[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,173)

Search Parameters:
Keywords = cancer immunotherapies

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 34340 KiB  
Article
Establishment and Verification of a Novel Gene Signature Connecting Hypoxia and Lactylation for Predicting Prognosis and Immunotherapy of Pancreatic Ductal Adenocarcinoma Patients by Integrating Multi-Machine Learning and Single-Cell Analysis
by Ying Zheng, Yang Yang, Qunli Xiong, Yifei Ma and Qing Zhu
Int. J. Mol. Sci. 2024, 25(20), 11143; https://doi.org/10.3390/ijms252011143 - 17 Oct 2024
Abstract
Pancreatic ductal adenocarcinoma (PDAC) has earned a notorious reputation as one of the most formidable and deadliest malignant tumors. Within the tumor microenvironment, cancer cells have acquired the capability to maintain incessant expansion and increased proliferation in response to hypoxia via metabolic reconfiguration, [...] Read more.
Pancreatic ductal adenocarcinoma (PDAC) has earned a notorious reputation as one of the most formidable and deadliest malignant tumors. Within the tumor microenvironment, cancer cells have acquired the capability to maintain incessant expansion and increased proliferation in response to hypoxia via metabolic reconfiguration, leading to elevated levels of lactate within the tumor surroundings. However, there have been limited studies specifically investigating the association between hypoxia and lactic acid metabolism-related lactylation in PDAC. In this study, multiple machine learning approaches, including LASSO regression analysis, XGBoost, and Random Forest, were employed to identify hub genes and construct a prognostic risk signature. The implementation of the CERES score and single-cell analysis was used to discern a prospective therapeutic target for the management of PDAC. CCK8 assay, colony formation assays, transwell, and wound-healing assays were used to explore both the proliferation and migration of PDAC cells affected by CENPA. In conclusion, we discovered two distinct subtypes characterized by their unique hypoxia and lactylation profiles and developed a risk score to evaluate prognosis, as well as response to immunotherapy and chemotherapy, in PDAC patients. Furthermore, we indicated that CENPA may serve as a promising therapeutic target for PDAC. Full article
(This article belongs to the Section Molecular Immunology)
Show Figures

Figure 1

Figure 1
<p>Identification of prognostic hypoxia- and lactylation-related genes (HALRGs) and mutation landscape. (<b>A</b>) Intersection of differentially expressed genes (DEGs) in PDAC samples with hypoxia- and lactylation-related genes. (<b>B</b>) Univariate Cox analysis of these genes. (<b>C</b>) Biological network integration of these prognostic genes analyzed by GeneMANIA. (<b>D</b>) Kaplan–Meier survival curve of certain prognostic genes.</p>
Full article ">Figure 2
<p>Pan-cancer analysis of the prognostic hypoxia- and lactylation-related genes. (<b>A</b>) Survival differences between high and low GSVA score groups across various cancers. (<b>B</b>) Association between GSVA scores and cancer-related pathway activity (*: <span class="html-italic">p</span>-value ≤ 0.05; #: FDR ≤ 0.05). (<b>C</b>,<b>D</b>) Summary of the relationship between gene expression and responsiveness of top 30 GDSC and CTRP drugs in the pan-cancer analysis.</p>
Full article ">Figure 3
<p>Unsupervised clustering analysis identified two PDAC subtypes with distinctive biological functional characteristics in the TCGA and GSE183795 cohorts. (<b>A</b>) Consensus matrix heatmap defining two subtypes (k = 2). (<b>B</b>) PCA indicating the significant differences in transcriptomes between the subtypes. (<b>C</b>) Survival analysis indicates cluster A has a poor prognosis compared to cluster B. (<b>D</b>) Using PROGENy (Pathway RespOnsive GENes for activity inference) to assess the pathway activation in the above two subtypes (ns <span class="html-italic">p</span> &gt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; **** <span class="html-italic">p</span> &lt; 0.0001). (<b>E</b>) KEGG enrichment analysis of the two subtypes. (<b>F</b>) GO enrichment analysis of the two subtypes.</p>
Full article ">Figure 4
<p>Identification of hub genes using various machine learning algorithms, and construction of a hypoxia- and lactylation-related prognostic signature for PDAC. (<b>A</b>,<b>B</b>) LASSO Cox regression was used to identify signature genes and develop a prognostic module for PDAC patients. (<b>C</b>) Bar graph of the coefficient index of the hub genes. (<b>D</b>) Heatmap of hub gene expression in the low- and high-risk groups. (<b>E</b>,<b>F</b>) Risk score distribution and survival status in the two risk groups. (<b>G</b>) Kaplan–Meier survival curve showing overall survival (OS) in the two risk groups. (<b>H</b>) ROC curves predicting the sensitivity and specificity of the risk score model for the 1-, 3-, and 5-year survival rates. (<b>I</b>) Time-dependent ROC analysis indicating the predictive power of the risk signature and other clinical characteristics. (<b>J</b>,<b>K</b>) Mutation landscape of the low- and high-risk groups.</p>
Full article ">Figure 5
<p>Validation of the prognostic module in independent external datasets (GSE62452, GSE78299, and GSE85916) and nomogram construction. (<b>A</b>–<b>C</b>) Kaplan–Meier analysis validating the predictive power of the prognostic model in the GSE62452, GSE78299, and GSE85916 datasets. (<b>D</b>–<b>F</b>) ROC curves demonstrating the sensitivity and specificity of the risk score model for the 1-, 3-, and 5-year survival rates in these test cohorts. (<b>G</b>,<b>H</b>) Nomogram construction integrating the risk score and clinical characteristics (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01). (<b>I</b>,<b>J</b>) Forest plots of the univariate and multivariate Cox regression analyses show that the risk score is an independent prognostic factor for PDAC in the training cohort.</p>
Full article ">Figure 6
<p>The immunogenomic landscape of signature genes and their predictive values for immunotherapy and chemotherapy. (<b>A</b>) Correlation between risk scores and immune cell abundance analyzed using various immune cell profiling methods. (<b>B</b>) Evaluation of the potential efficacy of immunotherapy in low- and high-risk groups, showing a less favorable response in the high-risk group.(*** <span class="html-italic">p</span> &lt; 0.001) (<b>C</b>) Correlation analysis between signature genes and genes associated with immune evasion. (<b>D</b>) Analysis of chemotherapeutic sensitivity between the low- and high-risk groups (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 7
<p>CERES score of signature genes and HAL score analysis at the single-cell level. (<b>A</b>) CERES score of signature genes. (<b>B</b>) UMAP-1 plot showing cell subtypes identified from scRNA-seq data. (<b>C</b>) Distribution of <span class="html-italic">CENPA</span> in metastasis, normal, and primary PADC scRNA samples. (<b>D</b>) Heatmap displaying variations in interaction numbers. (<b>E</b>) Bar graph showing key signaling pathways differing between the high- and low-scoring groups. (<b>F</b>) Circular plot visualizing differences in cell–cell communication networks between the high- and low-scoring groups.</p>
Full article ">Figure 8
<p>The expression profile of <span class="html-italic">CENPA</span> in PDAC; the knockdown of <span class="html-italic">CENPA</span> hampers the proliferation and migratory potential of PDAC cells. (<b>A</b>) Validation of <span class="html-italic">CENPA</span> expression in the HPA database. (<b>B</b>) The expression level of <span class="html-italic">CENPA</span> in the PDAC expression data cohort from the TCGA and GETx database. (<b>C</b>) Associations between <span class="html-italic">CENPA</span> expression and overall survival of PDAC patients. (<b>D</b>) Relative mRNA expression of <span class="html-italic">CENPA</span> in PDAC cell lines (BXPC-3, CAPAN-1, CAPAN-2, CFPAC-1, MIA PaCa-2, PANC-1, and SW1990) and HPDE normal pancreatic ductal epithelial cells. (<b>E</b>) <span class="html-italic">CENPA</span> knockdown in PANC-1 and MIA PaCa-2 cells verified by qRT-PCR and Western blot. The cck8 assay (<b>F</b>) and colony formation assay (<b>G</b>) show reduced cell viability in <span class="html-italic">CENPA</span> knockdown PANC-1 and MIA PaCa-2 cells. (<b>H</b>,<b>I</b>) Wound-healing and transwell assays indicate significantly reduced migration ability in <span class="html-italic">CENPA</span> knockdown PANC-1 and MIA PaCa-2 cells. <span class="html-italic">n</span> = 3, ns <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Error bars represent mean ± SD.</p>
Full article ">Figure 9
<p>Correlation between <span class="html-italic">CENPA</span> expression and drug sensitivity, and molecular docking of drugs correlated with the high expression of <span class="html-italic">CENPA</span>. (<b>A</b>) Correlation analysis between <span class="html-italic">CENPA</span> expression and drug sensitivity, conducted using BEST. (<b>B</b>) Molecular docking diagrams of <span class="html-italic">CENPA</span> with the two drugs showing the strongest binding affinity: betulinic acid (−8.1 kcal/mol) and GSK2126458 (−8.6 kcal/mol).</p>
Full article ">
18 pages, 2127 KiB  
Review
Oncolytic Viruses as Reliable Adjuvants in CAR-T Cell Therapy for Solid Tumors
by Ruxandra Ilinca Stilpeanu, Bianca Stefania Secara, Mircea Cretu-Stancu and Octavian Bucur
Int. J. Mol. Sci. 2024, 25(20), 11127; https://doi.org/10.3390/ijms252011127 - 16 Oct 2024
Viewed by 322
Abstract
Although impactful scientific advancements have recently been made in cancer therapy, there remains an opportunity for future improvements. Immunotherapy is perhaps one of the most cutting-edge categories of therapies demonstrating potential in the clinical setting. Genetically engineered T cells express chimeric antigen receptors [...] Read more.
Although impactful scientific advancements have recently been made in cancer therapy, there remains an opportunity for future improvements. Immunotherapy is perhaps one of the most cutting-edge categories of therapies demonstrating potential in the clinical setting. Genetically engineered T cells express chimeric antigen receptors (CARs), which can detect signals expressed by the molecules present on the surface of cancer cells, also called tumor-associated antigens (TAAs). Their effectiveness has been extensively demonstrated in hematological cancers; therefore, these results can establish the groundwork for their applications on a wide range of requirements. However, the application of CAR-T cell technology for solid tumors has several challenges, such as the existence of an immune-suppressing tumor microenvironment and/or inadequate tumor infiltration. Consequently, combining therapies such as CAR-T cell technology with other approaches has been proposed. The effectiveness of combining CAR-T cell with oncolytic virus therapy, with either genetically altered or naturally occurring viruses, to target tumor cells is currently under investigation, with several clinical trials being conducted. This narrative review summarizes the current advancements, opportunities, benefits, and limitations in using each therapy alone and their combination. The use of oncolytic viruses offers an opportunity to address the existing challenges of CAR-T cell therapy, which appear in the process of trying to overcome solid tumors, through the combination of their strengths. Additionally, utilizing oncolytic viruses allows researchers to modify the virus, thus enabling the targeted delivery of specific therapeutic agents within the tumor environment. This, in turn, can potentially enhance the cytotoxic effect and therapeutic potential of CAR-T cell technology on solid malignancies, with impactful results in the clinical setting. Full article
Show Figures

Figure 1

Figure 1
<p>The structure of a CAR-T Cell. (<b>a</b>) The T cell, the chimeric antigen receptor (CAR), and the genetically engineered CAR-T cell are illustrated. (<b>b</b>) The structure of the CAR-T cell and its interaction with the surface antigen of the tumor cell are presented. (<b>c</b>) Differences between the intracellular signaling domains of the 5 generations of CARs. The first generation presents a CD3ζ-derived signaling module. The second generation of CARs is the first to contain a co-stimulatory domain. Co-stimulatory molecules include CD28, 4-1BB (CD137), CD27, and OX40 (CD134). TM—transmembrane domain; TRUCK—T cell redirected for universal cytokine-mediated killing; CoS1,2—co-stimulatory domain; IL-2Rβ—IL-2 receptor β. This figure was created with GoodNotes and adapted from reference [<a href="#B13-ijms-25-11127" class="html-bibr">13</a>] for (<b>a</b>) and (<b>b</b>) and reference [<a href="#B14-ijms-25-11127" class="html-bibr">14</a>] for (<b>c</b>).</p>
Full article ">Figure 2
<p>Three possible combinations of CAR-T cell therapy with other therapies are presented. From right to left: (<b>a</b>) chemotherapy, (<b>b</b>) oncolytic viral therapy, and (<b>c</b>) radiation therapy. This figure was created with GoodNotes and adapted from reference [<a href="#B2-ijms-25-11127" class="html-bibr">2</a>].</p>
Full article ">Figure 3
<p>The effect of CAR-T cells on macrophages and the development of cytokine release syndrome. The interaction of macrophages with activated CAR-T cells further leads to the release of chemokines and cytokines such as IL-1, IL-6, and iNOS, causing supplementary inflammatory reactions, resulting in CRS. This figure was created with GoodNotes and adapted from reference [<a href="#B117-ijms-25-11127" class="html-bibr">117</a>].</p>
Full article ">
20 pages, 526 KiB  
Review
The Role of Resveratrol in Cancer Management: From Monotherapy to Combination Regimens
by Eduarda Ribeiro and Nuno Vale
Targets 2024, 2(4), 307-326; https://doi.org/10.3390/targets2040018 (registering DOI) - 16 Oct 2024
Viewed by 288
Abstract
Resveratrol, a naturally occurring polyphenol found in grapes, berries, and peanuts, has garnered significant attention for its potential anti-cancer properties. This review provides a comprehensive analysis of its role in cancer therapy, both as a standalone treatment and in combination with other therapeutic [...] Read more.
Resveratrol, a naturally occurring polyphenol found in grapes, berries, and peanuts, has garnered significant attention for its potential anti-cancer properties. This review provides a comprehensive analysis of its role in cancer therapy, both as a standalone treatment and in combination with other therapeutic approaches. This review explores the molecular mechanisms underlying resveratrol’s anti-cancer effects, including its antioxidant activity, modulation of cellular signaling pathways, antiproliferative properties, anti-inflammatory effects, and epigenetic influences. This review also examines in vitro and in vivo studies that highlight resveratrol’s efficacy against various cancer types. Furthermore, the synergistic effects of resveratrol when used in conjunction with conventional treatments like chemotherapy and radiotherapy, as well as targeted therapies and immunotherapies, are discussed. Despite promising preclinical results, this review addresses the challenges and limitations faced in translating these findings into clinical practice, including issues of bioavailability and toxicity. Finally, it outlines future research directions and the potential for resveratrol to enhance existing cancer treatment regimens. This review aims to provide a thorough understanding of resveratrol’s therapeutic potential and to identify areas for further investigation in the quest for effective cancer treatments. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of resveratrol (3,5,4′-trihydroxylstilbene).</p>
Full article ">
12 pages, 954 KiB  
Perspective
Hepatocellular Carcinoma Immunotherapy: Predictors of Response, Issues, and Challenges
by Alessandro Rizzo, Oronzo Brunetti and Giovanni Brandi
Int. J. Mol. Sci. 2024, 25(20), 11091; https://doi.org/10.3390/ijms252011091 - 15 Oct 2024
Viewed by 248
Abstract
Immune checkpoint inhibitors (ICIs), such as durvalumab, tremelimumab, and atezolizumab, have emerged as a significant therapeutic option for the treatment of hepatocellular carcinoma (HCC). In fact, the efficacy of ICIs as single agents or as part of combination therapies has been demonstrated in [...] Read more.
Immune checkpoint inhibitors (ICIs), such as durvalumab, tremelimumab, and atezolizumab, have emerged as a significant therapeutic option for the treatment of hepatocellular carcinoma (HCC). In fact, the efficacy of ICIs as single agents or as part of combination therapies has been demonstrated in practice-changing phase III clinical trials. However, ICIs confront several difficulties, including the lack of predictive biomarkers, primary and secondary drug resistance, and treatment-related side effects. Herein, we provide an overview of current issues and future challenges in this setting. Full article
Show Figures

Figure 1

Figure 1
<p>Pro- and anti-tumor cells in the tumor microenvironment.</p>
Full article ">Figure 2
<p>Immunotherapies in hepatocellular carcinoma patients.</p>
Full article ">
19 pages, 29193 KiB  
Article
MSI-H Detection by ddPCR in Endoscopic Ultrasound Fine Needle Biopsy (EUS-FNB) from Pancreatic Ductal Adenocarcinoma
by Maria Assunta Piano, Elisa Boldrin, Lidia Moserle, Nicoletta Salerno, Dalila Fanelli, Giulia Peserico, Maria Raffaella Biasin, Giovanna Magni, Veronica Varano, Giorgia Zalgelli, Vasileios Mourmouras, Antonio Rosato, Antonio Scapinello, Alberto Fantin and Matteo Curtarello
Int. J. Mol. Sci. 2024, 25(20), 11090; https://doi.org/10.3390/ijms252011090 (registering DOI) - 15 Oct 2024
Viewed by 543
Abstract
Pancreatic ductal adenocarcinoma (PDAC) is a highly aggressive disease with limited survival. Curative opportunities are only available for patients with resectable cancer. Palliative chemotherapy is the current standard of care for unresectable tumors. Numerous efforts have been made to investigate new therapeutic strategies [...] Read more.
Pancreatic ductal adenocarcinoma (PDAC) is a highly aggressive disease with limited survival. Curative opportunities are only available for patients with resectable cancer. Palliative chemotherapy is the current standard of care for unresectable tumors. Numerous efforts have been made to investigate new therapeutic strategies for PDAC. Immunotherapy has been found to be effective in treating tumors with high microsatellite instability (MSI-H), including PDAC. The ability of the Endoscopic Ultrasound Fine Needle Biopsy (EUS-FNB) to reliably collect tissue could enhance new personalized treatment by permitting genomic alterations analysis. The aim of this study was to investigate the feasibility of obtaining adequate DNA for molecular analysis from EUS-FNB formalin-fixed-paraffin-embedded (FFPE) specimens. For this purpose, FFPE-DNA obtained from 43 PDAC archival samples was evaluated to verify adequacy in terms of quantity and quality and was tested to evaluate MSI-H status by droplet digital PCR (ddPCR). All samples were suitable for ddPCR analysis. Unlike the 1–2% MSI-H frequency found with traditional techniques, ddPCR detected this phenotype in 16.28% of cases. This study suggests the ddPCR ability to identify MSI-H phenotype, with the possibility of improving the selection of patients who may benefit from immunotherapy and who would be excluded by performing traditional diagnostic methods. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Pie chart showing the percentages of formalin-fixed-paraffin-embedded (FFPE) macrodissected samples (16.3%, orange slice) and not macrodissected (83.7%, yellow/white slice). No data were available for 3 samples (white slice). (<b>B</b>) Schematic representation of the amount of tumor DNA used as input in droplet digital PCR (ddPCR) analyses. In orange FFPE macrodissected samples, in yellow is represented the range from 3.64 ng to 0.27 ng of tumor DNA used as input in FFPE not macrodissected samples.</p>
Full article ">Figure 2
<p>(<b>A</b>) Immunohistochemistry (IHC) showing normal expression of the four mismatch repair (MMR) (MLH1, MSH2, MSH6, PMS2) proteins (original magnification 60×) and (<b>B</b>) the two-dimensional plots of the five microsatellites marker loci (BAT-25 and BAT-26; NR-21 and NR-24 and MONO-27) analyzed by ddPCR showing instability in three loci (BAT-25, BAT-26 and NR-21) in a representative pancreatic ductal adenocarcinoma (PDAC) FFPE sample. Positive control (CTRLpos) has been used to recognize the exact position of the droplet cluster to call the microsatellite as positive. Orange droplets (orange circle) represent microsatellites with unaltered length, blue droplets (blue circle) represent the microsatellite unstable molecules, and grey droplets (grey circle) represent the ones with the no DNA template.</p>
Full article ">Figure 3
<p>IHC of MMR protein expression and the corresponding hematoxylin/eosin (HE) stain of representative Endoscopic Ultrasound Fine Needle Biopsy (EUS-FNB) specimens of 5 out of the 7 cases resulted in MSI-H according to ddPCR. In all cases, the IHC shows retained nuclear staining of all the four MMR proteins and hence the cases were defined as microsatellite stable (MSS). The HE staining shows a typical PDAC histomorphology (original magnification, 40×).</p>
Full article ">Figure A1
<p>A representative image of Immunohistochemistry (IHC) staining for the four mismatch repair (MMR) proteins (MLH1, MSH2, MSH6 and PMS2) and the corresponding 2D plots for the droplet digital PCR (ddPCR) microsatellite instability (MSI) assays showing the status of the 5 microsatellite marker loci (BAT-25 and BAT-26; NR-21 and NR-24 and MONO-27) of an MSI-H gastric cancer sample, and an microsatellite stable (MSS) pancreatic ductal adenocarcinoma (PDAC) sample. In the 2D plots, black circles identify the clusters of blue droplets, which correspond to microsatellites with altered lengths (instability). If ≥3 blue droplets were included in the black circles, the microsatellite locus was considered unstable. If at least ≥2 loci were unstable, the sample was considered MSI-H, otherwise the sample was considered MSS. Orange droplets represent microsatellites with unaltered lengths, grey droplets represent the no DNA template ones.</p>
Full article ">Figure A2
<p>PDAC patient’s overall survival (OS) across the two different groups MSI-H and MSS according to ddPCR analysis.</p>
Full article ">Figure A3
<p>Quality assessment of formalin-fixed-paraffin-embedded (FFPE)-DNA from Endoscopic Ultrasound Fine Needle Biopsy (EUS-FNB) using the Agilent TapeStation 4200 (Agilent Genomic DNA ScreenTape Assay). In the upper, electrophoretic runs of fifteen representative samples, and in the bottom, electropherograms of 2 out of 15 representative samples (sample#1 and sample#13). LD: Ladder, DIN: DNA Integrity Number.</p>
Full article ">
26 pages, 409 KiB  
Review
Advancing Treatment Options for Merkel Cell Carcinoma: A Review of Tumor-Targeted Therapies
by Helena M. Nammour, Karla Madrigal, Caroline T. Starling and Hung Q. Doan
Int. J. Mol. Sci. 2024, 25(20), 11055; https://doi.org/10.3390/ijms252011055 (registering DOI) - 15 Oct 2024
Viewed by 466
Abstract
Although rare, Merkel cell carcinoma (MCC) is a highly aggressive and increasingly prevalent neuroendocrine cancer of the skin. While current interventions, including surgical resection, radiation, and immunotherapy have been employed in treating many patients, those who remain unresponsive to treatment are met with [...] Read more.
Although rare, Merkel cell carcinoma (MCC) is a highly aggressive and increasingly prevalent neuroendocrine cancer of the skin. While current interventions, including surgical resection, radiation, and immunotherapy have been employed in treating many patients, those who remain unresponsive to treatment are met with sparse alternatives and a grim prognosis. For this reason, it is of interest to expand the repertoire of available therapies for MCC patients who remain resistant to current primary interventions. Recently, our improved mechanistic understanding of aberrant cell signaling observed in both MCPyV-positive and -negative MCC has facilitated exploration into several small molecules and inhibitors, among them receptor tyrosine kinase inhibitors (TKIs) and somatostatin analogs (SSAs), both of which have positively improved response rates and reduced tumor volumes upon application to treatment of MCC. The introduction of such targeted therapies into treatment protocols holds promise for more personalized care tailored towards patients of diverse subtypes, thereby improving outcomes and mitigating tumor burden, especially for treatment-resistant individuals. In this review, we characterize recent findings surrounding targeted treatments that have been applied to MCC and provide an overview of emerging perspectives on translatable options that can be further developed to offer additional therapeutic avenues for patients with the disease. 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 383
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 ">
18 pages, 318 KiB  
Review
Review of Current and Future Medical Treatments in Head and Neck Squamous Cell Carcinoma
by Aaron M. Lee, Alice N. Weaver, Phillip Acosta, Lauren Harris and Daniel W. Bowles
Cancers 2024, 16(20), 3488; https://doi.org/10.3390/cancers16203488 (registering DOI) - 15 Oct 2024
Viewed by 354
Abstract
Head and neck squamous cell carcinoma (HNSCC) is a complex cancer requiring a multidisciplinary approach. For patients with locally or regionally advanced disease, surgery and/or radiation are the cornerstones of definitive treatment. Medical therapy plays an important adjunct role in this setting, typically [...] Read more.
Head and neck squamous cell carcinoma (HNSCC) is a complex cancer requiring a multidisciplinary approach. For patients with locally or regionally advanced disease, surgery and/or radiation are the cornerstones of definitive treatment. Medical therapy plays an important adjunct role in this setting, typically consisting of a platinum-based regimen given as induction, concurrent, or adjuvant treatment. While relapsed/metastatic HNSCC has historically been a difficult-to-treat disease with poor outcomes, options have considerably improved with the incorporation of biologics and immune checkpoint inhibitors. Clinical trials are ongoing to investigate novel approaches, including new and combination immunotherapies, targeted therapies, therapeutic vaccines, antibody–drug conjugates, and cellular therapies. The results thus far have been mixed, highlighting the knowledge gaps that continue to challenge the medical oncologist treating HNSCC. Here, we present the most updated and broad review of the current treatment landscape in both locoregional and metastatic HNSCC and discuss the expansive future medical therapies under investigation. Full article
11 pages, 397 KiB  
Review
Neoadjuvant Immunotherapy in Head and Neck Cancers: A Paradigm Shift in Treatment Approach
by Alessia Zotta, Maria Luisa Marciano, Francesco Sabbatino, Alessandro Ottaiano, Marco Cascella, Monica Pontone, Massimo Montano, Ester Calogero, Francesco Longo, Morena Fasano, Teresa Troiani, Fortunato Ciardiello, Fabiana Raffaella Rampetta, Giovanni Salzano, Giovanni Dell’Aversana Orabona, Luigi Califano, Franco Ionna and Francesco Perri
Biomedicines 2024, 12(10), 2337; https://doi.org/10.3390/biomedicines12102337 - 14 Oct 2024
Viewed by 336
Abstract
Checkpoint inhibitors (ICIs) have demonstrated substantial efficacy in the treatment of numerous solid tumors, including head and neck cancer. Their inclusion in the therapeutic paradigm in metastatic lines of treatment has certainly improved the outcomes of these patients. Starting from this assumption, numerous [...] Read more.
Checkpoint inhibitors (ICIs) have demonstrated substantial efficacy in the treatment of numerous solid tumors, including head and neck cancer. Their inclusion in the therapeutic paradigm in metastatic lines of treatment has certainly improved the outcomes of these patients. Starting from this assumption, numerous studies have been conducted on ICIs in other earlier disease settings, including studies conducted in patients in neoadjuvant settings. However, how many and which studies are truly significant? Can they lay concrete foundations for further future studies and therefore allow us to continue to have this interesting future perspective? Through a review of the existing literature, coupled with insights gleaned from clinical practice and from the main recently published studies, we aim to examine the therapeutic potential of ICIs in patients affected by head and neck cancer in a neoadjuvant treatment setting and encourage researchers to set up successful future clinical trials. Full article
(This article belongs to the Special Issue Head and Neck Tumors, 3rd Edition)
13 pages, 1536 KiB  
Review
Cancer Immunotherapy: Targeting TREX1 Has the Potential to Unleash the Host Immunity against Cancer Cells
by Karim Hawillo, Samira Kemiha and Hervé Técher
Onco 2024, 4(4), 322-334; https://doi.org/10.3390/onco4040022 (registering DOI) - 14 Oct 2024
Viewed by 232
Abstract
Chromosomal instability and DNA damage are hallmarks of cancers that can result in the accumulation of micronuclei, cytosolic chromatin fragments (CCFs), or cytosolic DNA species (cytoDNA). The cyclic GMP-AMP synthase (cGAS) is a DNA sensor that recognizes cytosolic DNA and chromatin fragments and [...] Read more.
Chromosomal instability and DNA damage are hallmarks of cancers that can result in the accumulation of micronuclei, cytosolic chromatin fragments (CCFs), or cytosolic DNA species (cytoDNA). The cyclic GMP-AMP synthase (cGAS) is a DNA sensor that recognizes cytosolic DNA and chromatin fragments and subsequently triggers a systemic type I interferon response via the cGAS-STING pathway. Although cancer cells usually contain a high level of chromosomal instability, these cells can avoid the induction of the interferon (IFN) response either by silencing cGAS-STING or the upregulation of the three prime exonuclease 1 (TREX1). TREX1 restricts the spontaneous activation of the cGAS-STING pathway through the degradation of cytoDNA; this in turn limits tumor immunogenicity allowing cancer cells to evade immune detection. Deletion of TREX1 in different cancer types has been shown to decrease tumor growth and increase tumor immune infiltration in pre-clinical mice models. These recent studies also showed the efficacy of TREX1-targeting in combination with anti-PD-1 immune checkpoint blockade. Therefore, targeting TREX1 represents a unique therapeutic strategy to induce an amplified induction of a type I IFN response, promoting the host’s immune response against chromosomally unstable cancer cells. We here discuss these recent advances obtained in preclinical cancer models that pave the way to develop TREX1 inhibitors and to find new avenues to target the broad cGAS-STING pathway signaling in cancer therapy. Full article
Show Figures

Figure 1

Figure 1
<p><b>cGAS-STING signaling pathway</b>. The DNA sensor cGAS recognizes double-stranded DNA from exogenous and endogenous (self) origin. Exogenous DNA can for instance be exposed in the cytosol following viral infection. Endogenous DNA can accumulate in the cytosol as a consequence of DNA damage. cGAS synthesizes the cyclic dinucleotide cGAMP that subsequently binds to and activates STING (stimulator of interferon genes), which is anchored to the endoplasmic reticulum (ER). STING activation is followed by its translocation to the Golgi and sequential phosphorylation of TBK1 and IRF3, which in turn translocates into the nucleus to induce the expression of type I interferons (type I IFN). TREX1 is an exonuclease that is mainly localized into the cytoplasm and that negatively regulates cGAS-STING-IFN signaling by degrading cytosolic DNA (cytoDNA). See the main text for details and references.</p>
Full article ">Figure 2
<p><b>Cancer cells upregulate TREX1 expression to escape immune detection</b>. Some cancer cells increase the expression of TREX1 exonuclease. Although cancer cells experience DNA damage and have unstable genomes, cytosolic DNA (cytoDNA) is constantly degraded by TREX1. As a result, high TREX1 expression in cancer cells suppresses the cGAS-STING signaling. Inactive STING favors the escape to immune surveillance, notably by T cells. See main text for details and references.</p>
Full article ">Figure 3
<p>The impact of TREX1 inactivation and of an interferon response on the tumor microenvironment. Inactivation of TREX1 in cancer cells promotes the accumulation of cytosolic DNA resulting in cGAS-STING-IFN-I signaling. The pro-inflammatory response, the type I IFN response, notably induced by TREX1 inactivation leads to the attraction and recruitment of innate and adaptative immune cells. Either TREX1 ablation or STING agonists (MSA-2 and SR-717) have been shown to boost anti-cancer immunity notably by remodeling the tumor microenvironment, with infiltration of dendritic cells and activation of T lymphocytes. Immune cells and normal cells may then spread this inflammatory response. For instance, dendritic cells can activate STING signaling upon phagocytosis of tumor DNA. Prolonged type I IFN response could promote genomic instability in both cancer and normal cells. See main text for details and references.</p>
Full article ">
14 pages, 697 KiB  
Review
Pericardial Disease in Patients with Cancer: Clinical Insights on Diagnosis and Treatment
by Laia Lorenzo-Esteller, Raúl Ramos-Polo, Alexandra Pons Riverola, Herminio Morillas, Javier Berdejo, Sonia Pernas, Helena Pomares, Leyre Asiain, Alberto Garay, Evelyn Martínez Pérez, Santiago Jiménez-Marrero, Lidia Alcoberro, Ernest Nadal, Paula Gubern-Prieto, Francisco Gual-Capllonch, Encarna Hidalgo, Cristina Enjuanes, Josep Comin-Colet and Pedro Moliner
Cancers 2024, 16(20), 3466; https://doi.org/10.3390/cancers16203466 - 12 Oct 2024
Viewed by 552
Abstract
Pericardial disease is increasingly recognized in cancer patients, including acute pericarditis, pericardial effusion, and constrictive pericarditis, often indicating a poor prognosis. Acute pericarditis arises from direct tumor involvement, cancer therapies, and radiotherapy. Immune checkpoint inhibitor (ICI)-related pericarditis, though rare, entails significant mortality risk. [...] Read more.
Pericardial disease is increasingly recognized in cancer patients, including acute pericarditis, pericardial effusion, and constrictive pericarditis, often indicating a poor prognosis. Acute pericarditis arises from direct tumor involvement, cancer therapies, and radiotherapy. Immune checkpoint inhibitor (ICI)-related pericarditis, though rare, entails significant mortality risk. Treatment includes NSAIDs, colchicine, and corticosteroids or anti-IL1 drugs in refractory cases. Pericardial effusion is the most frequent manifestation, primarily caused by lung cancer, followed by breast cancer, lymphoma, leukemia, gastrointestinal tumors, and melanoma. Chemotherapy, immunotherapy, and radiotherapy may also cause fluid accumulation in the pericardial space. Symptomatic relief for pericardial effusion may require pericardiocentesis, prolonged catheter drainage, or a pericardial window. Instillation of intrapericardial cytostatic agents may reduce recurrence. Constrictive pericarditis, though less common, often develops from radiotherapy and requires multimodality imaging for diagnosis, with pericardiectomy as the definitive treatment. Primary pericardial tumors are rare, with metastases being more frequent. Patients with cancer and pericardial disease generally have poor survival, emphasizing the need for early detection. A multidisciplinary approach involving hematologists, oncologists, and cardiologists is crucial to tailoring pericardial disease treatment to a patient’s clinical status, thereby improving the quality of life and prognosis. Full article
(This article belongs to the Special Issue Feature Paper in Section 'Cancer Epidemiology and Prevention' in 2024)
Show Figures

Figure 1

Figure 1
<p>Clinical manifestations of pericardial disease.</p>
Full article ">Figure 2
<p>Pericardiocentesis approaches (subxiphoid, apical, and left parasternal).</p>
Full article ">
13 pages, 4324 KiB  
Review
The Recent Research Progress of the Tumor mRNA Vaccine
by Hao Zhao, Miying Li, Jiaren Zhou, Lidan Hu, Shaohong Lu and Pan Li
Vaccines 2024, 12(10), 1167; https://doi.org/10.3390/vaccines12101167 - 12 Oct 2024
Viewed by 837
Abstract
Tumors have long posed a significant threat to human life and health, and the messenger ribonucleic acid (mRNA) vaccine is seen as an attractive approach for cancer immunotherapy due to its developmental simplicity, rapid manufacture, and increased immune safety and efficiency. In this [...] Read more.
Tumors have long posed a significant threat to human life and health, and the messenger ribonucleic acid (mRNA) vaccine is seen as an attractive approach for cancer immunotherapy due to its developmental simplicity, rapid manufacture, and increased immune safety and efficiency. In this review, we have summarized details of the developmental history of mRNA vaccines, discussed the basic molecular structure and the effect on the stable and translation level of mRNA, analyzed the underlying immune efficiency and mechanisms on tumors, and assessed the current status of clinical research. We explored the treatment and application prospects of mRNA vaccines, aiming to provide perspectives on the future of mRNA tumor vaccines for ongoing clinical research. Full article
(This article belongs to the Special Issue mRNA Vaccines: Pioneering the Future of Vaccination)
Show Figures

Figure 1

Figure 1
<p>Time axis of mRNA vaccine development: summary of the time axis of mRNA vaccine development from 1990 to 2023. pDNA: Plasmid DNA; CNE: Cationic Nanoparticle Emulsion; LNP: Lipid Nanoparticle; DOTAP: N-1-(2,3-diethoxy) propyl-n, n, n-trimethylammonium sulfate; DC: Dendritic cells.</p>
Full article ">Figure 2
<p>Diagram of the structures of three different types of mRNA vaccine. This section introduces the fundamental structures of three distinct types of mRNA and the various roles played by their components. By modifying and optimizing specific key elements, mRNA can enhance its efficiency and stability during the translation process. UTR: Untranslated Region, uORF: upstream open reading frame, saRNA: self-amplifying mRNA, circRNA: circular mRNA, CDS: coding DNA sequence, GOI: gene of interest, NSP: non-structural protein.</p>
Full article ">Figure 3
<p>Chemical structure of 5′-cap of mRNA. In this specific RNA molecule, the methylation patterns presented by the initial few 5′ nucleotides are observed, and distinct cap variants such as CAP 0, CAP 1, CAP 2, and so on, can be clearly differentiated. The determination of these cap variants depends on the differences in the methylation patterns (as shown in the figure). B1, B2: nucleobases.</p>
Full article ">Figure 4
<p>The immune mechanisms of mRNA tumor vaccines in vivo. The mRNA vaccine consisting of mRNA encapsulated by liposomes is absorbed by dendritic cells (DCs). Within DCs, mRNA undergoes the steps of transcription and translation to generate antigens, which are subsequently presented to T cells through MHC I or MHC II. Moreover, through the synergistic action of cytokines including interleukin-1 (IL-1), interleukin-2 (IL-2), and interleukin-12 (IL-12), the cellular immune pathway is activated. In particular, antigens secreted by APCs can also activate B lymphocytes. Under the action of activated CD4<sup>+</sup> T cells, B lymphocytes will secrete corresponding neutralizing antibodies, further enhancing the immune effect on tumors. Immune activation progress: ① vaccine internalization, ② and ③ encoding protein, ④ protein delivery, ⑤, ⑥ and ⑦ activating immunity.</p>
Full article ">
21 pages, 2591 KiB  
Review
Tertiary Lymphoid Structures in Microorganism-Related Cancer
by Shuzhe Deng, Xinxin Yang, Lin He, Yunjing Hou and Hongxue Meng
Cancers 2024, 16(20), 3464; https://doi.org/10.3390/cancers16203464 - 12 Oct 2024
Viewed by 386
Abstract
Tertiary lymphoid structures (TLSs) are ectopic lymphoid tissues formed by the accumulation of lymphocytes and other components outside lymphoid organs. They have been shown to be widespread in cancers and have predictive effects on prognosis and immunotherapy efficacy; however, there is no standardized [...] Read more.
Tertiary lymphoid structures (TLSs) are ectopic lymphoid tissues formed by the accumulation of lymphocytes and other components outside lymphoid organs. They have been shown to be widespread in cancers and have predictive effects on prognosis and immunotherapy efficacy; however, there is no standardized measurement guide. This paper provides a reference for future research. Moreover, the induction strategy for the formation mechanism of TLSs is a new direction for future cancer treatment, such as cancer vaccines for microorganisms. The effects of microorganisms on cancer are dual. The role of microorganisms, including bacteria, parasites, viruses, and fungi, in promoting cancer has been widely confirmed. However, the specific mechanism of their tumor suppressor effect, particularly the promotion of TLS formation, is currently unknown. In this review, we summarize the role of TLSs in cancer related to microbial infection and provide new ideas for further understanding their mechanisms of action in cancer. Full article
(This article belongs to the Section Cancer Pathophysiology)
Show Figures

Figure 1

Figure 1
<p>Structure and composition of TLSs. Tertiary lymphoid structures (TLSs) mainly consist of aggregated CD20<sup>+</sup> B cells as the core and CD3<sup>+</sup> T cells wrapped around the outside. In the B-cell realm, CD83<sup>+</sup> dendritic cells (DCs) and CD21<sup>+</sup> follicular dendritic cells (FDCs) are the most frequent, followed by CD23<sup>+</sup>BCL6<sup>+</sup> germinal center (GC) B cells and CD38<sup>+</sup>/CD138<sup>+</sup> plasma cells. In the T-cell realm, CD8<sup>+</sup> T cells and CD4<sup>+</sup> T cells containing follicular helper T (T<sub>FH</sub>) cells and regulatory T (Treg) cells are mainly observed, whereas CD19<sup>+</sup> regulatory B (Breg) cells are observed in some instances. Importantly, high endothelial venules (HEVs) can be observed in the periphery. CD68<sup>+</sup> macrophages and neutrophils are also scattered in TLSs. Therefore, molecules with important functions should not be ignored. The figure was drawn using the MedPeer program.</p>
Full article ">Figure 2
<p>Detection and classification of TLSs. First, it should be determined whether tertiary lymphoid structures (TLSs) exist. TLS-negative status was determined when there were no immune cells in or around the tumor and there were focally distributed immune cells without aggregation. The main detection method of TLSs is hematoxylin and eosin (HE) staining, by which the quantity, density, and location can be observed. There are two types of location-based classification. The first is that TLSs can be divided into intratumoral and peritumoral based on the spatial location, and the former can be further subdivided into parenchymatous and stromal. The second is that TLSs can be divided into superficial and deep layers based on the infiltration location. Combined with immunohistochemistry (IHC) or immunofluorescence (IF) based on HE, TLS differentiation can be divided into lymphocyte cluster, immature follicle, and mature follicle. In addition, TLSs can be studied at the cellular and gene levels using flow cytometry (FCM) and sequencing technology. At present, evidence of intercellular interaction and multiple gene signatures have been obtained. The figure was drawn using the MedPeer program.</p>
Full article ">Figure 3
<p>TLSs in microorganism-related cancer (×400). Representative images of tertiary lymphoid structures (TLSs) detected in formalin fixation with paraffin-embedding (FFPE) tumor sections via hematoxylin and eosin (HE) staining or immunohistochemistry (IHC) staining showing CD3<sup>+</sup> T cells, CD20<sup>+</sup> B cells, CD21<sup>+</sup> follicular dendritic cells (FDCs), and CD23<sup>+</sup> germinal centers (GCs). These tumor tissues contained human papilloma virus (HPV)-associated head and neck squamous-cell carcinoma (HNSCC) and Epstein-barr virus (EBV)-associated gastric cancer, and there was no significant difference in the morphological structure of TLSs. These histological images are original, unpublished images from the authors’ examination of tumors.</p>
Full article ">Figure 4
<p>Generation mechanism and induced strategy of TLSs. Inflammation-related molecules such as CXC-chemokine ligand 13 (CXCL13) attract aggregation of lymphoid tissue inducer (LTi) cells which communicate with stromal cells through LTα1β2-LTβR, TNF-TNFR1, IL7-IL7R, and IL17-IL17R signaling pathways to promote the excretion of vascular endothelial growth factor C (VEGFC) to irritative high endothelial venules (HEV) production. Then, the excretion of chemokines and adhesion factors induce the homing of peripheral lymphocytes to the HEVs and control their entry into specific areas to form tertiary lymphoid structures (TLSs). It is noteworthy that the secretion of CXCL13 can induce the expression of lymphotoxin-α1β2 (LTα1β2) on B cells or LTi cells through the CXCL13-CXCR5 axis and form a positive feedback loop, driving expansions of stromal cells such as cancer-associated fibroblasts (CAFs) and TLSs. Except for the classical pathway, the mechanism of microbially induced TLS formation has been elucidated gradually. The formation and maturation of microbial-driven TLSs are closely related to the antigen recognition involved in follicular helper T (T<sub>FH</sub>) cells, which may be because T<sub>FH</sub> cells can produce TLS-dependent CXCL13 to recruit B cells. Based on the mechanism of TLSs, it is feasible to induce TLSs through intratumoral injection or biomaterial implantation of cytokines and chemokines. Direct infusion of lymphoid tissue organizer (LTo) cells and intramuscular injection of vaccines may also promote TLS formation. The figure was drawn using the MedPeer program.</p>
Full article ">
16 pages, 2235 KiB  
Review
Unraveling the Heterogeneity of Deficiency of Mismatch Repair Proteins in Endometrial Cancer: Predictive Biomarkers and Assessment Challenges
by Filomena M. Carvalho and Jesus P. Carvalho
Cancers 2024, 16(20), 3452; https://doi.org/10.3390/cancers16203452 - 11 Oct 2024
Viewed by 509
Abstract
Endometrial cancer (EC) poses a significant global health challenge, with increasing prevalence in 26 of 43 countries and over 13,000 deaths projected in the United States by 2024. This rise correlates with aging populations, the obesity epidemic, and changing reproductive patterns, including delayed [...] Read more.
Endometrial cancer (EC) poses a significant global health challenge, with increasing prevalence in 26 of 43 countries and over 13,000 deaths projected in the United States by 2024. This rise correlates with aging populations, the obesity epidemic, and changing reproductive patterns, including delayed childbearing. Despite the early diagnosis in 67% of cases, approximately 30% of cases present with regional or distant spread, leading to nearly 20% mortality rates. Unlike many cancers, EC mortality rates are escalating, outpacing therapeutic advancements until recently. One of the reasons for this was the lack of effective therapeutic options for advanced disease until recently. The introduction of immunotherapy has marked a turning point in EC treatment, particularly benefiting patients with defects in mismatch repair proteins (dMMRs). However, dMMR status alone does not ensure a favorable response, underscoring the need for precise patient selection. This review explores the pivotal role of mismatch repair proteins in EC, emphasizing their heterogeneity, the challenges in their assessment, and their potential as predictive biomarkers. Full article
(This article belongs to the Special Issue Pathology of Gynecological Cancers)
Show Figures

Figure 1

Figure 1
<p>The immunohistochemistry of mismatch repair proteins in endometrial cancer. (<b>a</b>) Intact expression of MSH6—it is important to note the higher intensity in tumor cells than in stromal cells; (<b>b</b>) the loss of MLH1 expression. Tumor cells are negative, while stromal cells are positive.</p>
Full article ">Figure 2
<p>The testing flowchart used to determine the MMR/MSI status in endometrial cancer, based on the work of Noh et al. [<a href="#B39-cancers-16-03452" class="html-bibr">39</a>].</p>
Full article ">Figure 3
<p>The subclonal loss of MLH1 in a case of dedifferentiated carcinoma. On the right, the undifferentiated component does not have staining, while the endometrioid component (left) and stroma show intact expression.</p>
Full article ">Figure 4
<p>The many facets of deficient mismatch repair/microsatellite instability-high endometrial carcinomas.</p>
Full article ">
7 pages, 480 KiB  
Case Report
Diaphragmatic Palsy Due to a Paraneoplastic Autoimmune Syndrome Revealed by Checkpoint Inhibitors
by Jean-Baptiste Destival, Jean-Marie Michot, Cécile Cauquil, Nicolas Noël, Salima Hacein-Bey-Abina, Pascale Chrétien and Olivier Lambotte
Reports 2024, 7(4), 84; https://doi.org/10.3390/reports7040084 - 11 Oct 2024
Viewed by 382
Abstract
Background and Clinical Significance: Immune checkpoint inhibitors (ICIs) have revolutionized cancer treatment but may underlie diverse and potentially life-threatening immune-related adverse events (irAEs). They may cause various conditions leading to respiratory failure, including myasthenic syndromes and myositis. However, diaphragmatic paralysis (DP) has rarely been [...] Read more.
Background and Clinical Significance: Immune checkpoint inhibitors (ICIs) have revolutionized cancer treatment but may underlie diverse and potentially life-threatening immune-related adverse events (irAEs). They may cause various conditions leading to respiratory failure, including myasthenic syndromes and myositis. However, diaphragmatic paralysis (DP) has rarely been reported. To describe patients with diaphragmatic paralysis in a pharmacovigilance registry, we searched the prospective REISAMIC registry at the Gustave Roussy Cancer Center (Villejuif, France) for cases of diaphragmatic palsy (DP) occurring from September 2014 to December 2021. Case Presentation: We identified three patients, in whom DP was confirmed by diaphragmatic ultrasonography, pulmonary function tests, and/or diaphragmatic electroneuromyogram. Diaphragmatic palsy was life-threatening in all patients, as it caused respiratory failure requiring mechanical ventilation. In all cases, a pre-existing subclinical paraneoplastic syndrome was detected. Onconeural antibodies (anti-titin and anti-VGCC) were detected in these patients before and after the initiation of ICI therapy, suggesting a mixed paraneoplastic syndrome with features overlapping those of myasthenic syndrome (myasthenia gravis in one patient and Lambert–Eaton syndrome in another) and myositis. Conclusions: Diaphragmatic palsy is a severe irAE potentially resulting from different mechanisms, including myositis and neuromuscular junction involvement (myasthenia gravis, Lambert–Eaton). Antineuronal antibodies associated with such conditions were already present in our patients prior to immunotherapy initiation, suggesting ICIs could trigger flare-ups of pre-existing silent paraneoplastic autoimmune conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Antibody testing in each patient, before and after the first administration of immune-checkpoint inhibitor (#1 ICI). D0 Hospitalization marks the day of admission for respiratory failure. + indicates a positive test, − indicates a negative test. Antibody titers are indicated in UA (antibody unit) or pM (pmmol/L). T means trace of antibody. D: day, M: month, VGCC: voltage-gated calcium channel, Ab: antibody.</p>
Full article ">
Back to TopTop