[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

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 (2,499)

Search Parameters:
Keywords = immunomodulation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
33 pages, 10253 KiB  
Systematic Review
Schistosomiasis–Microbiota Interactions: A Systematic Review and Meta-Analysis
by Philip Afful, Godwin Kwami Abotsi, Czarina Owusua Adu-Gyamfi, George Benyem, Gnatoulma Katawa, Samuel Kyei, Kathrin Arndts, Manuel Ritter and Kwame Kumi Asare
Pathogens 2024, 13(10), 906; https://doi.org/10.3390/pathogens13100906 - 16 Oct 2024
Abstract
Introduction: Schistosomiasis, a tropical disease affecting humans and animals, affected 251.4 million people in 2021. Schistosoma mansoni, S. haematobium, S. intercalatum, and S. japonicum are primary human schistosomes, causing tissue damage, granulomas, ulceration, hemorrhage, and opportunistic pathogen entry. The gut [...] Read more.
Introduction: Schistosomiasis, a tropical disease affecting humans and animals, affected 251.4 million people in 2021. Schistosoma mansoni, S. haematobium, S. intercalatum, and S. japonicum are primary human schistosomes, causing tissue damage, granulomas, ulceration, hemorrhage, and opportunistic pathogen entry. The gut and urinary tract microbiota significantly impact a host’s susceptibility to schistosomiasis, disrupting microbial balance; however, this relationship is not well understood. This systematic review and meta-analysis explores the intricate relationship between schistosomiasis and the host’s microbiota, providing crucial insights into disease pathogenesis and management. Methods: This systematic review used PRISMA guidelines to identify peer-reviewed articles on schistosomiasis and its interactions with the host microbiome, using multiple databases and Google Scholar, providing a robust dataset for analysis. The study utilized Meta-Mar v3.5.1; descriptive tests, random-effects models, and subgroups were analyzed for the interaction between Schistosomiasis and the microbiome. Forest plots, Cochran’s Q test, and Higgins’ inconsistency statistic (I2) were used to assess heterogeneity. Results: The human Schistosoma species were observed to be associated with various bacterial species isolated from blood, stool, urine, sputum, skin, and vaginal or cervical samples. A meta-analysis of the interaction between schistosomiasis and the host microbiome, based on 31 studies, showed 29,784 observations and 5871 events. The pooled estimates indicated a significant association between schistosomiasis and changes in the microbiome of infected individuals. There was considerable heterogeneity with variance effect sizes (p < 0.0001). Subgroup analysis of Schistosoma species demonstrated that S. haematobium was the most significant contributor to the overall heterogeneity, accounting for 62.1% (p < 0.01). S. mansoni contributed 13.0% (p = 0.02), and the coinfection of S. haematobium and S. mansoni accounted for 16.8% of the heterogeneity (p < 0.01), contributing to the variability seen in the pooled analysis. Similarly, praziquantel treatment (RR = 1.68, 95% CI: 1.07–2.64) showed high heterogeneity (Chi2 = 71.42, df = 11, p < 0.01) and also indicated that Schistosoma infections in males (RR = 1.46, 95% CI: 0.00 to 551.30) and females (RR = 2.09, 95% CI: 0.24 to 18.31) have a higher risk of altering the host microbiome. Conclusions: Schistosomiasis significantly disrupts the host microbiota across various bodily sites, leading to increased susceptibility to different bacterial taxa such as E. coli, Klebsiella, Proteus, Pseudomonas, Salmonella, Staphylococcus, Streptococcus, and Mycobacterium species (M. tuberculosis and M. leprae). This disruption enables these bacteria to produce toxic metabolites, which in turn cause inflammation and facilitate the progression of disease. The impact of schistosomiasis on the vaginal microbiome underscores the necessity for gender-specific approaches to treatment and prevention. Effective management of female genital schistosomiasis (FGS) requires addressing both the parasitic infection and the resulting microbiome imbalances. Additionally, praziquantel-treated individuals have different microbiome compositions compared to individuals with no praziquantel treatment. This suggests that combining praziquantel treatment with probiotics could potentially decrease the disease severity caused by an altered microbiome. Full article
Show Figures

Figure 1

Figure 1
<p>PRISMA flow chart for search and selection of included studies.</p>
Full article ">Figure 2
<p>The geographical distribution of the included studies. The red indicates the various studies and the countries where they were conducted.</p>
Full article ">Figure 3
<p>Forest plot showing the risk ratio of schistosomiasis and bacterial infections from 31 studies.</p>
Full article ">Figure 4
<p>Forest plot showing the risk difference of schistosomiasis and bacterial infections from 31 studies.</p>
Full article ">Figure 5
<p>Forest plot showing the risk ratio of schistosomiasis and bacterial infection based on <span class="html-italic">Schistosoma</span> species.</p>
Full article ">Figure 6
<p>Forest plot showing the risk difference of schistosomiasis and bacterial infection based on <span class="html-italic">Schistosoma</span> species.</p>
Full article ">Figure 7
<p>Forest plot showing the risk ratio of schistosomiasis and bacterial infection based on praziquantel treatment.</p>
Full article ">Figure 8
<p>Forest plot showing the risk difference of schistosomiasis and bacterial infection based on praziquantel treatment.</p>
Full article ">Figure 9
<p>Forest plot showing the risk ratio of schistosomiasis and bacterial infection based on gender.</p>
Full article ">Figure 10
<p>Forest plot showing the risk difference of schistosomiasis and bacterial infection based on gender.</p>
Full article ">Figure 11
<p>Boxplot showing the effect size distribution of <span class="html-italic">Schistosoma</span> species among schistosomiasis and bacterial coinfection; (<b>a</b>) effect size calculation from risk ratio, (<b>b</b>) effect size calculation from risk difference.</p>
Full article ">Figure 12
<p>Funnel plot (Trim and Fill) showing asymmetrical distribution for schistosomiasis and bacterial infection from the 31 studies; (<b>a</b>) assess publication bias based on risk ratio analysis, (<b>b</b>) assess publication bias based on risk difference analysis.</p>
Full article ">Figure 12 Cont.
<p>Funnel plot (Trim and Fill) showing asymmetrical distribution for schistosomiasis and bacterial infection from the 31 studies; (<b>a</b>) assess publication bias based on risk ratio analysis, (<b>b</b>) assess publication bias based on risk difference analysis.</p>
Full article ">Figure 13
<p>Interplay of schistosome infection, microbiome, and immune system.</p>
Full article ">
18 pages, 3914 KiB  
Article
Overcoming Irinotecan Resistance by Targeting Its Downstream Signaling Pathways in Colon Cancer
by Shashank Saurav, Sourajeet Karfa, Trung Vu, Zhipeng Liu, Arunima Datta, Upender Manne, Temesgen Samuel and Pran K. Datta
Cancers 2024, 16(20), 3491; https://doi.org/10.3390/cancers16203491 (registering DOI) - 15 Oct 2024
Viewed by 337
Abstract
Among the most popular chemotherapeutic agents, irinotecan, regarded as a prodrug belonging to the camptothecin family that inhibits topoisomerase I, is widely used to treat metastatic colorectal cancer (CRC). Although immunotherapy is promising for several cancer types, only microsatellite-instable (~7%) and not microsatellite-stable [...] Read more.
Among the most popular chemotherapeutic agents, irinotecan, regarded as a prodrug belonging to the camptothecin family that inhibits topoisomerase I, is widely used to treat metastatic colorectal cancer (CRC). Although immunotherapy is promising for several cancer types, only microsatellite-instable (~7%) and not microsatellite-stable CRCs are responsive to it. Therefore, it is important to investigate the mechanism of irinotecan function to identify cellular proteins and/or pathways that could be targeted for combination therapy. Here, we have determined the effect of irinotecan treatment on the expression/activation of tumor suppressor genes (including p15Ink4b, p21Cip1, p27Kip1, and p53) and oncogenes (including OPN, IL8, PD-L1, NF-κB, ISG15, Cyclin D1, and c-Myc) using qRT-PCR, Western blotting, immunofluorescence (IF), and RNA sequencing of tumor specimens. We employed stable knockdown, neutralizing antibodies (Abs), and inhibitors of OPN, p53, and NF-κB to establish downstream signaling and sensitivity/resistance to the cytotoxic activities of irinotecan. Suppression of secretory OPN and NF-κB sensitized colon cancer cells to irinotecan. p53 inhibition or knockdown was not sufficient to block or potentiate SN38-regulated signaling, suggesting p53-independent effects. Irinotecan treatment inhibited tumor growth in syngeneic mice. Analyses of allograft tumors from irinotecan-treated mice validated the cell culture results. RNA-seq data suggested that irinotecan-mediated activation of NF-κB signaling modulated immune and inflammatory genes in mice, which may compromise drug efficacy and promote resistance. In sum, these results suggest that, for CRCs, targeting OPN, NF-κB, PD-L1, and/or ISG15 signaling may provide a potential strategy to overcome resistance to irinotecan-based chemotherapy. Full article
Show Figures

Figure 1

Figure 1
<p>SN38 induces expression of several apoptotic-related genes in addition to the conventional p53 pathway. DLD-1, SW480, and FET cells were treated with SN38 at 100 nM concentration for 48 h. (<b>A</b>) Western blots show higher levels of p21<sup>Cip1</sup>, p27<sup>Kip1</sup>, and Bax and lower levels of c-Myc and CyclinD1 upon SN38 treatment. Bar diagrams show the relative mRNA expression of p15<sup>Ink4b</sup>, p21<sup>Cip1</sup>, p27<sup>Kip1</sup>, and p53 in (<b>B</b>) DLD-1, (<b>C</b>) SW480, and (<b>D</b>) FET cells after SN38 treatment. (<b>E</b>) p53-silenced DLD-1, SW480, and FET cells were treated with SN38 at 50 or 100 nM concentrations for 48 h. Western blots show the p53-independent, SN38-induced differential levels of p21<sup>Cip1</sup>, p27<sup>Kip1</sup>, and Bax. <span class="html-italic">β</span>-Actin served as a loading control. Statistical analysis of the samples was by a Student’s <span class="html-italic">t</span>-test. <span class="html-italic">p</span> &lt; 0.05 was considered to be 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, and **** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>SN38 upregulates pro-oncogenic factors, including survivin, PD-L1, osteopontin, and ISG15. (<b>A</b>) DLD-1, SW480, and FET cells were treated with SN38 at increasing concentrations for 48 h. Conditioned media from samples were concentrated and normalized with cell counts and protein concentrations. Western blots show increased cellular levels of OPN, survivin, and PD-L1 and secreted levels of OPN. (<b>B</b>) p53-silenced DLD-1, SW480, and FET cells were treated with SN38 at 50 or 100 nM concentrations for 48 h. Western blots show the p53-independent, SN38-induced levels of OPN, PD-L1, and survivin. Cells were treated with an OPN Ab (2 μg/mL) in combination with various concentrations of SN38 for 48 h. Graph showing lower cell survival (%) (MTT assay) of (<b>C</b>) SW480 and (<b>D</b>) FET cells. (<b>E</b>) Western blots showing higher levels of ISG15 after 48 h of SN38 treatment of DLD-1, SW480, and FET cells. <span class="html-italic">β</span>-Actin served as a loading control. Statistical analysis of the samples was by a Student’s <span class="html-italic">t</span>-test. <span class="html-italic">p</span> &lt; 0.05 was considered to be 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, and **** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>SN38 induces NF-κB activity and its nuclear localization. DLD-1, SW40, and FET cells were pretreated with Bay (1 μM) or QNZ (1 μM) for 3 h followed by in combination with SN38 (100 nM) for 24 h. (<b>A</b>) Western blots of nuclear lysates show elevated nuclear localization of p-p65 (NF-κB), an effect inhibited by QNZ. PARP served as a nuclear protein loading control. IF staining of (<b>B</b>) DLD-1, (<b>C</b>) SW480, and (<b>D</b>) FET cells show increased nuclear localization of p-p65 (NF-κB), an effect inhibited by QNZ (scale bar = 10 μm). IF data were analyzed, and NF-κB nuclear-positive cells were counted. Graph showing increased numbers of NF-κB-positive cells (%) (<span class="html-italic">n</span> = 1000) in (<b>E</b>) DLD-1, (<b>F</b>) SW480, and (<b>G</b>) FET cells, which were increased by SN38 treatment and partially regulated by Bay and effectively regulated by QNZ. (<b>H</b>) DLD-1, SW480, and FET cells were co-transfected with pGL2-NF-κB-luciferase and CMV-<span class="html-italic">β</span>-Gal followed by treatment with Bay or QNZ in combination with SN38 under the above conditions. Luciferase activity was measured and normalized with <span class="html-italic">β</span>-galactosidase activity. Bar diagram showing the increased relative luciferase activity, which was partially regulated by Bay and effectively regulated by QNZ. (<b>I</b>) Bar diagram showing the cell survival (%) (MTT assay) of Bay- or QNZ-treated DLD-1, SW480, and FET cells upon treatment with SN38 for 48 h. Statistical analysis of the samples was by a Student’s <span class="html-italic">t</span>-test. <span class="html-italic">p</span> &lt; 0.05 was considered to be significant (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>SN38 induces immunomodulatory molecules through non-canonical NF-κB signaling. (<b>A</b>) OPN-silenced DLD-1 cells were treated with SN38 at 50 or 100 nM concentrations for 48 h. Western blotting shows OPN silencing decreases the level of PD-L1. (<b>B</b>) DLD-1 and SW480 cells were treated with PFT<span class="html-italic">α</span> (200 nM) (a p53 transactivation inhibitor) in combination with SN38 for 48 h. Western blots show no effects of p53 inhibition on PD-L1 levels. (<b>C</b>) DLD-1, (<b>D</b>) SW480, and (<b>E</b>) FET cells were pretreated with Bay (1 μM) or QNZ (1 μM) for 3 h followed by in combination with SN38 (100 nM) for 24 h. Bar diagrams showing increased mRNA expressions of IL8, CCL3, CCL5, and RANKL, an effect reduced by QNZ. Relative mRNA expressions of these proteins upon SN38 treatment relative to untreated cells, and the effects of SN38 treatment were compared with SN38 in combination with Bay or QNZ. (<b>F</b>) DLD-1, SW480, and FET cells were pretreated with Bay (1 μM) or QNZ (1 μM) for 3 h followed by in combination with SN38 (100 nM) for 48 h. Western blots show inhibition of SN38 on OPN, survivin, and ISG15. <span class="html-italic">β</span>-Actin served as a loading control. Statistical analysis of the samples was by a Student’s <span class="html-italic">t</span>-test. <span class="html-italic">p</span> &lt; 0.05 was considered to be 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, and **** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Irinotecan regulates tumor growth by differential regulation of pro- and anti-oncogenic factors. MC38 tumor-bearing mice were treated every third day with irinotecan (5 mg/kg) (<span class="html-italic">n</span> = 5) or irinotecan (15 mg/kg) (<span class="html-italic">n</span> = 5) for 24 days. Tumor volumes were calculated by the equation V = L × W<sup>2</sup> × 0.5, where L is the length and W is the width of a tumor. (<b>A</b>) Graph showing the kinetics of tumor growth. (<b>B</b>) Western blotting showing high levels of CD44, OPN, PD-L1, p21<sup>Cip1</sup>, survivin, and ISG15 in tumor lysates after treatment of tumor-bearing mice with irinotecan (15 mg/kg). IF analysis showing increased (<b>C</b>) OPN, (<b>D</b>) p21<sup>Cip1</sup>, (<b>E</b>) p65, (<b>F</b>) survivin, (<b>G</b>) PD-L1, (<b>H</b>) c-Myc, and (<b>I</b>) ISG15 and (<b>J</b>) decreased CyclinD1 in the tumor tissues of mice after treatment with irinotecan (15 mg/kg) (scale bar = 10 μm). <span class="html-italic">β</span>-Actin served as a loading control. Statistical analysis of the samples was by a Student’s <span class="html-italic">t</span>-test. <span class="html-italic">p</span> &lt; 0.05 was considered to be significant (**** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Irinotecan treatment regulates immune and inflammatory genes. RNA sequencing analysis after irinotecan (15 mg/kg) treatment compared to the vehicle control group (<b>A</b>) showing the expression of 3518 upregulated and 3650 downregulated genes. (<b>B</b>) Venn diagram showing the expression of 590 distinctive genes after irinotecan treatment and 560 distinctive genes in the vehicle control group. (<b>C</b>) GO pathway analysis showing the differential expression of various genes involved in biological processes, cellular components, and metabolic pathways. (<b>D</b>,<b>E</b>) Heatmaps showing high expressions of Sox2, p53, c-Myc, PD-L1, Snail, OPN, Oct4, p15<sup>Ink4b</sup>, Bax, survivin, p21<sup>Cip1</sup>, and Slug after treatment with irinotecan. Bar diagram showing mRNA expression from RNA seq data (<b>F</b>) increased TLR, (<b>G</b>) CXCR, (<b>H</b>) CXCL, (<b>I</b>) CCL, and (<b>J</b>) interleukins and their receptors (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>Graphical summary: p53-independent effects of irinotecan in colorectal cancer. ↑ Denotes upregulation and ↓ denotes downregulation of protein expression or signaling.</p>
Full article ">
14 pages, 3014 KiB  
Article
Effects of Dietary Yeast β-1,3/1,6-D-Glucan on Immunomodulation in RAW 264.7 Cells and Methotrexate-Treated Rat Models
by Joohee Son, Yeseul Hwang, Eun-Mi Hong, Marion Schulenberg, Hyungyung Chai, Hee-Geun Jo and Donghun Lee
Int. J. Mol. Sci. 2024, 25(20), 11020; https://doi.org/10.3390/ijms252011020 - 14 Oct 2024
Viewed by 353
Abstract
A new subclass of nutraceuticals, called immunoceuticals, is dedicated to immunological regulation. Although yeast-derived β-1,3/1,6-D-glucan shows promise as an immunoceutical candidate, further studies are needed to define its precise immune-enhancing processes and to standardize its use. Following methotrexate (MTX)-induced immunosuppression in rats, we [...] Read more.
A new subclass of nutraceuticals, called immunoceuticals, is dedicated to immunological regulation. Although yeast-derived β-1,3/1,6-D-glucan shows promise as an immunoceutical candidate, further studies are needed to define its precise immune-enhancing processes and to standardize its use. Following methotrexate (MTX)-induced immunosuppression in rats, we evaluated the immunomodulatory efficacy of a highly pure and standardized β-1,3/1,6-D-glucan sample (YBG) in RAW 264.7 macrophages. In in vitro and in vivo models, YBG demonstrated remarkable immunomodulatory effects, such as repair of immune organ damage, elevation of blood cytokine levels, and enhanced phagocytosis and nitric oxide production in RAW 264.7 cells. These results are consistent with the established immunostimulatory properties of β-glucan. It is noteworthy that this research indicates the potential of YBG as an immunomodulatory nutraceutical, as it is among the first to demonstrate immunological augmentation in an immunosuppression setting produced by MTX. Based on these observations, further investigation of YBG is warranted, particularly given its potential to emerge as a combination immunoceutical to mitigate immunosuppression and reduce the risk of infection in rheumatoid arthritis (RA) patients receiving long-term MTX therapy. Full article
(This article belongs to the Special Issue Nutrients and Active Substances in Natural Products)
Show Figures

Figure 1

Figure 1
<p>Effects of the YBG on body weight and organ indices in MTX-induced immunosuppressed rats. The SD rats were allocated into four groups, each consisting of eight animals. DW was utilized as a vehicle, and rats in YBG groups were given oral YBG (45, 90 mg/kg/day) once daily for 14 days. Subsequently, MTX (2 mg/kg/day, po) was administered for three consecutive days to induce immunosuppression. (<b>A</b>) The body weight of SD rats, (<b>B</b>) spleen index, and (<b>C</b>) thymus index. All results are shown as the mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01 in comparison to the sham group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s post hoc test. CON: control; YBG: yeast β-glucan; MTX: methotrexate.</p>
Full article ">Figure 2
<p>Effects of YBG on blood leukocyte number in MTX-induced immunosuppressed rats. SD rats were treated with distilled water or oral YBG (45, 90 mg/kg/day) for 14 days. Subsequently, MTX (2 mg/kg/day, po) was given for 3 consecutive days to induce immunosuppression. (<b>A</b>) Total white blood cells, (<b>B</b>) neutrophils, (<b>C</b>) lymphocytes, and (<b>D</b>) monocytes. All results are shown as mean ± SEM. # <span class="html-italic">p</span> &lt; 0.05 and ### <span class="html-italic">p</span> &lt; 0.001 in comparison to the sham group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s post hoc test. CON: control; YBG: yeast β-glucan; MTX: methotrexate.</p>
Full article ">Figure 3
<p>Effects of YBG on splenocyte proliferation induced by Con-A and LPS in MTX-induced immunosuppressed rats. SD rats were given saline or oral YBG (45, 90 mg/kg/day) for 14 days. Subsequently, MTX (2 mg/kg/day, po) was administered orally for 3 consecutive days to induce immunosuppression. (<b>A</b>) Con-A induced T lymphocyte proliferation; (<b>B</b>) LPS induced B lymphocyte proliferation. All results are shown as mean ± SEM. # <span class="html-italic">p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 in comparison to the sham group, * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s post hoc test. CON: control; YBG: yeast β-glucan; MTX: methotrexate;.</p>
Full article ">Figure 4
<p>Effects of YBG on NK cell activity in MTX-induced immunosuppressed rats. SD rats were administered distilled water or oral YBG (45, 90 mg/kg/day) for 14 days. Subsequently, MTX (2 mg/kg/day, po) was given for 3 consecutive days to induce immunosuppression. NK-sensitive fibroblast, YAC-1 cells were utilized as target cells. All results are shown as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s post hoc test. CON: control; YBG: yeast β-glucan; MTX: methotrexate.</p>
Full article ">Figure 5
<p>The effect of YBG on cytokine release in MTX-induced immunosuppressed rats. SD rats were given distilled water or oral YBG (45, 90 mg/kg/day) for 14 days. Subsequently, MTX (2 mg/kg/day) was administrated orally for 3 days to induce immunosuppression. (<b>A</b>) mRNA expression of IL-2 and (<b>B</b>) IL-6 in the spleen; (<b>C</b>) serum TNF-α level. qRT-PCR was utilized to examine the mRNA expression levels of IL-2 and IL-6, while serum TNF-α levels were measured using ELISA. The data are shown as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 in comparison to the sham group; * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s test; CON: control; YBG: yeast β-glucan; MTX: methotrexate; IL: interleukin; TNF: tumor necrosis factor.</p>
Full article ">Figure 6
<p>The effect of YBG on immunoglobulin production in MTX-induced immunosuppressed rats. SD rats were given distilled water or oral YBG (45, 90 mg/kg/day) for 14 days. Subsequently, MTX (2 mg/kg/day) was orally administered for 3 days to induce immunosuppression. (<b>A</b>) IgG and (<b>B</b>) IgM levels in the serum were estimated using ELISA. All results are shown as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s test. CON: control; YBG: yeast β-glucan; Ig: immunoglobulin.</p>
Full article ">Figure 7
<p>Effects of the YBG on cell viability and phagocytic activity in RAW 264.7. (<b>A</b>) cell viability and (<b>B</b>) phagocytosis. The phagocytic activity was examined using a neutral red uptake assay. All results are shown as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 in comparison to the CON group by one-way ANOVA and Dunnett’s test. CON: control; YBG: yeast β-glucan.</p>
Full article ">Figure 8
<p>Effects of the YBG on nitric-oxide production and proinflammatory factors of RAW 264.7. (<b>A</b>) NO production and mRNA expressions of (<b>B</b>) NOS2, (<b>C</b>) COX-2, and (<b>D</b>) TNF-α; (<b>E</b>) protein secretion of iNOS, COX-2, and TNF- α from RAW 264.7. The mRNA expression levels were examined using qRT-PCR, while protein productions were estimated using a Western blot assay. All results are shown as mean ± SEM. * <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; 0.001 in comparison to the CON group by one-way ANOVA and Dunnett’s test, CON: control; YBG: yeast β-glucan; NOS2: nitric oxide synthase 2; COX-2: cyclooxygenase 2; TNF: tumor necrosis factor.</p>
Full article ">
17 pages, 2391 KiB  
Article
Limosilactobacillus reuteri DSM 17938 Produce Bioactive Components during Formulation in Sucrose
by Ludwig Ermann Lundberg, Manuel Mata Forsberg, James Lemanczyk, Eva Sverremark-Ekström, Corine Sandström, Stefan Roos and Sebastian Håkansson
Microorganisms 2024, 12(10), 2058; https://doi.org/10.3390/microorganisms12102058 - 12 Oct 2024
Viewed by 448
Abstract
Improved efficacy of probiotics can be achieved by using different strategies, including the optimization of production parameters. The impact of fermentation parameters on bacterial physiology is a frequently investigated topic, but what happens during the formulation, i.e., the step where the lyoprotectants are [...] Read more.
Improved efficacy of probiotics can be achieved by using different strategies, including the optimization of production parameters. The impact of fermentation parameters on bacterial physiology is a frequently investigated topic, but what happens during the formulation, i.e., the step where the lyoprotectants are added prior to freeze-drying, is less studied. In addition to this, the focus of process optimization has often been yield and stability, while effects on bioactivity have received less attention. In this work, we investigated different metabolic activities of the probiotic strain Limosilactobacillus reuteri DSM 17938 during formulation with the freeze-drying protectant sucrose. We discovered that the strain consumed large quantities of the added sucrose and produced an exopolysaccharide (EPS). Using NMR, we discovered that the produced EPS was a glucan with α-1,4 and α-1,6 glycosidic bonds, but also that other metabolites were produced. The conversion of the lyoprotectant is hereafter designated lyoconversion. By also analyzing the samples with GCMS, additional potential bioactive compounds could be detected. Among these were tryptamine, a ligand for the aryl hydrocarbon receptor, and glycerol, a precursor for the antimicrobial compound reuterin (3-hydroxypropionaldehyde). To exemplify the bioactivity potential of lyoconversion, lyoconverted samples as well as purified EPS were tested in a model for immunomodulation. Both lyoconverted samples and purified EPS induced higher expression levels of IL-10 (2 times) and IL-6 (4–6 times) in peripheral blood mononuclear cells than non-converted control samples. We further found that the initial cultivation of DSM 17938 with sucrose as a sugar substrate, instead of glucose, improved the ability to convert sucrose in the lyoprotectant into EPS and other metabolites. Lyoconversion did not affect the viability of the bacteria but was detrimental to freeze-drying survival, an issue that needs to be addressed in the future. In conclusion, we show that the metabolic activities of the bacteria during the formulation step can be used as a tool to alter the activity of the bacteria and thereby potentially improve probiotic efficacy. Full article
(This article belongs to the Section Microbial Biotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic overview of samples and methods used in this work.</p>
Full article ">Figure 2
<p>1D <sup>1</sup>H NMR spectra recorded at 333 K of freeze-drying supernatants of sucrose- and glucose-grown bacteria, either incubated in lyoprotectant at room temperature overnight (RT) or directly frozen (DF).</p>
Full article ">Figure 3
<p>Anomeric region of the 1D 1H NMR spectra. Sucrose- and glucose-grown bacteria were incubated at room temperature overnight (RT) or directly frozen (DF). The chemical shifts of the anomeric proton signal of glucose in sucrose, in α-(1-4)-linked EPSs, free α-Glc, and α-(1-6)-linked EPSs are indicated in the table (<b>A</b>). The NMR spectra are reported in (<b>B</b>). ? indicates a peak that could not be annotated.</p>
Full article ">Figure 4
<p>The relative amount of sucrose, glucose, and fructose in the lyoconverted samples determined by GCMS. Significance levels used were ** <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 <span class="html-italic">p</span> &gt; 0.05. Prism GraphPad version 9.0 was used for statistical analysis.</p>
Full article ">Figure 5
<p>GCMS metabolomic data displayed as a PCA plot with sample distributions and relative concentrations of selected metabolites. (<b>A</b>) Sample distributions show four distinct clusters corresponding to the sample treatments. (<b>B</b>) Briefly, the sucrose-grown lyoconverted samples had higher relative concentrations of most of the identified metabolites including glycerol, glyceraldehyde, tryptamine, and malonic acid, whereas glucose-grown lyoconverted samples had increased relative concentrations of mevalonic acid. Significance levels used were *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. Prism GraphPad version 9.0 was used for statistical analysis.</p>
Full article ">Figure 6
<p>The effect of lyoconversion on pH (<b>A</b>), concentrations of organic acids (<b>B</b>), CFU after formulation and freezing (<b>C</b>), and CFU after freeze-drying (<b>D</b>). Samples incubated at room temperature overnight are denoted as RT, and direct-frozen samples are denoted as DF. Analysis of organic acids by HPLC indicated metabolic activity during the lyoconversion, while the concentrations of lactate and acetate increased substantially in the RT samples, especially for sucrose-grown bacteria. Significance levels used were * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; **** <span class="html-italic">p</span> &lt; 0.0001, ns <span class="html-italic">p</span> &gt; 0.05. Prism GraphPad version 9.0 was used for statistical analysis.</p>
Full article ">Figure 7
<p>Induction of IL-6 and IL-10 by the different types of samples was measured in unstimulated PBMCs. Cytokine secretion was significantly increased in response to lyoconverted (RT) samples compared to direct-frozen (DF) samples. The multiplicity of bacteria equivalents was 10 for IL-6 secretion and 100 for IL-10 secretion. Friedman’s test was used for statistical analysis. Significance levels used were * <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. Prism GraphPad version 9.0 was used for statistical analysis.</p>
Full article ">Figure 8
<p>IL-6 and IL-10 induction in PBMCs after incubation with purified EPSs from sucrose-grown lyoconverted samples. Both IL-6 and IL-10 levels were significantly increased compared to unstimulated immune cells. The multiplicity of bacterial equivalent was 100. Wilcoxon matched-pair signed-rank test was used for statistical analysis. Significance levels used were * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01. Prism GraphPad version 9.0 was used for statistical analysis.</p>
Full article ">
16 pages, 1299 KiB  
Review
State of the Field: Cytotoxic Immune Cell Responses in C. neoformans and C. deneoformans Infection
by Elizabeth C. Okafor and Kirsten Nielsen
J. Fungi 2024, 10(10), 712; https://doi.org/10.3390/jof10100712 - 12 Oct 2024
Viewed by 271
Abstract
Cryptococcus neoformans is an environmental pathogen that causes life-threatening disease in immunocompromised persons. The majority of immunological studies have centered on CD4+ T-cell dysfunction and associated cytokine signaling pathways, optimization of phagocytic cell function against fungal cells, and identification of robust antigens [...] Read more.
Cryptococcus neoformans is an environmental pathogen that causes life-threatening disease in immunocompromised persons. The majority of immunological studies have centered on CD4+ T-cell dysfunction and associated cytokine signaling pathways, optimization of phagocytic cell function against fungal cells, and identification of robust antigens for vaccine development. However, a growing body of literature exists regarding cytotoxic cells, specifically CD8+ T-cells, Natural Killer cells, gamma/delta T-cells, NK T-cells, and Cytotoxic CD4+ T-cells, and their role in the innate and adaptive immune response during C. neoformans and C. deneoformans infection. In this review, we (1) provide a comprehensive report of data gathered from mouse and human studies on cytotoxic cell function and phenotype, (2) discuss harmonious and conflicting results on cellular responses in mice models and human infection, (3) identify gaps of knowledge in the field ripe for exploration, and (4) highlight how innovative immunological tools could enhance the study of cytotoxic cells and their potential immunomodulation during cryptococcosis. Full article
(This article belongs to the Special Issue Fungal Immunology and Vaccinology)
Show Figures

Figure 1

Figure 1
<p>Murine and Human CD8<sup>+</sup> T-cell Sensing, Signaling, and Response to <span class="html-italic">C. neoformans</span> and <span class="html-italic">C. deneoformans.</span> Summary of the current body of literature regarding the mechanism of (<b>A</b>) murine and (<b>B</b>) human CD8<sup>+</sup> T-cell targeting, degranulation, and cytokine secretion after exposure to <span class="html-italic">C</span>. <span class="html-italic">neoformans</span> or <span class="html-italic">deneoformans.</span> (<b>A</b>) Despite depletion of murine CD8<sup>+</sup> T-cells in the lung and systemic circulation leading to increased CFUs, the exact receptor(s) and fungal ligand(s) which trigger cytotoxicity and the cytolytic proteins secreted are unknown. During infection, infiltrating pulmonary and systemic CD8<sup>+</sup> T-cells are robust producers of IFN-γ. (<b>B</b>) Similar to murine CD8<sup>+</sup> T-cells, the exact receptor(s) and fungal ligand(s) which facilitate cytotoxicity of human CD8<sup>+</sup> T-cells are unknown. Upon exposure to fungal cells, human CD8<sup>+</sup> T-cells increase transcription of the cytolytic protein granulysin and depletion of granulysin abrogates CD8<sup>+</sup> T-cell anti-cryptococcal effects. The mechanism by which granules containing granulysin are selectively secreted, over those containing perforin or other cytolytic molecules, remains to be determined. IL-15 stimulation enhances human CD8<sup>+</sup> T-cell anti-cryptococcal effects while IL-2R expression increases upon coculture with fungal cells. However, the cytokines released by activated CD8<sup>+</sup> T-cells during human cryptococcal meningitis have not been identified.</p>
Full article ">Figure 2
<p>Murine and Human NK cell Sensing, Signaling, and Response to <span class="html-italic">C. neoformans</span> and <span class="html-italic">C. deneoformans.</span> Summary of the current body of literature regarding the mechanism of (<b>A</b>) murine and (<b>B</b>) human NK cell targeting, degranulation, and cytokine secretion after exposure to <span class="html-italic">C</span>. <span class="html-italic">neoformans</span> or <span class="html-italic">deneoformans</span>. (<b>A</b>) Murine NK cells interact with encapsulated <span class="html-italic">Cryptococcus</span> cells via “microvilli” or cellular protrusions, though the receptor or integrin interacting with the fungal cell capsule remains unknown. Though anti-cryptococcal activity of murine NK cells has been documented, the mechanism is undetermined but likely involves intracellular signaling pathways which are not mediated by DAP12 motifs. Murine NK cells are robust producers of the inflammatory cytokine IFN-γ during infection; the production of other inflammatory cytokines is unknown. (<b>B</b>) Human NK cell activating receptor NKp30 recognizes cryptococcal antigen β-1,3-glucan on the fungal cell wall which leads to the phosphorylation of intracellular tyrosine motifs by Src family kinases Fyn and Lyn and increase perforin transcription. Stimulation with IL-12 enhances NKp30 expression on NK cells from donors with HIV. The downstream signaling cascade remains to be elucidated. Granules containing granulysin and perforin are transported to the NK cell membrane via Erg5-kinesin. Perforin is released at the immunological synapse facilitating NK cell anti-cryptococcal effects and reducing fungal growth. However, it is unclear how perforin penetrates the <span class="html-italic">Cryptococcus</span> fungal cell wall. Upon exposure to Cryptococcus cells, NK cells increase secretion of chemoattractant IP-10 and reduce secretion of the inflammatory cytokine TNF-α and activating cytokine GM-CSF.</p>
Full article ">
24 pages, 1325 KiB  
Article
Did We Overreact? Insights on COVID-19 Disease and Vaccination in a Large Cohort of Immune-Mediated Inflammatory Disease Patients during Sequential Phases of the Pandemic (The BELCOMID Study)
by Jeroen Geldof, Marie Truyens, João Sabino, Marc Ferrante, Jo Lambert, Hilde Lapeere, Tom Hillary, An Van Laethem, Kurt de Vlam, Patrick Verschueren, Triana Lobaton, Elizaveta Padalko and Séverine Vermeire
Vaccines 2024, 12(10), 1157; https://doi.org/10.3390/vaccines12101157 - 11 Oct 2024
Viewed by 628
Abstract
Introduction: As the COVID-19 pandemic becomes an endemic state, still many questions remain regarding the risks and impact of SARS-CoV-2 infection and vaccination in patients with immune-mediated inflammatory diseases (IMIDs) who were excluded from the phase 3 COVID-19 vaccination trials. Methods: The BELCOMID [...] Read more.
Introduction: As the COVID-19 pandemic becomes an endemic state, still many questions remain regarding the risks and impact of SARS-CoV-2 infection and vaccination in patients with immune-mediated inflammatory diseases (IMIDs) who were excluded from the phase 3 COVID-19 vaccination trials. Methods: The BELCOMID study collected patient data and serological samples from a large, multicentric IMID patient cohort that was prospectively followed during sequential stages of the pandemic. Patients were stratified according to vaccination status into five groups across three sampling periods. Interactions between SARS-CoV-2 infection, COVID-19 vaccination status, IMID-treatment modalities and IMID course were explored. Results: In total, 2165 patients with IBD, a dermatological or rheumatological IMID participated. SARS-CoV-2 infection rates increased over the course of the pandemic and were highest in IMID patients that had refused every vaccine. After baseline COVID-19 vaccination, serologic spike (S)-antibody responses were attenuated by particular types of immune-modulating treatment: anti-TNF, rituximab, JAKi, systemic steroids, combined biologic/immunomodulator treatment. Nonetheless, S-antibody concentration increased progressively in patients who received a booster vaccination, reaching 100% seroconversion rate in patients who had received two booster vaccines. Previous SARS-CoV-2 infection was found as a predictor of higher S-antibody response. Patients who had refused every vaccine showed the lowest rates of S-seroconversion (53.8%). Multiple logistic regression did not identify previous SARS-CoV-2 infection as a risk factor for IMID flare-up. Furthermore, no increased risk of IMID flare-up was found with booster vaccination. Conclusions: Altogether, the BELCOMID study provides evidence for the efficacy and safety of COVID-19 vaccination and confirms the importance of repeated booster vaccination in IMID patients. Full article
(This article belongs to the Special Issue Immunotherapy and Vaccine Development for Viral Diseases)
Show Figures

Figure 1

Figure 1
<p>BELCOMID timeline. Interim results from inclusion periods 1 and 2 were published previously [<a href="#B7-vaccines-12-01157" class="html-bibr">7</a>]. The current manuscript considers all 3 inclusion periods and 5 vaccination groups.</p>
Full article ">Figure 2
<p>Estimated mean S-serology evolution across vaccination groups.</p>
Full article ">
22 pages, 6432 KiB  
Article
Priming Mesenchymal Stem Cells with Lipopolysaccharide Boosts the Immunomodulatory and Regenerative Activity of Secreted Extracellular Vesicles
by Aina Areny-Balagueró, Marta Camprubí-Rimblas, Elena Campaña-Duel, Anna Solé-Porta, Adrián Ceccato, Anna Roig, John G. Laffey, Daniel Closa and Antonio Artigas
Pharmaceutics 2024, 16(10), 1316; https://doi.org/10.3390/pharmaceutics16101316 - 10 Oct 2024
Viewed by 401
Abstract
Background: Mesenchymal stem cells (MSCs)-derived extracellular vesicles (EVs) have been proposed as an alternative to live-cell administration for Acute Respiratory Distress Syndrome (ARDS). MSC-EVs can be chiefly influenced by the environment to which the MSCs are exposed. Here, lipopolysaccharide (LPS) priming of MSCs [...] Read more.
Background: Mesenchymal stem cells (MSCs)-derived extracellular vesicles (EVs) have been proposed as an alternative to live-cell administration for Acute Respiratory Distress Syndrome (ARDS). MSC-EVs can be chiefly influenced by the environment to which the MSCs are exposed. Here, lipopolysaccharide (LPS) priming of MSCs was used as a strategy to boost the natural therapeutic potential of the EVs in acute lung injury (ALI). Methods: The regenerative and immunemodulatory effect of LPS-primed MSC-EVs (LPS-EVs) and non-primed MSC-EVs (C-EVs) were evaluated in vitro on alveolar epithelial cells and macrophage-like THP-1 cells. In vivo, ALI was induced in adult male rats by the intrapulmonary instillation of HCl and LPS. Rats (n = 8 to 22/group) were randomized to receive a single bolus (1 × 108 particles) of LPS-EVs, C-EVs, or saline. Lung injury severity was assessed at 72 h in lung tissue and bronchoalveolar lavage. Results: In vitro, LPS-EVs improved wound regeneration and attenuated the inflammatory response triggered by the P. aeruginosa infection, enhancing the M2 macrophage phenotype. In in vivo studies, LPS-EVs, but not C-EVs, significantly decreased the neutrophilic infiltration and myeloperoxidase (MPO) activity in lung tissue. Alveolar macrophages from LPS-EVs-treated animals exhibited a reduced expression of CXCL-1, a key neutrophil chemoattractant. However, both C-EVs and LPS-EVs reduced alveolar epithelial and endothelial permeability, mitigating lung damage. Conclusions: EVs from LPS-primed MSCs resulted in a better resolution of ALI, achieving a greater balance in neutrophil infiltration and activation, while avoiding the complete disruption of the alveolar barrier. This opens new avenues, paving the way for the clinical implementation of cell-based therapies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization of C-MSCs and LPS-primed MSCs isolated from rat bone marrow. (<b>A</b>) Immunophenotype of C-MSCs and LPS-MSCs, staining CD105 (FITC, green), CD90 (FITC, green) and CD44 (Texas Red, red) markers. Quantification of the percentage of positive cells for each marker (<span class="html-italic">n</span> = 7–11). The nuclei of the cells were stained with Hoechst (UV light, blue), 20× magnification. Scale bar: 100 µm. (<b>B</b>) Representative images of undifferentiated MSCs and MSCs differentiation towards adipogenic, osteogenic and chondrogenic lineages in vitro. Magnification: 10×. Scale bar: 500 µm. (<b>C</b>) Quantification of the EVs-like particles (<b>right</b>) (<span class="html-italic">n</span> = 5) and protein (<b>left</b>) (<span class="html-italic">n</span> = 7) concentration in C-MSCs and LPS-MSCs media. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 3); * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01. Abbreviations: extracellular vesicles, EVs.</p>
Full article ">Figure 2
<p>Characterization of C- and LPS-MSCs-derived EVs. (<b>A</b>) Representative particle size distributions of C-EVs and LPS-EVs’ pools by Nanosight analysis. (<b>B</b>) Representative cryo-TEM images of C-EVs and LPS-EVs, 12kX magnification. Scale bar: 0.5 µm. (<b>C</b>) Detection of Alix, TSG101 and CD81 surface markers in C-EVs and LPS-EVs by Western Blot analysis. Detection of Calnexin only in cell lysate samples as negative control. Abbreviations: cell lysate, C. Lys.; extracellular vesicles, EVs.</p>
Full article ">Figure 3
<p>Effect of MSCs-derived EVs on wound healing and cell proliferation in vitro. (<b>A</b>) Representative optical images of wound healing in HPAEpiC monolayer. Magnification: 20×. Scale bar: 50 µm. Percentage of wound closure in HPAEpiC 24 h after being treated with C-EVs and LPS-EVs. (<b>B</b>) Percentage of cell viability of HPAEpiC 24 h after being treated with C-EVs and LPS-EVs, considering that non-treated cells (control) had 100% cellular viability. Data are presented as mean ± SEM of eight (<b>A</b>) and four (<b>B</b>) independent experiments with two replicates of each condition; * <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. Abbreviations: extracellular vesicles, EVs.</p>
Full article ">Figure 4
<p>Effect of MSCs-derived EVs on infected THP-1 cells in vitro. Representation of mRNA expression of pro-inflammatory cytokines, chemoattractant mediators, and M1 and M2 macrophage phenotype markers: IL-1β (<b>A</b>), IL-6 (<b>B</b>), IL-8 (<b>C</b>), CD86 (<b>D</b>) and CD206 (<b>E</b>) in THP-1 cells activated with LPS and infected by <span class="html-italic">P. aeruginosa</span> and treated with C-EVs or LPS-EVs. (<b>F</b>) Ratio of CD86/CD206 expression (M1/M2 ratio). The relative expression of target genes was normalized to RPL37a expression. Data are presented as mean ± SEM of six independent experiments with two or three replicates of each condition. * <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. Abbreviations: interleukin, IL; extracellular vesicles, EVs.</p>
Full article ">Figure 5
<p>Effect of MSCs-derived EVs on animals’ weight and lung permeability. (<b>A</b>) Monitoring of the animals’ body weight every 24 h, considering 100% as the starting body weight for each group (# <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01 control group vs. HCl + LPS group). (<b>B</b>) Ratio of lung weight/body weight measured at the end of the experiment (grams/grams) (<span class="html-italic">n</span> = 12–22). (<b>C</b>) Total protein concentration (µg/mL) and (<b>D</b>) cells’ concentration (cells/µL) in the BAL fluid at the end point (72 h) (<span class="html-italic">n</span> = 8–13); Data are presented as mean ± SEM * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001. Abbreviations: extracellular vesicles, EVs.</p>
Full article ">Figure 6
<p>Effect of MSCs-derived EVs on neutrophil infiltration in the intra-alveolar space in vivo at 72 h. Percentage of (<b>A</b>) monocytes, (<b>B</b>) neutrophils and (<b>C</b>) lymphocytes in BAL by flow cytometry. (<b>D</b>) Representative images of BAL cells cytospins stained with Diff-Quick. Neutrophils (red arrows), macrophages (green arrows) and lymphocytes (yellow arrows) are indicated in each image. Magnification: 20×. (<b>E</b>) MPO activity quantification in lung tissue homogenates (mU/total tissue protein (g)). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 8–14); * <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 7
<p>Effect of MSCs-derived EVs on lung tissue inflammation in vivo at 72 h. mRNA expression of pro-inflammatory cytokines, (<b>A</b>) IL-1β and (<b>B</b>) IL-6, and chemoattractant mediators, (<b>C</b>) CCL-2 and (<b>D</b>) CXCL-1 (mRNA expression correlated vs. GAPDH). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 9–14). * <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. Abbreviations: interleukin, IL.</p>
Full article ">Figure 8
<p>Effect of MSCs-derived EVs on alveolar macrophage inflammation in vivo at 72 h. mRNA expression of pro-inflammatory cytokines, (<b>A</b>) IL-1β; chemoattractant mediators, (<b>B</b>) CXCL-1; and M2 phenotype markers, (<b>C</b>) Arg-1 and (<b>D</b>) MR (mRNA expression correlated vs. GAPDH). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 8–14). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01. Abbreviations: interleukin, IL; arginase 1, Arg-1; mannose receptor, MR.</p>
Full article ">Figure 9
<p>Effect of MSCs-derived EVs on the lung injury in vivo at 72 h. Representative images of H&amp;E histological lung sections of control and injured animals with (<b>A</b>) 2.5× and (<b>B</b>) 10× magnification. Scale bars: 200 and 100 µm, respectively. (<b>C</b>) Quantification of the lung injury score (LIS), evaluating hemorrhage, peribronchial infiltration, interstitial edema, pneumocyte hyperplasia and intra-alveolar infiltration, as described in <a href="#app1-pharmaceutics-16-01316" class="html-app">Supplementary Table S3</a>. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 9–14). *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
13 pages, 1352 KiB  
Review
Immune Modulation by Epstein–Barr Virus Lytic Cycle: Relevance and Implication in Oncogenesis
by Nevena Todorović, Maria Raffaella Ambrosio and Amedeo Amedei
Pathogens 2024, 13(10), 876; https://doi.org/10.3390/pathogens13100876 - 8 Oct 2024
Viewed by 609
Abstract
EBV infects more than 90% of people globally, causing lifelong infection. The phases of the EBV life cycle encompass primary infection, latency, and subsequent reactivation or lytic phase. The primary infection usually happens without noticeable symptoms, commonly in early life stages. If it [...] Read more.
EBV infects more than 90% of people globally, causing lifelong infection. The phases of the EBV life cycle encompass primary infection, latency, and subsequent reactivation or lytic phase. The primary infection usually happens without noticeable symptoms, commonly in early life stages. If it manifests after childhood, it could culminate in infectious mononucleosis. Regarding potential late consequences, EBV is associated with multiple sclerosis, rheumatoid arthritis, chronic active EBV infection, lymphomas, and carcinomas. Previous reports that the lytic phase plays a negligible or merely secondary role in the oncogenesis of EBV-related tumors are steadily losing credibility. The right mechanisms through which the lytic cycle contributes to carcinogenesis are still unclear, but it is now recognized that lytic genes are expressed to some degree in different cancer-type cells, implicating their role here. The lytic infection is a persistent aspect of virus activity, continuously stimulating the immune system. EBV shows different strategies to modulate and avoid the immune system, which is thought to be a key factor in its ability to cause cancer. So, the principal goal of our review is to explore the EBV’s lytic phase contribution to oncogenesis. Full article
(This article belongs to the Special Issue Oncogenic Viruses)
Show Figures

Figure 1

Figure 1
<p>Presentation of EBV-related oncogenesis with correlated mechanisms, induced by its lytic cycle.</p>
Full article ">
21 pages, 619 KiB  
Review
Investigating the Anti-Inflammatory Activity of Various Brown Algae Species
by Selin Ersoydan and Thomas Rustemeyer
Mar. Drugs 2024, 22(10), 457; https://doi.org/10.3390/md22100457 - 5 Oct 2024
Viewed by 638
Abstract
This literature review investigated the anti-inflammatory properties of brown algae, emphasizing their potential for dermatological applications. Due to the limitations and side effects associated with corticosteroids and immunomodulators, interest has been growing in harnessing therapeutic qualities from natural products as alternatives to traditional [...] Read more.
This literature review investigated the anti-inflammatory properties of brown algae, emphasizing their potential for dermatological applications. Due to the limitations and side effects associated with corticosteroids and immunomodulators, interest has been growing in harnessing therapeutic qualities from natural products as alternatives to traditional treatments for skin inflammation. This review explored the bioactive compounds in brown algae, specifically looking into two bioactive compounds, namely, fucoidans and phlorotannins, which are widely known to exhibit anti-inflammatory properties. This review synthesized the findings from various studies, highlighting how these compounds can mitigate inflammation by mechanisms such as reducing oxidative stress, inhibiting protein denaturation, modulating immune responses, and targeting inflammatory pathways, particularly in conditions like atopic dermatitis. The findings revealed species-specific variations influenced by the molecular weight and sulphate content. Challenges related to skin permeability were addressed, highlighting the potential of nanoformulations and penetration enhancers to improve delivery. While the in vivo results using animal models provided positive results, further clinical trials are necessary to confirm these outcomes in humans. This review concludes that brown algae hold substantial promise for developing new dermatological treatments and encourages further research to optimize extraction methods, understand the molecular mechanisms, and address practical challenges such as sustainability and regulatory compliance. This review contributes to the growing body of evidence supporting the integration of marine-derived compounds into therapeutic applications for inflammatory skin diseases. Full article
(This article belongs to the Special Issue From Sea to Skin: Advancements in Marine-Based Cosmeceuticals)
Show Figures

Figure 1

Figure 1
<p>Inclusion and exclusion criteria.</p>
Full article ">
21 pages, 18608 KiB  
Article
Distribution of Immunomodulation, Protection and Regeneration Factors in Cleft-Affected Bone and Cartilage
by Mārtiņš Vaivads and Māra Pilmane
Diagnostics 2024, 14(19), 2217; https://doi.org/10.3390/diagnostics14192217 - 4 Oct 2024
Viewed by 431
Abstract
Background: Craniofacial clefts can form a significant defect within bone and cartilage, which can negatively affect tissue homeostasis and the remodeling process. Multiple proteins can affect supportive tissue growth, while also regulating local immune response and tissue protection. Some of these factors, like [...] Read more.
Background: Craniofacial clefts can form a significant defect within bone and cartilage, which can negatively affect tissue homeostasis and the remodeling process. Multiple proteins can affect supportive tissue growth, while also regulating local immune response and tissue protection. Some of these factors, like galectin-10 (Gal-10), nuclear factor kappa-light-chain-enhancer of activated B cells protein 65 (NF-κB p65), heat shock protein 60 (HSP60) and 70 (HSP70) and cathelicidin (LL-37), have not been well studied in cleft-affected supportive tissue, while more known tissue regeneration regulators like type I collagen (Col-I) and bone morphogenetic proteins 2 and 4 (BMP-2/4) have not been assessed jointly with immunomodulation and protective proteins. Information about the presence and interaction of these proteins in cleft-affected supportive tissue could be helpful in developing biomaterials and improving cleft treatment. Methods: Two control groups and two cleft patient groups for bone tissue and cartilage, respectively, were organized with five patients in each group. Immunohistochemistry with the semiquantitative counting method was implemented to determine Gal-10-, NF-κB p65-, HSP60-, HSP70-, LL-37-, Col-I- and BMP-2/4-positive cells within the tissue. Results: Factor-positive cells were identified in each study group. Multiple statistically significant correlations were identified. Conclusions: A significant increase in HSP70-positive chondrocytes in cleft patients could indicate that HSP70 might be reacting to stressors caused by the local tissue defect. A significant increase in Col-I-positive osteocytes in cleft patients might indicate increased bone remodeling and osteocyte activity due to the presence of a cleft. Correlations between factors indicate notable differences in molecular interactions within each group. Full article
(This article belongs to the Section Pathology and Molecular Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Hematoxylin and eosin (H&amp;E) stained control tissue and cleft-affected supportive tissue. (<b>A</b>) Control bone tissue, H&amp;E, 200×. (<b>B</b>) Cleft-affected bone tissue with surrounding periosteum, H&amp;E, 200×. (<b>C</b>) Control cartilage tissue with surrounding perichondrium, H&amp;E, 200×. (<b>D</b>) Cleft-affected cartilage tissue, H&amp;E, 200×.</p>
Full article ">Figure 2
<p>Immunohistochemistry (IMH) of galectin-10 (Gal-10)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) A few Gal-10-positive osteocytes (arrow) in control group bone tissue, Gal-10 IMH, 200×. (<b>B</b>) A few Gal-10-positive osteocytes (arrows) in cleft-affected bone tissue, Gal-10 IMH, 200×. (<b>C</b>) A few Gal-10-positive chondroblasts and chondrocytes (arrow) in the control group hyaline cartilage, Gal-10 IMH, 200×. (<b>D</b>) A rare occurrence of Gal-10-positive chondrocytes (arrows) in cleft-affected hyaline cartilage, Gal-10 IMH, 200×.</p>
Full article ">Figure 3
<p>Immunohistochemistry of nuclear factor kappa-light-chain-enhancer of activated B cells protein 65 (NF-κB p65)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) A few NF-κB p65-positive osteocytes (arrow) in control bone tissue, NF-κB p65 IMH, 200×. (<b>B</b>) A few NF-κB p65-positive osteocytes (arrows) in cleft-affected bone tissue, NF-κB p65 IMH, 200×. (<b>C</b>) Few to moderate number of NF-κB p65-positive chondrocytes and chondroblasts (arrows) in control hyaline cartilage, NF-κB p65 IMH, 200×. (<b>D</b>) Numerous NF-κB p65-positive chondrocytes in cleft-affected cartilage, NF-κB p65 IMH, 200×.</p>
Full article ">Figure 4
<p>Immunohistochemistry of heat shock protein 60 (HSP60)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) A few HSP60-positive osteocytes (arrows) in bone tissue of a control patient, HSP60 IMH, 200×. (<b>B</b>) A few HSP60-positive osteocytes (arrows) in cleft-affected bone tissue, HSP60 IMH, 200×. (<b>C</b>) Few to moderate number of HSP60-positive chondrocytes and chondroblasts (arrow) in control patient hyaline cartilage, HSP60 IMH, 200×. (<b>D</b>) Moderate to numerous HSP60-positive chondrocytes and chondroblasts in cleft-affected cartilage, HSP60 IMH, 200×.</p>
Full article ">Figure 5
<p>Immunohistochemistry of heat shock protein 70 (HSP70)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) Moderate number of HSP70-positive osteocytes in control bone tissue, HSP70 IMH, 200×. (<b>B</b>) A few HSP70-positive bone cells (arrow) in cleft-affected bone tissue, HSP70 IMH, 200×. (<b>C</b>) Moderate number of HSP70-positive chondrocytes in control hyaline cartilage, HSP70 IMH, 200×. (<b>D</b>) Numerous HSP70-positive chondrocytes in cleft-affected cartilage, HSP70 IMH, 200×.</p>
Full article ">Figure 6
<p>Immunohistochemistry of cathelicidin (LL-37)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) A few LL-37-positive osteocytes (arrows) in control bone tissue, LL-37 IMH, 200×. (<b>B</b>) Few to moderate number of LL-37-positive osteocytes (arrows) in cleft-affected bone tissue, LL-37 IMH, 200×. (<b>C</b>) A few LL-37-positive chondrocytes (arrow) in control cartilage tissue, LL-37 IMH, 200×. (<b>D</b>) Moderate number of LL-37-positive chondrocytes (arrows) in cleft-affected cartilage, LL-37 IMH, 200×.</p>
Full article ">Figure 7
<p>Immunohistochemistry of type I collagen (Col-I)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) A rare occurrence of Col-I-positive osteocytes (arrows) in control bone tissue, Col-I IMH, 200×. (<b>B</b>) A few Col-I-positive osteocytes (arrows) in cleft-affected bone tissue, Col-I IMH, 200×. (<b>C</b>) A few Col-I-positive chondrocytes (arrows) in control hyaline cartilage, Col-I IMH, 200×. (<b>D</b>) Few to moderate number of Col-I-positive chondrocytes (arrows) in cleft-affected hyaline cartilage, Col-I IMH, 200×.</p>
Full article ">Figure 8
<p>Immunohistochemistry of bone morphogenetic protein 2/4 (BMP-2/4)-containing cells in control and cleft-affected supportive tissue. (<b>A</b>) Moderate number of BMP-2/4-positive osteocytes in control bone tissue, BMP-2/4 IMH, 200×. (<b>B</b>) A few BMP-2/4-positive osteocytes (arrows) in cleft-affected bone tissue, BMP-2/4 IMH, 200×. (<b>C</b>) Moderate number of BMP-2/4-positive chondrocytes in control hyaline cartilage, BMP-2/4 IMH, 200×. (<b>D</b>) Numerous to abundant BMP-2/4-positive chondrocytes in cleft-affected hyaline cartilage, BMP-2/4 IMH, 200×.</p>
Full article ">
19 pages, 2775 KiB  
Article
A Customizable Platform to Integrate CAR and Conditional Expression of Immunotherapeutics in T Cells
by Huong T. X. Nguyen, Yabin Song, Satendra Kumar and Fu-Sen Liang
Int. J. Mol. Sci. 2024, 25(19), 10568; https://doi.org/10.3390/ijms251910568 - 30 Sep 2024
Viewed by 729
Abstract
The potential of chimeric antigen receptor (CAR)-based immunotherapy as a promising therapeutic approach is often hindered by the presence of highly immunosuppressive tumor microenvironments (TME). Combination therapies with either co-administration or built-in expression of additional TME-modulating therapeutic molecules to potentiate the functions of [...] Read more.
The potential of chimeric antigen receptor (CAR)-based immunotherapy as a promising therapeutic approach is often hindered by the presence of highly immunosuppressive tumor microenvironments (TME). Combination therapies with either co-administration or built-in expression of additional TME-modulating therapeutic molecules to potentiate the functions of CAR-T cells can cause systemic toxicities due to the lack of control over the delivery of biologics. Here, we present a proof-of-concept engineered platform in human Jurkat T cells that combines CAR with a therapeutic gene circuit capable of sensing β-galactosidase (a reported cancer-associated signal) and subsequently activate the production of customized therapeutic gene products. We have demonstrated the integration of the chemically induced proximity (CIP) and associated signal sensing technologies with CAR in this study. A β-galactosidase-activatable prodrug was designed by conjugating a galactose moiety with a CIP inducer abscisic acid (ABA). We showed that Jurkat T cells engineered with CAR and the ABA-inducible genetic circuits can respond to recombinant β-galactosidase to drive the production and secretion of various immunotherapeutics including an anti-cancer agent, an immunomodulatory cytokine, and immune checkpoint inhibitors. Our design is highly modular and could be adapted to sense different cancer-related signals to locally produce antitumor therapeutics that can potentially boost CAR-T efficacy and persistence. Full article
(This article belongs to the Special Issue Current Molecular Progress on Cell and Gene Therapies)
Show Figures

Figure 1

Figure 1
<p>The design of CAR and inducible gene circuit integrated T cell platform. The programmable in situ production of therapeutic outputs is designed to be triggered by cancer-associated signals to enhance the efficacy and persistence of CAR-expressing T cells.</p>
Full article ">Figure 2
<p>ABA-induced production of sTRAIL in engineered HEK-293T cells. (<b>A</b>) Cartoon illustrates sTRAIL-induced apoptosis in death receptor-expressing cancer cells. (<b>B</b>) ABA-inducible therapeutic gene expression constructs. (<b>C</b>) Induction of sTRAIL in HEK-293T cells in response to ABA (1 μM). The concentration of secreted TRAIL in the cell culture supernatant was measured by ELISA. (<b>D</b>) Percentage of MDA-MB-231 cell apoptosis after being treated with DMSO, recombinant human TRAIL (rhTRAIL), or media containing sTRAIL for 24 h. Target cells were co-stained with Annexin V and propidium iodide to exclude necrotic dead cells. Data are calculated from 3 biological replicates and presented as mean ± S.E.M. **** <span class="html-italic">p</span> ≤ 0.0001, *** <span class="html-italic">p</span> ≤ 0.001, ns, <span class="html-italic">p</span> ≥ 0.05, two-tailed Student <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>Generation and validation of recombinant lentiviral vector-encoding inducible gene cassette. (<b>A</b>) Lentiviral vector of ABA-inducible gene expression system. (<b>B</b>) ABA-induced EGFP expression in transfected HEK-293T cells. Cells were treated with 0.5% DMSO or 1 µM ABA, and the induced EGFP expression was observed under a fluorescence microscope after 24 h. The scale bar is 200 µm. Mean fluorescence intensity (MFI) was calculated from 3 random areas in a well using microscope-integrated cellular analysis tool. Error bars are mean ± S.E.M.</p>
Full article ">Figure 4
<p>ABA-induced expression of pro-inflammatory cytokine IL-12 in engineered Jurkat T cells. (<b>A</b>) Three different configurations of lentiviral vectors were tested. The best construct (3) is outlined in red and used in the following studies. (<b>B</b>) Jurkat T cells were then transduced with viruses encoding different inducible IL-12 (iIL-12) vectors and treated with DMSO or ABA at indicated concentrations for 24 h. IL-12 released in the cell culture supernatant was quantified by ELISA. Data are shown as mean ± S.E.M from 3 technical triplicates. **** <span class="html-italic">p</span> ≤ 0.0001, *** <span class="html-italic">p</span> ≤ 0.001, ns, <span class="html-italic">p</span> ≥ 0.05, two-tailed Student <span class="html-italic">t</span>-test. (<b>C</b>) Dosage response of inducible IL-12 in engineered Jurkat T cells with varying ABA concentrations.</p>
Full article ">Figure 5
<p>ABA-induced expression of pro-inflammatory cytokine IL-12 in CAR<sup>+</sup>iIL-12 Jurkat T cells. (<b>A</b>) Lentiviral vectors encoding HER2-CAR and ABA-inducible IL-12 expression cassette. (<b>B</b>) The comparison of IL-12 expression in Jurkat T cells transduced with a single virus encoding iIL-12 only or two viruses encoding both iIL-2 and CAR (CAR<sup>+</sup> iIL-12) genes. Cells were treated with DMSO or ABA (1 μM), and IL-12 secretion was quantified by ELISA at 24 h post-treatment. Data are shown as mean ± S.E.M from 3 technical replicates. **** <span class="html-italic">p</span> ≤ 0.0001, two-tailed Student <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p>ABA-induced expression of immune checkpoint inhibitor pembrolizumab (α-PD1 mAb) in CAR<sup>+</sup>iPembro Jurkat T cells. (<b>A</b>) Lentiviral constructs for HER2-CAR and ABA-inducible pembrolizumab. (<b>B</b>) CAR<sup>+</sup> T cells were co-cultured with MDA-MB-231 cancer cells at 5:1 E:T ratio in the presence of 500 nM ABA or PBS for 24 h. Jurkat T cells were stained for Pembro using A647-conjugated α-HA antibody. The shift in fluorescence signal towards the right indicated the binding of the secreted Pembro to the PD-1 ligand. (<b>C</b>) The levels of free PD-1 were detected by staining the cells with commercial FITC-conjugated α-PD1 antibody (clone EH12.2H7) and analyzed via flow cytometry. Data are mean ± S.E.M from 3 biological replicates. *** <span class="html-italic">p</span> ≤ 0.001, two-tailed Student <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>ABA-inducible expression of immune checkpoint blockage α-CTLA4 scFv in CAR-expressing Jurkat T cells. (<b>A</b>) Lentiviral constructs for HER2-CAR and ABA-inducible α-CTLA4 scFv. (<b>B</b>) The secreted α-CTLA4 scFv in the cell culture supernatant was detected using flow cytometry-based binding assay and HEK-293T target cells overexpressing the CTLA4 ligand. The scFv-ligand binding was followed by staining target cells with A647-conjugated α-Myc antibody for detection of the bound scFv to target cells. (<b>C</b>) Flow cytometry data showing the induced expression and secretion of α-CTLA4 scFv upon the addition of 500 nM ABA or PBS and selective binding of the scFv to CTLA-4<sup>+</sup> cells. The shown histogram is a representative of 3 biological replicates.</p>
Full article ">Figure 8
<p>Synthesis and characterization of ABA prodrug (ABA-Gal) for <span class="html-italic">β</span>-galactosidase (<span class="html-italic">β</span>-gal). (<b>A</b>) Synthesis and the activation of ABA-Gal prodrug. Red arrow indicates cleavage of the prodrug by <span class="html-italic">β</span>-gal. (<b>B</b>) The stability of ABA-Gal and its generation of ABA by <span class="html-italic">β</span>-gal were analyzed using HPLC. Prodrug at 200 μM was incubated with <span class="html-italic">β</span>-gal at the indicated concentration in PBS buffer (pH 7.4) at 37 °C for different time points. (<b>C</b>) HEK-293T EGFP reporter cells responded to <span class="html-italic">β</span>-gal when treated with ABA-Gal, leading to EGFP expression. Reporter cells were treated with 20 μM ABA or 20 μM prodrug with and without 10 μM <span class="html-italic">β</span>-gal. Treated cells were incubated for 24 h, and EGFP expression was observed under a fluorescence microscope. The scale bar is 200 µm.</p>
Full article ">Figure 9
<p>TME-gated expression and secretion of customized biologics in engineered Jurkat T cells. (<b>A</b>) The secretion of IL-12 by CAR<sup>+</sup>iIL12 cells at 15 h post-drug treatment, as detected via ELISA. Data are shown as mean ± S.E.M from 3 technical triplicates. **** <span class="html-italic">p</span> ≤ 0.0001, ns, <span class="html-italic">p</span> ≥ 0.05, two-tailed Student <span class="html-italic">t</span>-test. (<b>B</b>) Competitive binding assay showing % PD1-positive Jurkat T cells in a 24 h co-culture of CAR<sup>+</sup>iPembro cells and HER2-expressing cancer cell line MDA-MB-231. Data are mean ± S.E.M from 2 biological replicates. * <span class="html-italic">p</span> &lt; 0.05, ns, <span class="html-italic">p</span> ≥ 0.05, one-way ANOVA with Dunnett multiple comparisons test. (<b>C</b>) The secretion of α-CTLA4 scFv by CAR<sup>+</sup>iCTLA4 scFv and the binding of scFv to CTLA4-expressing HEK-293T target cells, followed by staining target cells with A647-conjugated anti-Myc antibody and detected via flow cytometry. The histograms shown are representative of 2 biological replicates.</p>
Full article ">
28 pages, 1521 KiB  
Review
The Immunomodulatory Effects of Selenium: A Journey from the Environment to the Human Immune System
by Rebecka A. Sadler, Bonnie A. Mallard, Umesh K. Shandilya, Mohammed A. Hachemi and Niel A. Karrow
Nutrients 2024, 16(19), 3324; https://doi.org/10.3390/nu16193324 - 30 Sep 2024
Viewed by 955
Abstract
Selenium (Se) is an essential nutrient that has gained attention for its impact on the human immune system. The purpose of this review is to explore Se’s immunomodulatory properties and to make up-to-date information available so novel therapeutic applications may emerge. People acquire [...] Read more.
Selenium (Se) is an essential nutrient that has gained attention for its impact on the human immune system. The purpose of this review is to explore Se’s immunomodulatory properties and to make up-to-date information available so novel therapeutic applications may emerge. People acquire Se through dietary ingestion, supplementation, or nanoparticle applications. These forms of Se can beneficially modulate the immune system by enhancing antioxidant activity, optimizing the innate immune response, improving the adaptive immune response, and promoting healthy gut microbiota. Because of these many actions, Se supplementation can help prevent and treat pathogenic diseases, autoimmune diseases, and cancers. This review will discuss Se as a key micronutrient with versatile applications that supports disease management due to its beneficial immunomodulatory effects. Further research is warranted to determine safe dosing guidelines to avoid toxicity and refine the application of Se in medical treatments. Full article
(This article belongs to the Section Micronutrients and Human Health)
Show Figures

Figure 1

Figure 1
<p>Selenium sources and uses. Se: selenium; SeMeth: selenomethionine; SeCys: selenocysteine [<a href="#B4-nutrients-16-03324" class="html-bibr">4</a>,<a href="#B5-nutrients-16-03324" class="html-bibr">5</a>,<a href="#B6-nutrients-16-03324" class="html-bibr">6</a>].</p>
Full article ">Figure 2
<p>Absorption, distribution, and metabolism of selenium. Se: selenium; SeMeth: selenomethionine; SeCys: selenocysteine; SelP: selenoprotein P [<a href="#B21-nutrients-16-03324" class="html-bibr">21</a>,<a href="#B28-nutrients-16-03324" class="html-bibr">28</a>,<a href="#B29-nutrients-16-03324" class="html-bibr">29</a>].</p>
Full article ">Figure 3
<p>Antioxidant activity of selected selenoproteins [<a href="#B52-nutrients-16-03324" class="html-bibr">52</a>].</p>
Full article ">
21 pages, 4104 KiB  
Article
Optimization of Ultrasonic-Assisted Extraction, Characterization and Antioxidant and Immunoregulatory Activities of Arthrospira platensis Polysaccharides
by Na Wang, Jingyi Qin, Zishuo Chen, Jiayi Wu and Wenzhou Xiang
Molecules 2024, 29(19), 4645; https://doi.org/10.3390/molecules29194645 - 30 Sep 2024
Viewed by 752
Abstract
This study aimed to enhance the ultrasonic-assisted extraction (UAE) yield of seawater Arthrospira platensis polysaccharides (APPs) and investigate its structural characteristics and bioactivities. The optimization of UAE achieved a maximum crude polysaccharides yield of 14.78%. The optimal extraction conditions were a liquid–solid ratio [...] Read more.
This study aimed to enhance the ultrasonic-assisted extraction (UAE) yield of seawater Arthrospira platensis polysaccharides (APPs) and investigate its structural characteristics and bioactivities. The optimization of UAE achieved a maximum crude polysaccharides yield of 14.78%. The optimal extraction conditions were a liquid–solid ratio of 30.00 mL/g, extraction temperature of 81 °C, ultrasonic power at 92 W and extraction time at 30 min. After purification through cellulose DEAE-52 and Sephadex G-100 columns, two polysaccharide elutions (APP-1 and APP-2) were obtained. APP-2 had stronger antioxidant and immunoregulatory activities than APP-1, thus the characterization of APP-2 was conducted. APP-2 was an acidic polysaccharide consisting of rhamnose, glucose, mannose and glucuronic acid at a ratio of 1.00:24.21:7.63:1.53. It possessed a molecular weight of 72.48 kDa. Additionally, APP-2 had linear and irregular spherical particles and amorphous structures, which contained pyranoid polysaccharides with alpha/beta glycosidic bonds. These findings offered the foundation for APP-2 as an antioxidant and immunomodulator applied in the food, pharmaceutical and cosmetic industries. Full article
(This article belongs to the Special Issue Natural Products from Plant: From Determination to Application)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effects of liquid–solid ratio (<b>a</b>), extraction temperature (<b>b</b>), ultrasonic power (<b>c</b>) and extraction time (<b>d</b>) on the extraction yield of <span class="html-italic">Arthrospira platensis</span> polysaccharides (n = 3).</p>
Full article ">Figure 2
<p>3D response surface plots (<b>a</b>–<b>f</b>) of the interaction effects of variables (A: liquid–solid ratio, mL/g; B: extraction temperature, °C; C: ultrasonic power, W; and D: extraction time, min) on <span class="html-italic">Arthrospira platensis</span> polysaccharide yield.</p>
Full article ">Figure 3
<p>Contour plots (<b>a</b>–<b>f</b>) showing the effect of variables (A: liquid–solid ratio, mL/g; B: extraction temperature, °C; C: ultrasonic power, W; and D: extraction time, min) on the extraction yield of <span class="html-italic">Arthrospira platensis</span> polysaccharides. The number of contour lines indicates the elevation of the location. A denser arrangement of contour lines signifies significant changes in the terrain, whereas a sparser arrangement suggests relatively flat topography.</p>
Full article ">Figure 4
<p>DEAE-cellulose (<b>a</b>) and Sephadex G-100 elution profiles (<b>b</b>,<b>c</b>) of polysaccharides from <span class="html-italic">Arthrospira platensis</span>.</p>
Full article ">Figure 5
<p>Antioxidant activities of APP-1 and APP-2. Scavenging effects against DPPH (<b>a</b>), ABTS (<b>b</b>) and Hydroxyl (<b>c</b>) free radicals, and Ferrous chelation rate (<b>d</b>).</p>
Full article ">Figure 6
<p>Effects of APP-1 and APP-2 on the viability (<b>a</b>), phagocytosis activity (<b>b</b>), secretions of NO (<b>c</b>), TNF-α (<b>d</b>), IL-6 (<b>e</b>) and IL-1β (<b>f</b>) of RAW 264.7 cells. * <span class="html-italic">p</span> &lt; 0.05 vs. cell blank control, ** <span class="html-italic">p</span> &lt; 0.01 vs. cell blank control.</p>
Full article ">Figure 7
<p>UV spectra (<b>a</b>), FTIR spectra (<b>b</b>) and XRD pattern (<b>c</b>) of APP-2.</p>
Full article ">Figure 8
<p>SEM images of APP-2 at 2000 (<b>a</b>) and 10,000 (<b>b</b>) magnifications. AFM images of APP-2 (<b>c</b>–<b>e</b>).</p>
Full article ">
15 pages, 3734 KiB  
Article
Effect of Dietary Astragalus polysaccharides (APS) on the Growth Performance, Antioxidant Responses, Immunological Parameters, and Intestinal Microbiota of Coral Trout (Plectropomus leopardus)
by Xiaoqi Hao, Heizhao Lin, Ziyang Lin, Keng Yang, Jing Hu, Zhenhua Ma and Wei Yu
Microorganisms 2024, 12(10), 1980; https://doi.org/10.3390/microorganisms12101980 - 30 Sep 2024
Viewed by 484
Abstract
The potential effects of Astragalus polysaccharides (APS) were evaluated in coral trout (Plectropomus leopardus). Five APS levels (0%, 0.05%, 0.10%, 0.15%, and 0.20%) were added to the diet of coral trout, and a 56-day growth trial (initial weight 18.62 ± 0.05 [...] Read more.
The potential effects of Astragalus polysaccharides (APS) were evaluated in coral trout (Plectropomus leopardus). Five APS levels (0%, 0.05%, 0.10%, 0.15%, and 0.20%) were added to the diet of coral trout, and a 56-day growth trial (initial weight 18.62 ± 0.05 g) was conducted. Dietary APS enhanced growth performance, with the highest improvement observed in fish fed the 0.15% APS diet. This concentration also enhanced the antioxidant capacity and immunomodulation of the fish by regulating the expression of genes associated with antioxidant enzymes and immune responses. Intestinal microbiota analysis revealed that APS supplementation significantly increased the Chao1 index and relative abundance of beneficial bacteria (Firmicutes and Bacillus). A high level of APS (0.20%) did not provide additional benefits for growth and health compared to a moderate level (0.15%). These findings indicate that an optimal APS dose promotes growth, enhances antioxidant activity, supports immune function, and improves intestinal microbiota in coral trout. Based on a cubic regression analysis of the specific growth rate, the optimal APS level for the maximal growth of coral trout was determined to be 0.1455%. Full article
(This article belongs to the Special Issue Aquatic Microorganisms and Their Application in Aquaculture)
Show Figures

Figure 1

Figure 1
<p>Prediction of the <span class="html-italic">Astragalus polysaccharides</span> (APS) dose required for maximal SGR of coral trout (<span class="html-italic">Plectropomus leopardus</span>) using cubic regression analysis. Values are shown as the mean ± SE (<span class="html-italic">n</span> = 3). Different letters indicate significant differences between the treatments.</p>
Full article ">Figure 2
<p>Intestinal morphology in coral trout (<span class="html-italic">Plectropomus leopardus</span>) fed different <span class="html-italic">Astragalus polysaccharides</span> (APS) dietary levels (long arrow, villus length; short arrow, muscle thickness). (<b>A</b>) 0% APS; (<b>B</b>) 0.05% APS; (<b>C</b>) 0.10% APS; (<b>D</b>) 0.15% APS; (<b>E</b>) 0.20% APS; (<b>F</b>) villus length; (<b>G</b>) muscle thickness. Scale bar, 200 μm. Values are shown as the mean ± SE (<span class="html-italic">n</span> = 3). Different letters indicate a significant difference.</p>
Full article ">Figure 3
<p>Expression of liver antioxidant enzyme (<b>A</b>) and immune indicator (<b>B</b>) genes in coral trout (<span class="html-italic">Plectropomus leopardus</span>) fed different <span class="html-italic">Astragalus polysaccharides</span> (APS) dietary levels. Values are shown as the mean ± SE (<span class="html-italic">n</span> = 3). The full names of the PCR target genes (<span class="html-italic">SOD-1</span>, <span class="html-italic">SOD-2</span>, <span class="html-italic">CAT</span>, <span class="html-italic">GSH-Px1a</span>, <span class="html-italic">ACP6</span>, <span class="html-italic">AKP</span>, <span class="html-italic">C3</span>, <span class="html-italic">C4-b</span>, <span class="html-italic">IgM</span>, <span class="html-italic">LZ-c</span>, and <span class="html-italic">GAPDH</span>) are provided in <a href="#app1-microorganisms-12-01980" class="html-app">Table S2</a>. Different letters indicate a significant difference.</p>
Full article ">Figure 4
<p>Alpha-diversity and beta-diversity of the gut bacterial community in coral trout (<span class="html-italic">Plectropomus leopardus</span>) fed different <span class="html-italic">Astragalus polysaccharides</span> (APS) dietary levels. (<b>A</b>) Chao1, Shannon, and Simpson index. Values are shown as the mean ± SE (<span class="html-italic">n</span> = 3) via the student’s two-tailed <span class="html-italic">t</span>-test. Different symbols (*, <span class="html-italic">p</span> &lt; 0.05; ns, not significant) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Rarefaction curve of the intestines in coral trout fed different APS dietary levels. (<b>C</b>) PCoA (Principal Co-ordinates Analysis) of the intestines in coral trout fed different APS dietary levels based on weighted UniFrac distances. Control, 0% APS group; APS, 0.15% APS group.</p>
Full article ">Figure 5
<p>Bacterial compositions in coral trout (<span class="html-italic">Plectropomus leopardus</span>) fed different <span class="html-italic">Astragalus polysaccharides</span> (APS) dietary levels. (<b>A</b>,<b>B</b>) Distribution of the top 10 microbial phylum levels in different groups, with the legend shown on the right. (<b>C</b>,<b>D</b>) Distribution of the top 10 microbial genus levels in different groups, with the legend shown on the right. Control, 0% APS group; APS, 0.15% APS group.</p>
Full article ">
11 pages, 839 KiB  
Review
Halocins and C50 Carotenoids from Haloarchaea: Potential Natural Tools against Cancer
by Rosa María Martínez-Espinosa
Mar. Drugs 2024, 22(10), 448; https://doi.org/10.3390/md22100448 - 29 Sep 2024
Viewed by 780
Abstract
Haloarchaea are a group of moderate and extreme halophilic microorganisms, belonging to the Archaea domain, that constitute relevant microbial communities in salty environments like coastal and inland salted ponds, marshes, salty lagoons, etc. They can survive in stress conditions such as high salinity [...] Read more.
Haloarchaea are a group of moderate and extreme halophilic microorganisms, belonging to the Archaea domain, that constitute relevant microbial communities in salty environments like coastal and inland salted ponds, marshes, salty lagoons, etc. They can survive in stress conditions such as high salinity and, therefore, high ionic strength, high doses of ultraviolet radiation (UV), high temperature, and extreme pH values. Consequently, most of the species can be considered polyextremophiles owing to their ability to respond to the multiple extreme conditions characterizing their natural habitats. They cope with those stresses thanks to several molecular and metabolic adaptations. Thus, some of the molecules produced by haloarchaea show significantly different biological activities and physicochemical properties compared to their bacterial counterparts. Recent studies have revealed promising applications in biotechnology and medicine for these biomolecules. Among haloarchaeal biomolecules, rare natural pigments (C50 carotenoids) and small peptides called halocins and microhalocins have attracted attention worldwide due to their effects on animal and human commercial tumoral cells, apart from the role as antibiotics described for halocins or the immunomodulatory activity reported from C50 carotenoids like bacterioruberin. This review summarizes recent knowledge on these two types of biomolecules in connection with cancer to shed new light on the design of drugs and new therapies based on natural compounds. Full article
(This article belongs to the Special Issue Discovery of Marine-Derived Anticancer Agents)
Back to TopTop