB  
Article
HCoV-229E Mpro Suppresses RLR-Mediated Innate Immune Signalling Through Cleavage of NEMO and Through Other Mechanisms
by Xavier Martiáñez-Vendrell, Puck B. van Kasteren, Sebenzile K. Myeni and Marjolein Kikkert
Int. J. Mol. Sci. 2025, 26(3), 1197; https://doi.org/10.3390/ijms26031197 - 30 Jan 2025
Viewed by 309
Abstract
In order to detect and respond to invading pathogens, mammals have evolved a battery of pattern recognition receptors. Among these, RIG-I-like receptors (RLR) are cytosolic RNA sensors that play an essential role in the innate immune response against RNA viruses, including coronaviruses. In [...] Read more.
In order to detect and respond to invading pathogens, mammals have evolved a battery of pattern recognition receptors. Among these, RIG-I-like receptors (RLR) are cytosolic RNA sensors that play an essential role in the innate immune response against RNA viruses, including coronaviruses. In return, coronaviruses have acquired diverse strategies to impair RLR-mediated immune responses to enable productive infection. Viral innate immune evasion mechanisms have been well studied for highly pathogenic human coronaviruses (HCoVs), and often, these activities are thought to be linked to the severe symptoms these viruses can cause. Whether other coronaviruses, including human common cold coronaviruses, display similar activities has remained understudied. Here, we present evidence that the main protease (Mpro) of common cold HCoV-229E acts as an interferon (IFN) and NF-κB antagonist by disrupting RLR-mediated antiviral signalling. Furthermore, we show that HCoV-229E, HCoV-OC43 and MERS-CoV Mpros are able to directly cleave NEMO. We also show that HCoV-229E Mpro induces the cleavage and/or degradation of multiple other RLR pathway components, including MDA5, TBK1 and IKKε. Finally, we show that HCoV-229E infection leads to a delayed innate immune response that is accompanied by a decrease in NEMO protein levels. Our results suggest that NEMO degradation during HCoV-229E infection could be mediated, in part, by cellular degradation pathways, in addition to viral Mpro-mediated cleavage. Altogether, our research unveils innate immune evasion activities of the Mpros of low-pathogenic coronaviruses, which, despite their low pathogenicity, appear to share functionalities previously described for highly pathogenic HCoVs. Full article
(This article belongs to the Special Issue Viral Infections and Host Immune Responses)
Show Figures

Figure 1

Figure 1
<p><b>HCoV-229E Mpro expression blocks IRF3 nuclear translocation and IFN-β and NF-κB induction.</b> (<b>A</b>) HEK293T cells seeded in 24-well clusters containing coverslips were mock-transfected or transfected with a plasmid coding for RIG-I(2CARD) (100 ng/well) in combination with an empty vector (EV), a plasmid encoding V5-tagged HCoV-229E Mpro WT or Mpro C144A. At 16 hpt, cells were fixed and stained with antibodies against IRF3 (green) and V5 (red), and nuclei were counterstained with Hoechst dye (blue). Dashed empty arrows indicate cells expressing Mpro and cytosolic IRF3, and solid white arrows indicate cells expressing Mpro and nuclear IRF3. (<b>B</b>) HEK293T cells in 24-well plates were co-transfected with a combination of plasmids encoding the IFN-β-luciferase reporter (50 ng/well), renilla luciferase as control for transfection efficiency (5 ng/well), RIG-I(2CARD) (25 ng/well), and HCoV-229E Mpro WT (5, 25 or 50 ng/well) or Mpro C144A (50 ng/well). At 16 hpt, luciferase activity was measured. (<b>C</b>) HEK293T cells were co-transfected with a combination of plasmids encoding the NF-κB-luciferase reporter (50 ng/well), renilla luciferase as control for transfection efficiency (5 ng/well), RIG-I(2CARD) (25 ng/well), and HCoV-229E Mpro WT (5, 25 or 50 ng/well) or Mpro C144A (50 ng/well). At 16 hpt, luciferase activity was measured. Data are shown as triplicates obtained in one of three independent experiments that yielded similar results. Differences between groups were assessed by one-way ANOVA. Mean ± SEM is shown. ns, not significant; ***, <span class="html-italic">p</span> &lt; 0.0001; ****, <span class="html-italic">p</span> &lt; 0.00001.</p>
Full article ">Figure 2
<p><b>The Mpros of HCoV-229E, HCoV-OC43 and MERS-CoV cleave NEMO by means of their catalytic activity.</b> (<b>A</b>) HEK293T cells were co-transfected with a plasmid coding for human Myc-NEMO-HA (2 µg/well) and increasing amounts of a V5-tagged HCoV-229E WT Mpro mammalian expression plasmid (0.25, 0.5, 1 and 2 µg/well) or catalytic mutant Mpro (2 µg/well). (<b>B</b>) Myc-NEMO-HA (2 µg/well) and V5-tagged HCoV-229E WT Mpro (2 µg/well) mammalian expression plasmids were co-transfected in HEK293T cells, and cells were treated with increasing concentrations of Mpro inhibitor GC376. (<b>C</b>) HEK293T cells were left untreated or were treated with 25 µM zVAD for 2 h prior to transfection and were then co-transfected with the Myc-NEMO-HA plasmid (2 µg/well) and the V5-HCoV-229E Mpro plasmid. (<b>D</b>) HEK293T cells were transfected with a combination of plasmids coding for double-tagged NEMO and WT Mpro of either HCoV-OC43, MERS-CoV or SARS-CoV-2 (the latter as a positive control). (<b>A</b>–<b>D</b>) At 24 h post-transfection (hpt), cells were lysed, and protein lysates were immunoblotted for Myc, HA, V5 and β-actin. Immunoblots are representative of at least three independent experiments. FL, full-length; WT, wild type; C&gt;A, catalytic mutant Mpro; solid black arrows indicate cleaved fragments identified by immunoblotting; black asterisks indicate unspecific bands; red asterisks indicate less abundant cleaved fragments.</p>
Full article ">Figure 3
<p><b>Identification of HCoV-229E Mpro cleavage sites within NEMO.</b> (<b>A</b>) Previously identified cleavage sites within NEMO for PEDV [<a href="#B23-ijms-26-01197" class="html-bibr">23</a>], PDCoV [<a href="#B24-ijms-26-01197" class="html-bibr">24</a>], FIPV [<a href="#B25-ijms-26-01197" class="html-bibr">25</a>], SARS-CoV and SARS-CoV-2 [<a href="#B26-ijms-26-01197" class="html-bibr">26</a>,<a href="#B27-ijms-26-01197" class="html-bibr">27</a>]. (<b>B</b>) Schematic representation of NEMO protein in which the glutamine residues mutated into an alanine are indicated by black stars. From N-terminal to C-terminal: αH1, Helical domain 1; CC1, Coiled coil 1; αH2, Helical domain 2; CC2, Coiled coil 2; NUB, NEMO ubiquitin binding; LZ, Leucin zipper; ZF, Zinc finger. (<b>C</b>) HEK293T cells were transfected with plasmids coding for double-tagged NEMO single mutants or with a combination of a plasmids for the expression of the different NEMO mutants and V5-HCoV-229E Mpro. (<b>D</b>) Plasmids coding for NEMO double mutant (x2A, Q205A + Q231A), triple mutant (x3A, Q83A + Q205A + Q231A) or quintuple mutant (x5A, Q83A + Q205A + Q231A + Q304A + Q313A) were transfected alone or in combination with V5-HCoV-229E Mpro. (<b>E</b>) The indicated NEMO mutants were co-expressed in HEK293T cells either alone or together with the V5-tagged Mpros of HCoV-OC43 or MERS-CoV. (<b>C</b>–<b>E</b>) At 24 hpt, cells were lysed, and protein lysates were immunoblotted for Myc, HA, V5 and β-actin. Immunoblots are representative of at least three independent experiments. FL, full-length; WT, wild type; C&gt;A, catalytic mutant Mpro; solid black arrows indicate cleaved fragments identified by immunoblotting; black asterisks indicate unspecific bands; red asterisks indicate less abundant cleaved fragments.</p>
Full article ">Figure 4
<p><b>HCoV-229E Mpro-uncleavable NEMO barely recovers NF-κB induction in the presence of Mpro.</b> (<b>A</b>) Plasmids coding for NF-κB-luciferase reporter (50 ng/well), renilla luciferase as a control for transfection efficiency (5 ng/well), FLAG-NEMO K277A (400 ng/well), and HCoV-229E Mpro WT (50 ng/well) or Mpro C144A (50 ng/well) were co-transfected in HEK293T cells in 24-well plates. At 16 hpt, cells were lysed, and luciferase activity was measured. (<b>B</b>) HEK293T cells were co-transfected with mammalian expression vectors for FLAG-NEMO K277A and HCoV-229E Mpro WT or C144A. At 24 hpt, cells were lysed, and protein lysates were immunoblotted for FLAG, V5 and β-actin. (<b>C</b>) HEK293T cells were co-transfected with plasmids coding for NF-κB-luciferase reporter (50 ng/well), renilla luciferase as a control for transfection efficiency (5 ng/well), FLAG-NEMO K277A or FLAG-K277A x5A (400 ng/well), and HCoV-229E Mpro WT (50 ng/well) or Mpro C144A (50 ng/well). At 16 hpt, cells were lysed, and luciferase activity was measured. Dual-luciferase reporter assays were performed as triplicates and repeated at least three times. Differences between groups were assessed by one-way ANOVA (panel (<b>A</b>)) or two-way ANOVA (panel (<b>C</b>)). Immunoblots are representative of two independent experiments. Mean ± SEM is shown. FL, full-length; WT, wild type; C&gt;A, catalytic mutant Mpro; solid black arrows indicate cleaved fragments identified by immunoblotting; EV, empty vector; *, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.0001; ****, <span class="html-italic">p</span> &lt; 0.00001.</p>
Full article ">Figure 5
<p><b>The Mpro of HCoV-229E disrupts the RLR signalling pathway at multiple levels.</b> (<b>A</b>) HEK293T cells were co-transfected with plasmids encoding the IFN-β-luciferase reporter (50 ng/well), renilla luciferase as a control for transfection efficiency (5 ng/well), and HCoV-229E Mpro WT (50 ng/well) or Mpro C144A (50 ng/well), together with a plasmid expressing MDA5 (25 ng/well), MAVS (25 ng/well), TBK1 (50 ng/well), IKKε (50 ng/well) or IRF3(5D) (50 ng/well). At 16 hpt, cells were lysed, and luciferase activity was measured. (<b>B</b>) HEK293T cells were co-transfected with a combination of plasmids encoding the NF-κB-luciferase reporter (50 ng/well); renilla luciferase as a control for transfection efficiency (25 ng/well); HCoV-229E Mpro WT (50 ng/well) or Mpro C144A (50 ng/well); and MDA5 (25 ng/well), MAVS (25 ng/well), TBK1 (50 ng/well) or IKKε (50 ng/well). At 16 hpt, cells were lysed, and luciferase activity was measured. (<b>C</b>) Plasmids coding for FLAG-tagged RLR ligands RIG-I (2 μg/well), MDA5 (2 μg/well), MAVS (2 μg/well), TBK1 (2 μg/well), IKKε (2 μg/well) or GFP-IRF3 (2 μg/well) were co-transfected with plasmids coding for HCoV-229E Mpro WT or C144A. At 24 hpt, cells were lysed, and protein lysates were immunoblotted for FLAG or GFP, V5 and β-actin. Data are shown as triplicates obtained in one of three independent experiments that yielded similar results. Differences between groups were assessed by one-way ANOVA. Mean ± SEM is shown. Immunoblots are representative of two independent experiments. FL, full-length; EV, empty vector; WT, wild type; C&gt;A, catalytic mutant Mpro; red asterisks indicate cleavage products; ns, not significant; *, <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; 0.00001.</p>
Full article ">Figure 6
<p><b>The impairment of IFN-β and NF-κB induction by HCoV-229E Mpro is not attributed to Mpro-mediated cytotoxicity.</b> (<b>A</b>,<b>B</b>) HEK293T cells in 24-well plates were transfected with a plasmid encoding eGFP (negative control, 500 ng/well), increasing amounts (5, 25, 50, 100 and 500 ng/well) of a plasmid coding for HCoV-229E Mpro WT or a plasmid encoding HCoV-229E Mpro C144A (500 ng/well) or treated with ethanol (EtOH) as a positive control for cell death. At 16 hpt, cell viability (<b>A</b>) and/or lytic cell death (<b>B</b>) were assessed by MTS assay and LDH release assay, respectively. (<b>C</b>) HEK293T cells were transfected with a plasmid coding for V5-tagged HCoV-229E Mpro WT (50 ng/well) or a plasmid coding for Mpro C144A (50 ng/well). At 16 hpt, cells were fixed and stained with an antibody against V5. (<b>D</b>) HEK293T cells were transfected with a plasmid encoding eGFP (negative control, 500 ng/well), increasing amounts (5, 25, 50, 100 and 500 ng/well) of a plasmid coding for HCoV-229E Mpro WT or a plasmid encoding HCoV-229E Mpro C144A (500 ng/well). At 40 hpt, lytic cell death was assessed. (<b>E</b>) HEK293T cells treated with DMSO or fmk-zVAD (25 µM) were co-transfected with a combination of plasmids encoding the IFN-β-luciferase reporter (50 ng/well), renilla luciferase as control for transfection efficiency (5 ng/well), RIG-I(2CARD) (25 ng/well), and HCoV-229E Mpro WT (50 ng/well) or Mpro C144A (50 ng/well). At 16 hpt, luciferase activity was measured. (<b>F</b>) HEK293T cells treated with DMSO or fmk-zVAD (25 µM) were co-transfected with a combination of plasmids encoding the NF-κB-luciferase reporter (50 ng/well), renilla luciferase (5 ng/well), NEMO K277A (400 ng/well), and HCoV-229E Mpro WT (50 ng/well) or Mpro C144A (50 ng/well). At 16 hpt, luciferase activity was measured. Experiments were performed once (<b>A</b>,<b>E</b>,<b>F</b>) or twice (<b>B</b>,<b>C</b>) as triplicates. Differences between groups were assessed by one-way ANOVA (panels (<b>A</b>,<b>B</b>,<b>D</b>)) or two-way ANOVA (panels (<b>E</b>,<b>F</b>)). Mean ± SEM is shown. LDH, lactate dehydrogenase; EV, empty vector; WT, wild type; C&gt;A, catalytic mutant Mpro; ns, not significant; *, <span class="html-italic">p</span> &lt; 0.01; ****, <span class="html-italic">p</span> &lt; 0.00001.</p>
Full article ">Figure 7
<p><b>HCoV-229E infection leads to delayed innate immune responses and to decreased NEMO protein levels.</b> (<b>A</b>) Huh7 cells were infected with HCoV-229E at an MOI of 5, supernatants were harvested at the indicated time points and HCoV-229E infectious virus particles were quantified by plaque assay. (<b>B</b>) Huh7 cells were infected with HCoV-229E at an MOI of 5; then, 8, 24 and 48 hpi, cell lysates were harvested, and IFN-β, IFN-λ and IL-6 mRNA levels were determined by RT-qPCR. (<b>C</b>) Huh7 cells were infected with HCoV-229E at an MOI of 5, and protein lysates were collected at the indicated time points and analysed by immunoblotting for NEMO, HCoV-229E nucleocapsid (N) protein and β-actin. (<b>D</b>) Huh7 cells were infected with HCoV-229E at an MOI of 5; then, 8, 24 and 48 hpi, cell lysates were harvested, and NEMO mRNA levels were determined by RT-qPCR. (<b>E</b>) BEAS-2B cells were infected with HCoV-229E at an MOI of 5, and at 24 hpi, protein lysates were collected and analysed by immunoblotting for NEMO, HCoV-229E nucleocapsid protein and β-actin. (<b>F</b>) Huh7 cells were mock-infected or infected with HCoV-229E at an MOI of 5 and treated with vehicle control (DMSO), zVAD (25 µM) or bafilomycin A1 (100 nM) for 24 h or with MG132 (20 µM) for 12 h (from t = 12–24 hpi). At 24 hpi, protein lysates were collected and analysed by immunoblotting for NEMO, HCoV-229E nucleocapsid protein and β-actin. Experiments were performed as triplicates and repeated at least twice. Differences between groups were assessed by one-way ANOVA. Immunoblots are representative of two (<b>F</b>) or more (<b>C</b>,<b>E</b>) independent experiments. ns, not significant; ***, <span class="html-italic">p</span> &lt; 0.0001; ****, <span class="html-italic">p</span> &lt; 0.00001. Mean ± SEM is shown. Amounts of NEMO are normalized by β-actin and expressed relative to the corresponding controls and are indicated below Western blot panels.</p>
Full article ">
15 pages, 6493 KiB  
Article
Glutathione Depletion Exacerbates Hepatic Mycobacterium tuberculosis Infection
by Kayvan Sasaninia, Aishvaryaa Shree Mohan, Ali Badaoui, Ira Glassman, Sonyeol Yoon, Arshavir Karapetyan, Afsal Kolloli, Ranjeet Kumar, Santhamani Ramasamy, Selvakumar Subbian and Vishwanath Venketaraman
Biology 2025, 14(2), 131; https://doi.org/10.3390/biology14020131 - 27 Jan 2025
Viewed by 731
Abstract
Extrapulmonary tuberculosis (EPTB) accounts for approximately 17% of all Mycobacterium tuberculosis (M.tb) infections globally. Immunocompromised individuals, such as those with HIV infection or type 2 diabetes mellitus (T2DM), are at an increased risk for EPTB. Previous studies have demonstrated that patients [...] Read more.
Extrapulmonary tuberculosis (EPTB) accounts for approximately 17% of all Mycobacterium tuberculosis (M.tb) infections globally. Immunocompromised individuals, such as those with HIV infection or type 2 diabetes mellitus (T2DM), are at an increased risk for EPTB. Previous studies have demonstrated that patients with HIV and T2DM exhibit diminished synthesis of glutathione (GSH) synthesizing enzymes. In a murine model, we showed that the diethyl maleate (DEM)-induced depletion of GSH in the lungs led to increased M.tb burden and an impaired pulmonary granulomatous response to M.tb infection. However, the effects of GSH depletion during active EPTB in the liver and spleen have yet to be elucidated. In this study, we evaluated hepatic GSH and malondialdehyde (MDA) levels, as well as cytokine profiles, in untreated and DEM-treated M.tb-infected wild-type (WT) C57BL/6 mice. Additionally, we assessed hepatic and splenic M.tb burdens and tissue pathologies. DEM treatment resulted in a significant decrease in the levels of the reduced form of GSH and an increase in MDA, oxidized GSH, and interleukin (IL)-6 levels. Furthermore, DEM-induced GSH decrease was associated with decreased production of IL-12 and IL-17 and elevated production of interferon-gamma (IFN-γ), tumor necrosis factor (TNF)-α, and transforming growth factor (TGF)-β. A significant increase in M.tb growth was detected in the liver and spleen in DEM-treated M.tb-infected mice. Large, disorganized lymphocyte infiltrates were detected in the hepatic tissues of DEM-treated mice. Overall, GSH diminishment impaired the granulomatous response to M.tb in the liver and exacerbated M.tb growth in both the liver and spleen. These findings provide critical insights into the immunomodulatory role of GSH in TB pathogenesis and suggest potential therapeutic avenues for the treatment of extrapulmonary M.tb infections. Full article
Show Figures

Figure 1

Figure 1
<p>Levels of a reduced form (rGSH) and oxidized form (GSSG) of glutathione in the liver of untreated (n = 3) and DEM-treated (n = 3) <span class="html-italic">M.tb</span>-infected C57BL/6 mice. (<b>A</b>) rGSH levels 2, 4, and 8 weeks post-<span class="html-italic">M.tb</span> infection; (<b>B</b>) GSSG levels 2, 4, and 8 weeks post-<span class="html-italic">M.tb</span> infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Dots indicate replicates of each mouse. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using the Kruskal–Wallis test. The placement of an asterisk (*) indicates statistical significance between compared groups. A single asterisk (*) denotes a <span class="html-italic">p</span>-value &lt; 0.05. Double asterisks (**) imply a <span class="html-italic">p</span>-value below 0.005. Triple asterisks (***) indicate a <span class="html-italic">p</span>-value &lt; 0.0005. ns indicates no significance.</p>
Full article ">Figure 2
<p>Malondialdehyde (MDA) levels in the liver of untreated and DEM-treated C57BL/6 mice; (<b>A</b>) 2 weeks post-<span class="html-italic">M.tb</span> infection, (<b>B</b>) 4 weeks post-<span class="html-italic">M.tb</span> infection, and (<b>C</b>) 8 weeks post-<span class="html-italic">M.tb</span> infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Dots indicate replicates of each mouse. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using an unpaired <span class="html-italic">t</span>-test with Welch’s correction. The placement of an asterisk (*) indicates statistical significance between compared groups with <span class="html-italic">p</span>-value &lt; 0.05. ns indicate no significance.</p>
Full article ">Figure 3
<p>Levels of IL-6 in the liver of untreated and DEM-treated C57BL/6 mice; (<b>A</b>) 2 weeks post-<span class="html-italic">M.tb</span> infection, (<b>B</b>) 4 weeks post-<span class="html-italic">M.tb</span> infection, and (<b>C</b>) 8 weeks post-<span class="html-italic">M.tb</span> infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Dots indicate replicates of each mouse. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using an unpaired <span class="html-italic">t</span>-test with Welch’s correction. The placement of an asterisk (*) indicates statistical significance between compared groups. A single asterisk (*) denotes a <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 4
<p>Levels of Th1 cytokines in the liver of untreated and DEM-treated <span class="html-italic">M.tb</span>-infected C57BL/6 mice. (<b>A</b>–<b>C</b>) IL-12 levels 2, 4, and 8 weeks post-<span class="html-italic">M.tb</span> infection; (<b>D</b>–<b>F</b>) IL-2 levels 2, 4, and 8 weeks post-<span class="html-italic">M.tb</span> infection; (<b>G</b>–<b>I</b>) TNF-α levels 2, 4, and 8 weeks post-M.tb infection; (<b>J</b>–<b>L</b>) IFN-γ levels 2, 4, and 8 weeks post-<span class="html-italic">M.tb</span> infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Dots represent replicates of each mouse. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using an unpaired <span class="html-italic">t</span>-test with Welch’s correction. The placement of an asterisk (*) indicates statistical significance between compared groups. A single asterisk (*) denotes a <span class="html-italic">p</span>-value &lt; 0.05. Double asterisk (**) indicates a <span class="html-italic">p</span>-value &lt; 0.005. ns indicates statistical non-significance.</p>
Full article ">Figure 5
<p>Levels of IL-17 in the liver of untreated and DEM-treated <span class="html-italic">M.tb</span>-infected C57BL/6 mice; (<b>A</b>) 2 weeks post-infection, (<b>B</b>) 4 weeks post-infection, and (<b>C</b>) 8 weeks post-infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Dots represent replicates of each mouse. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using an unpaired <span class="html-italic">t</span>-test with Welch’s correction. The placement of an asterisk (*) indicates statistical significance between compared groups. ns denotes non-significance. Double asterisks (**) denote a <span class="html-italic">p</span>-value &lt; 0.005. Quadruple asterisks (****) denote a <span class="html-italic">p</span>-value &lt; 0.0001.</p>
Full article ">Figure 6
<p>Levels of TGF-β in the liver of untreated and DEM-treated <span class="html-italic">M.tb</span>-infected C57Bl/6 mice; (<b>A</b>) 2 weeks post-infection, (<b>B</b>) 4 weeks post-infection, and (<b>C</b>) 8 weeks post-infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using an unpaired <span class="html-italic">t</span>-test with Welch’s correction. The placement of an asterisk (*) indicates statistical significance between compared groups. A single asterisk (*) denotes a <span class="html-italic">p</span>-value &lt; 0.05. Double asterisks (**) denote a <span class="html-italic">p</span>-value &lt; 0.005.</p>
Full article ">Figure 7
<p>Survival of <span class="html-italic">M.tb</span> in the liver and spleen of untreated and DEM-treated <span class="html-italic">M.tb</span>-infected C57BL/6 mice. (<b>A</b>) CFU/mL <span class="html-italic">M.tb</span> in the liver of <span class="html-italic">M.tb</span>-infected mice 8 weeks post-infection. (<b>B</b>) CFU/mL <span class="html-italic">M.tb</span> in the spleen of <span class="html-italic">M.tb</span>-infected mice 4 weeks post-infection. (<b>C</b>) CFU/mL <span class="html-italic">M.tb</span> in the liver of <span class="html-italic">M.tb</span>-infected mice 8 weeks post-infection. The sample size (n) includes three female mice (n = 3) each in the untreated and DEM-treated groups. Dots represent replicates of each mouse. Comparisons between untreated and DEM-treated groups were analyzed via GraphPad prism software using an unpaired <span class="html-italic">t</span>-test with Welch’s correction. The placement of an asterisk (*) indicates statistical significance between compared groups. A single asterisk (*) denotes a <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 8
<p>Hematoxylin and eosin (H&amp;E) staining of the liver section of untreated and DEM-treated C57BL/6 mice 4 weeks post-<span class="html-italic">M.tb</span> infection. Thick arrows indicate immune cell infiltration into a granulomatous lesion; thin arrows indicate binuclear hepatocytes. Tissue sections were imaged at 100× and 400× the original magnification. (<b>A</b>) Hepatic tissue of untreated <span class="html-italic">M.tb</span> infected mice imaged at 100×. (<b>B</b>) Hepatic tissue of untreated <span class="html-italic">M.tb</span> infected mice imaged at 400×. (<b>C</b>). Hepatic tissue of DEM treated <span class="html-italic">M.tb</span> infected mice imaged at 100×. (<b>D</b>) Hepatic tissue of DEM treated <span class="html-italic">M.tb</span>-infected mice imaged at 400×. The scale bar for 100× panels refers to 500 µm, and for 400× panels, it refers to 200 µm.</p>
Full article ">Figure 9
<p>Hematoxylin and eosin (H&amp;E) staining of the liver section of untreated and DEM-treated C57BL/6 mice 8 weeks post-<span class="html-italic">M.tb</span> infection. Thick arrows indicate immune cell infiltration into a granulomatous lesion; thin arrows indicate binuclear hepatocytes. Tissue sections were imaged at 100× and 400× the original magnification. (<b>A</b>) Hepatic tissue of untreated <span class="html-italic">M.tb</span> infected mice imaged at 100×. (<b>B</b>) Hepatic tissue of untreated <span class="html-italic">M.tb</span> -infected mice imaged at 400×. (<b>C</b>). Hepatic tissue of DEM treated M.tb infected mice imaged at 100×. (<b>D</b>) Hepatic tissue of DEM treated <span class="html-italic">M.tb</span>-infected mice imaged at 400×. The scale bar for 100× panels refers to 500 µm, and for 400× panels, it refers to 200 µm.</p>
Full article ">Figure 10
<p>Overview of study findings: DEM-induced GSH depletion leads to an increase in MDA, dysregulated cytokine response, increased <span class="html-italic">M.tb</span> burden, and impaired granulomatous control of infection in the liver of WT C57BL/6 mice.</p>
Full article ">
21 pages, 8609 KiB  
Article
Signature Construction Associated with Tumor-Infiltrating Macrophages Identifies IRF8 as a Novel Biomarker for Immunotherapy in Advanced Gastric Cancer
by Wanqian Liao, Yu Wang, Rui Wang, Bibo Fu, Xiangfu Chen, Ying Ouyang, Bing Bai, Ying Jin, Yunxin Lu, Furong Liu, Yang Zhang, Dongni Shi and Dongsheng Zhang
Int. J. Mol. Sci. 2025, 26(3), 1089; https://doi.org/10.3390/ijms26031089 - 27 Jan 2025
Viewed by 461
Abstract
Advanced gastric cancer (AGC) is characterized by poor prognosis and limited responsiveness to immunotherapy. Tumor-associated macrophages (TAMs) play a pivotal role in cancer progression and therapeutic outcomes. In this study, we developed a novel gene signature associated with M1-like TAMs using data from [...] Read more.
Advanced gastric cancer (AGC) is characterized by poor prognosis and limited responsiveness to immunotherapy. Tumor-associated macrophages (TAMs) play a pivotal role in cancer progression and therapeutic outcomes. In this study, we developed a novel gene signature associated with M1-like TAMs using data from the Gene Expression Omnibus (GEO) and The Cancer Genome Atlas (TCGA) to predict prognosis and immunotherapy response. This gene signature was determined as an independent prognostic indicator for AGC, with high-risk patients exhibiting an immunosuppressive tumor immune microenvironment (TIME) and poorer survival outcomes. Furthermore, Interferon regulatory factor 8 (IRF8) was identified as a key gene and validated through in vitro and in vivo experiments. IRF8 overexpression reshaped the suppressive TIME, leading to an increased presence of M1-like TAMs, IFN-γ+ CD8+ T cells, and Granzyme B+ CD8+ T cells. Notably, the combination of IRF8 overexpression and anti-PD-1 therapy significantly inhibited tumor growth in syngeneic mouse models. AGC patients with elevated IRF8 expression were found to be more responsive to anti-PD-1 treatment. These findings highlight potential biomarkers for prognostic evaluation and immunotherapy in AGC, offering insights that could guide personalized treatment strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Identification of M1-like TAM-related clusters. (<b>A</b>) Survival analysis of the patients with high and low infiltration of M1-like TAMs grouped by the median CIBERSORT-based M1 macrophage scores. (<b>B</b>) WGCNA identified an M1-like TAM-related module. (<b>C</b>) The heatmap of the consensus matrix showing the optimal values for clusters was K = 3. (<b>D</b>) The KM curve revealed the difference in OS of the three clusters. (<b>E</b>) Differential gene GO and KEGG enrichment analysis. GO, Gene Ontology; KEGG, Kyoto Encyclopedia of Genes and Genomes; KM, Kaplan–Meier; TAMs, tumor-associated macrophages; WGCNA, weighted correlation network analysis.</p>
Full article ">Figure 2
<p>Construction of risk model in GEO training cohort and validation in TCGA-STAD cohort. (<b>A</b>) Distribution of risk scores, survival status, and gene expression heatmap of the 11 genes for the GEO training cohort. (<b>B</b>) KM plot of OS in high- and low-risk groups for GEO training cohort. (<b>C</b>) The boxplot presents the differences in risk scores among the three clusters. (<b>D</b>,<b>E</b>) Distribution of risk scores, survival status, gene expression heatmap, and Kaplan–Meier curve of OS in high- and low-risk groups for TCGA-STAD validation cohort. GEO, Gene Expression Omnibus; KM, Kaplan–Meier; OS, overall survival; TCGA-STAD, The Cancer Genome Atlas-stomach adenocarcinoma.</p>
Full article ">Figure 3
<p>Different immune landscapes between high- and low-risk groups for GEO training cohort. (<b>A</b>) Comparison of immune cell infiltration between high- and low-risk groups using CIBERSORT analysis. (<b>B</b>) Different expression of immune-related genes between high- and low-risk groups. <span class="html-italic">p</span> values are marked as: ns, not 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. GEO, Gene Expression Omnibus.</p>
Full article ">Figure 4
<p>Mutational landscape analysis of risk model in TCGA-STAD cohort. (<b>A</b>) Comparison of TMB between high-risk and low-risk groups. (<b>B</b>) Pearson correlation analysis between the risk score and TMB. The red shaded area represents the 95% confidence interval around the regression line. (<b>C</b>,<b>D</b>) Waterfall plots presenting the top 20 genes with the highest mutation frequency in high-risk (<b>C</b>) and low-risk (<b>D</b>) groups. Each column represents an individual patient. The bar chart at the top displays the TMB for each patient, while the numbers on the right indicate the mutation frequency for each gene. The stacked bar chart on the right shows the distribution of mutation types for each gene, including missense mutations, nonsense mutations, frame-shift insertions/deletions, and others. TCGA-STAD, The Cancer Genome Atlas-stomach adenocarcinoma; TMB, tumor mutation burden.</p>
Full article ">Figure 5
<p>High-risk scores predict poor immunotherapy response based on TIDE analysis. (<b>A</b>,<b>D</b>) Boxplots comparing TIDE scores between high-risk and low-risk groups in the GEO cohort (<b>A</b>) and TCGA-STAD cohort (<b>D</b>) using the Wilcoxon test. (<b>B</b>,<b>E</b>) Waterfall plots showing the distribution of immunotherapy responders (red) and non-responders (blue) based on TIDE scores in the GEO cohort (<b>B</b>) and TCGA-STAD cohort (<b>E</b>). (<b>C</b>,<b>F</b>) Stacked bar charts illustrating the proportions of responders and non-responders in high-risk and low-risk groups in the GEO cohort (<b>C</b>) and TCGA-STAD cohort (<b>F</b>), analyzed by the Chi-square test. GEO, Gene Expression Omnibus; TCGA-STAD, The Cancer Genome Atlas-stomach adenocarcinoma; TIDE, Tumor Immune Dysfunction and Exclusion.</p>
Full article ">Figure 6
<p>Identifying <span class="html-italic">IRF8</span> as a hub gene and its impact on AGC prognosis. (<b>A</b>) Venn plot showing the intersection of the hub gene of M1-TAM-related WGCNA and LASSO module. (<b>B</b>) KM survival analysis for high and low <span class="html-italic">IRF8</span> expression in the GEO cohort. (<b>C</b>) The difference between high and low <span class="html-italic">IRF8</span> expression from the GEO cohort through the Wilcox test. (<b>D</b>,<b>E</b>) Comparison of immune cell infiltration and immune-related genes between high and low <span class="html-italic">IRF8</span> expression. (<b>F</b>) GSVA analysis revealed the top 15 significantly enriched KEGG pathways for high and low <span class="html-italic">IRF8</span> expression. <span class="html-italic">p</span> values are marked as ns, not 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. AGC, advanced gastric cancer; GEO, Gene Expression Omnibus; GSVA, Gene set variation analysis; <span class="html-italic">IRF8</span>, Interferon regulatory factor 8; KEGG, Kyoto Encyclopedia of Genes and Genomes; KM, Kaplan–Meier; LASSO, Least Absolute Shrinkage and Selection Operator; TAMs, tumor-associated macrophages; WGCNA, weighted correlation network analysis.</p>
Full article ">Figure 7
<p><span class="html-italic">IRF8</span> overexpression inhibits tumor progression in different GC models. (<b>A</b>) Schematic of the GC subcutaneous tumor model. C57BL/6 mice were subcutaneously injected with MC38 cells (5 × 10<sup>6</sup> cells per mouse) and then treated with or without doxycycline. (<b>B</b>,<b>C</b>) Tumor, final tumor volume, and tumor weight in C57BL/6 mice bearing <span class="html-italic">IRF8</span> overexpression or vector MC38 cells. (<b>D</b>) Schematic of the GC orthotopic tumor model. MC38 cells were injected into the stomach wall (10<sup>6</sup> cells per mouse) and then treated with or without doxycycline. At the experimental endpoint, mice were sacrificed to harvest tumors and spleens for flow cytometric analysis. (<b>E</b>,<b>F</b>) Representative images of abdominal metastatic tumor nodules and orthotopic tumors of orthotopic GC models in each group. (<b>G</b>) Orthotopic tumor weight and the number of GC models. (<b>H</b>) Tumors were imaged every 10 days using bioluminescence imaging to determine tumor growth and invasion. Representative images are shown on the right. (<b>I</b>) Representative images of IHC staining of <span class="html-italic">IRF8</span> in orthotopic tumors. Data are presented as mean ± s.d. Statistical analysis was performed using two-sided Student’s <span class="html-italic">t</span>-test for comparisons in (<b>C</b>,<b>G</b>) and two-way ANOVA in (<b>H</b>). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; n = 5 per group. DOX, doxycycline; GC, gastric cancer; IHC, Immunohistochemistry; <span class="html-italic">IRF8</span>, Interferon regulatory factor 8; OE, overexpression.</p>
Full article ">Figure 8
<p><span class="html-italic">IRF8</span> overexpression reshapes the immune-suppressive TIME in vivo and in vitro. Percentages of (<b>A</b>) CD3<sup>+</sup>CD4<sup>+</sup> T cells and CD3<sup>+</sup>CD8<sup>+</sup> T cells, (<b>B</b>,<b>C</b>) Granzyme B<sup>+</sup>/CD8<sup>+</sup> T cells, and (<b>D</b>,<b>E</b>) CD86<sup>+</sup>/F4/80<sup>+</sup> cells and CD206<sup>+</sup>/F4/80<sup>+</sup> cells in tumors and spleens of the orthotopic GC models were measured by flow cytometry at the experimental endpoint (n = 5 per group). (<b>F</b>) In vitro, after being treated with the CM of <span class="html-italic">IRF8</span>-overexpressed or empty vector MKN45, the polarization of THP-1 was detected through flow cytometry analysis. (<b>G</b>) For IFN-γ and Granzyme B, α-CD3/CD28-pretreated CD3<sup>+</sup>T cells were co-cultured with TAMs induced by CM from <span class="html-italic">IRF8</span>-overexpressed or control samples in vitro. Cells were collected to perform flow cytometry analysis (n = 3 per group). All data are presented as mean ± s.d. Two-sided Student’s <span class="html-italic">t</span>-test was used for statistical analysis. ns, not 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. GC, gastric cancer; <span class="html-italic">IRF8</span>, Interferon regulatory factor 8; OE, overexpression; TIME, tumor immune microenvironment.</p>
Full article ">Figure 9
<p><span class="html-italic">IRF8</span> overexpression enhances the efficacy of anti-PD1 therapy. (<b>A</b>) Schematic of GC subcutaneous tumor model establishment and treatment diagram. C57BL/6 mice were subcutaneously injected with MC38 cells (5 × 10<sup>6</sup> cells per mouse) and then treated with or without doxycycline. Anti-PD1 was delivered intraperitoneally (i.p.) (5 mg/kg) on day 7 after tumor inoculation. (<b>B</b>,<b>C</b>) Tumor, final tumor volume, and tumor weight in C57BL/6 mice were measured at the experimental endpoint. (<b>D</b>) Schematic of GC subcutaneous tumor model establishment and treatment diagram. A total of 615 mice were subcutaneously injected with MFC cells (5 × 10<sup>6</sup> cells per mouse) and then treated with or without doxycycline. Anti-PD1 was delivered intraperitoneally (i.p.) (5 mg/kg) on day 7 after tumor inoculation. (<b>E</b>,<b>F</b>) Tumor, final tumor volume, and tumor weight in 615 mice were measured at the experimental endpoint. All data are presented as mean ± s.d. Two-sided Student’s <span class="html-italic">t</span>-test was used for statistical analysis. ns, not significant; * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001; n = 5 per group. (<b>G</b>,<b>H</b>) Boxplot and Waterfall plot showing the different <span class="html-italic">IRF8</span> expression between responders and non-responders from RPJEB25780. (<b>I</b>) Bar plot showing the proportion of responders and non-responders in distinct <span class="html-italic">IRF8</span> expression subgroups. DOX, doxycycline; GC, gastric cancer; i.p., intraperitoneally; <span class="html-italic">IRF8</span>, Interferon regulatory factor 8; OE, overexpression.</p>
Full article ">
Back to TopTop