[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (547)

Search Parameters:
Keywords = oxaliplatin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 2138 KiB  
Article
Global Leadership Initiative on Malnutrition Criteria and Immunonutritional Status Predict Chemoadherence and Survival in Stage II/III Gastric Cancer Treated with XELOX Chemotherapy
by Jong Hyuk Yun, Geum Jong Song, Myoung Won Son and Moon Soo Lee
Nutrients 2024, 16(20), 3468; https://doi.org/10.3390/nu16203468 - 14 Oct 2024
Viewed by 384
Abstract
Backgroud: Adjuvant chemotherapy is crucial for the treatment of advanced gastric cancer. However, various factors negatively impact chemoadherence, with malnutrition after gastrectomy being a critical determinant. This study aims to analyze the impact of malnutrition, assessed through the Global Leadership Initiative on Malnutrition [...] Read more.
Backgroud: Adjuvant chemotherapy is crucial for the treatment of advanced gastric cancer. However, various factors negatively impact chemoadherence, with malnutrition after gastrectomy being a critical determinant. This study aims to analyze the impact of malnutrition, assessed through the Global Leadership Initiative on Malnutrition (GLIM) and other immunonutritional indices, on chemoadherence and its subsequent effect on survival. Methods: This retrospective study included 116 patients who underwent curative gastrectomy and received oxaliplatin and capecitabine (XELOX). Preoperative nutritional status was assessed using the GLIM criteria along with other immunonutritional indices, such as the prognostic nutritional index (PNI), C-reactive protein-to-albumin ratio (CAR), neutrophil–lymphocyte ratio (NLR), controlling nutritional status (CONUT) score, and modified Glasgow Prognostic Score (mGPS). Chemotherapy adherence was measured using relative dose intensity (RDI). Statistical analyses included least absolute shrinkage and selection operator (LASSO) regression to identify the key predictors of RDI and Cox proportional hazards models and assess the impact on survival. Results: Overall, 116 patients were included in this analysis. In the multivariate analysis using LASSO regression, higher GLIM severity was independently associated with a lower RDI (coefficient = −0.0216; p < 0.01). Other significant factors influencing RDI included older age (p < 0.01), female sex (p = 0.02), higher mGPS (p = 0.03), higher CONUT score (p = 0.04), and higher CAR (p = 0.05), all of which were associated with a lower RDI. The Cox proportional hazards analysis revealed that higher RDI was significantly associated with better survival (hazard ratio [HR] = 0.06; p < 0.005). Conclusions: This study highlights the critical role of immunonutritional status, particularly as measured using the GLIM criteria, in maintaining adherence to chemotherapy and improving survival outcomes in patients with gastric cancer. Routine preoperative nutritional assessments using GLIM can help identify high-risk patients, and early nutritional interventions may improve chemotherapy adherence and outcomes. These findings support the integration of nutritional strategies, specifically targeting those identified by the GLIM, into standard care to enhance the efficacy and survival of chemotherapy. Full article
Show Figures

Figure 1

Figure 1
<p>Body composition analysis at the L3 level using abdominal pelvic CT (APCT). Evaluation of the skeletal muscle index (SMI), psoas muscle index (PMI), subcutaneous fat index (SFI), and visceral fat index (VFI) areas using non-contrast-enhanced APCT imaging. The figure compares a patient with severely reduced muscle mass according to the GLIM criteria (<b>B</b>) against a patient with normal muscle mass (<b>A</b>).</p>
Full article ">Figure 2
<p>Study flowchart of patients with gastric cancer who underwent gastrectomy.</p>
Full article ">Figure 3
<p>Univariate analysis of the associations between nutritional inflammatory indices and relative dose intensity (RDI). (<b>A</b>)<b>:</b> The modified Glasgow Prognostic Score (mGPS) was significantly negatively associated with RDI (H = 7.473, <span class="html-italic">p</span> = 0.024). (<b>B</b>)<b>:</b> The C-reactive protein-to-albumin ratio (CAR) showed a significant negative correlation with RDI (Spearman’s r = −0.285, <span class="html-italic">p</span> = 0.0019). (<b>C</b>)<b>:</b> The Global Leadership Initiative on Malnutrition (GLIM) criteria showed a significant decrease in RDI with increasing malnutrition severity (H = 9.468, <span class="html-italic">p</span> = 0.0088).</p>
Full article ">Figure 4
<p>Correlation matrix and LASSO regression analysis of predictors related to RDI. (<b>A</b>) Correlation matrix of relevant clinical and nutritional variables, including age, sex, GLIM criteria, stage, psoas muscle index (PMI), skeletal muscle index (SMI), neutrophil-to-lymphocyte ratio (NLR), prognostic nutritional index (PNI), controlling nutritional status (CONUT), modified Glasgow Prognostic Score (mGPS), C-reactive protein-to-albumin ratio (CAR), and relative dose intensity (RDI). The color gradient represents the strength and direction of the correlation (from −1.0 to +1.0), (<b>B</b>) LASSO (least absolute shrinkage and selection operator) regression coefficient plot showing the influence of significant predictors on RDI. The variables include age, sex (female), GLIM criteria, mGPS, CONUT, and CAR, with negative coefficients indicating a stronger association with lower RDI.</p>
Full article ">Figure 4 Cont.
<p>Correlation matrix and LASSO regression analysis of predictors related to RDI. (<b>A</b>) Correlation matrix of relevant clinical and nutritional variables, including age, sex, GLIM criteria, stage, psoas muscle index (PMI), skeletal muscle index (SMI), neutrophil-to-lymphocyte ratio (NLR), prognostic nutritional index (PNI), controlling nutritional status (CONUT), modified Glasgow Prognostic Score (mGPS), C-reactive protein-to-albumin ratio (CAR), and relative dose intensity (RDI). The color gradient represents the strength and direction of the correlation (from −1.0 to +1.0), (<b>B</b>) LASSO (least absolute shrinkage and selection operator) regression coefficient plot showing the influence of significant predictors on RDI. The variables include age, sex (female), GLIM criteria, mGPS, CONUT, and CAR, with negative coefficients indicating a stronger association with lower RDI.</p>
Full article ">Figure 5
<p>Kaplan–Meier survival curves based on RDI and stage. (<b>A</b>) Kaplan–Meier survival curve showing the difference in survival probability between patients with relative dose intensity (RDI) ≥ 0.8 and those with RDI &lt; 0.8. Shaded areas represent the 95% confidence intervals for each group, (<b>B</b>) Kaplan–Meier survival curves stratified by both RDI (high vs. low) and cancer stage (stage II vs. stage III). The survival curves compare the following four groups: RDI low, stage II; RDI low, stage III; RDI high, stage II; and RDI high, stage III. Shaded areas represent the 95% confidence intervals for each group.</p>
Full article ">
25 pages, 9863 KiB  
Article
Targeting PARP-1 and DNA Damage Response Defects in Colorectal Cancer Chemotherapy with Established and Novel PARP Inhibitors
by Philipp Demuth, Lea Thibol, Anna Lemsch, Felix Potlitz, Lukas Schulig, Christoph Grathwol, Georg Manolikakes, Dennis Schade, Vassilis Roukos, Andreas Link and Jörg Fahrer
Cancers 2024, 16(20), 3441; https://doi.org/10.3390/cancers16203441 - 10 Oct 2024
Viewed by 541
Abstract
The DNA repair protein PARP-1 emerged as a valuable target in the treatment of tumor entities with deficiencies of BRCA1/2, such as breast cancer. More recently, the application of PARP inhibitors (PARPi) such as olaparib has been expanded to other cancer entities [...] Read more.
The DNA repair protein PARP-1 emerged as a valuable target in the treatment of tumor entities with deficiencies of BRCA1/2, such as breast cancer. More recently, the application of PARP inhibitors (PARPi) such as olaparib has been expanded to other cancer entities including colorectal cancer (CRC). We previously demonstrated that PARP-1 is overexpressed in human CRC and promotes CRC progression in a mouse model. However, acquired resistance to PARPi and cytotoxicity-mediated adverse effects limit their clinical applicability. Here, we detailed the role of PARP-1 as a therapeutic target in CRC and studied the efficacy of novel PARPi compounds in wildtype (WT) and DNA repair-deficient CRC cell lines together with the chemotherapeutics irinotecan (IT), 5-fluorouracil (5-FU), and oxaliplatin (OXA). Based on the ComPlat molecule archive, we identified novel PARPi candidates by molecular docking experiments in silico, which were then confirmed by in vitro PARP activity measurements. Two promising candidates (X17613 and X17618) also showed potent PARP-1 inhibition in a CRC cell-based assay. In contrast to olaparib, the PARPi candidates caused no PARP-1 trapping and, consistently, were not or only weakly cytotoxic in WT CRC cells and their BRCA2- or ATR-deficient counterparts. Importantly, both PARPi candidates did not affect the viability of nonmalignant human colonic epithelial cells. While both olaparib and veliparib increased the sensitivity of WT CRC cells towards IT, no synergism was observed for X17613 and X17618. Finally, we provided evidence that all PARPi (olaparib > veliparib > X17613 > X17618) synergize with chemotherapeutic drugs (IT > OXA) in a BRCA2-dependent manner in CRC cells, whereas ATR deficiency had only a minor impact. Collectively, our study identified novel lead structures with potent PARP-1 inhibitory activity in CRC cells but low cytotoxicity due to the lack of PARP-1 trapping, which synergized with IT in homologous recombination deficiency. Full article
(This article belongs to the Special Issue Cancer Chemotherapy: Combination with Inhibitors (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Binding modes of veliparib (<b>1</b>), PDB: 7AAC), olaparib ((<b>2</b>), PDB: 7AAD), and selected compounds from virtual screening ((<b>3</b>–<b>9</b>), PDB: 4PJT) to PARP-1. The binding to either G863 or S904, as also found for veliparib, was used as a constraint in docking. All active compounds are able to form this bond and adopt a similar binding mode. Through the indole NH, there is an interaction with E988 by a bridging water molecule. No preference between the binding modes of the S- (<b>4</b>–<b>6</b>) or R-enantiomers (<b>7</b>–<b>9</b>) is observed, while the scoring values also differ only slightly. (<b>B</b>) Chemical structure of the most active compounds X17613, X17618, X17620, and X17621, according to in vitro screening and the two established PARPi veliparib and olaparib.</p>
Full article ">Figure 2
<p>(<b>A</b>) Concentration–response curves of four potential PARP-1 inhibitors with the highest activity in the PARP-1 screening assay kit. All concentrations were tested in duplicates. IC<sub>50</sub> values were derived using a nonlinear regression model in GraphPad Prism 9 (<span class="html-italic">n</span> = 2). (<b>B</b>) Investigation of PARP inhibition by X17613, X17618, X17620, and X17621 in HCT116 cells. Cells were challenged with 1 mM H<sub>2</sub>O<sub>2</sub> for 5 min and pretreated or not with the indicated compounds for 2 h. PAR synthesis was identified by confocal IF microscopy using the PAR 10H antibody. The signal intensity of five images per concentration was evaluated by ImageJ (<span class="html-italic">n</span> ≥ 3). (<b>C</b>) Representative confocal microscopy images at 100× magnification after PAR staining in HCT116 cells treated with the indicated concentrations of X17613 for 2 h with or without subsequent PARP activation by H<sub>2</sub>O<sub>2</sub> treatment for 5 min. Scale bar: 100 µm. (<b>D</b>) Confocal microscopy images at 630× magnification after pan-PAR staining in HCT116 cells treated according to (<b>C</b>). Scale bar: 20 µm. Data are presented as mean +/− SEM. * <span class="html-italic">p</span> &lt; 0.01, ** <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>(<b>A</b>,<b>B</b>) Analysis of PARP-1 trapping in HCT116 and Caco-2 cells. Immunoblot detection of PARP-1 after pre-treatment with X17613, X17618, and olaparib followed by MMS exposure for 1 h and chromatin isolation. The cytosolic marker Hsp90 and the chromatin marker Histone H3 served as respective loading controls. Representative Western blot images and densitometric evaluation are shown (<span class="html-italic">n</span> = 3). Data are shown as mean + SEM. (<b>C</b>) Cell viability determined by the resazurin reduction assay (RRA) in HCT116 PARP-1<sup>−/−</sup> and HCT116 PARP-1<sup>+/+</sup> cells after PARPi treatment for 72 h. A nonlinear regression curve fit was conducted using GraphPad Prism 9 (<span class="html-italic">n</span> ≥ 3). (<b>D</b>) Viability in Caco-2 cells after exposure to PARPi as indicated. (<span class="html-italic">n</span> = 3). All data are shown as mean +/− SEM.</p>
Full article ">Figure 4
<p>(<b>A</b>) Toxicity of PARPi in HCT116 cells depending on BRCA2 status. HCT116 WT and HCT116 BRCA2<sup>−/−</sup> cells were incubated with PARPi for 72 h and viability was assessed using the resazurin reduction assay (RRA). Nonlinear regression curve fit was conducted using GraphPad Prism 9 (<span class="html-italic">n</span> ≥ 3). (<b>B</b>) Toxicity of PARPi in DLD-1 cells depending on ATR status. DLD-1 WT and DLD-1 ATR<sup>s/s</sup> cells were incubated with PARPi for 72 h and viability was assessed using the RRA. Nonlinear regression curve fit was conducted using GraphPad Prism 9 (<span class="html-italic">n</span> ≥ 3). Data are depicted as mean +/− SEM.</p>
Full article ">Figure 5
<p>(<b>A</b>) Viability in HCT116 PARP-1<sup>−/−</sup> and HCT116 PARP-1<sup>+/+</sup> cells after treatment with PARPi olaparib or veliparib in combination with chemotherapeutic drugs irinotecan (IT, 0.5 µM), 5-fluorouracil (5-FU, 0.25 µM), and oxaliplatin (OXA, 0.5 µM) for 72 h (<span class="html-italic">n</span> ≥ 3). (<b>B</b>) Viability in HCT116 PARP-1<sup>−/−</sup> and HCT116 PARP-1<sup>+/+</sup> cells after treatment with PARPi X17613 in combination with chemotherapeutic drugs for 72 h. Data (<span class="html-italic">n</span> ≥ 3) are given as mean +/− SEM. (<b>C</b>,<b>D</b>) Viability in Caco-2 cells after treatment with PARPi in combination with chemotherapeutic drugs irinotecan (IT, 10 µM), 5-fluorouracil (5-FU, 5 µM), and oxaliplatin (OXA, 1 µM). Data (<span class="html-italic">n</span> = 3) are shown as mean +/− SEM. 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; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001; <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p>(<b>A</b>) Viability in HCT116 WT and HCT116 BRCA2<sup>−/−</sup> cells after treatment with PARPi olaparib or veliparib in combination with chemotherapeutic drugs (IT, 0.25 µM), 5-fluorouracil (5-FU, 0.1 µM), and oxaliplatin (OXA, 0.25 µM) for 72 h (<span class="html-italic">n</span> ≥ 3). (<b>B</b>) Viability in HCT116 WT and HCT116 BRCA2<sup>−/−</sup> cells after treatment with PARPi X17613 in combination with chemotherapeutic drugs (IT, 0.25 µM), 5-fluorouracil (5-FU, 0.1 µM), and oxaliplatin (OXA, 0.25 µM) for 72 h (<span class="html-italic">n</span> ≥ 3). (<b>C</b>) Representative brightfield microscopic images at 20X magnification of HCT116 WT and HCT116 BRCA2<sup>−/−</sup> cells after treatment with X17613 (50 µM), IT (0.25 µM), or a combination of both for 24 h. (<b>D</b>,<b>E</b>) γH2AX formation in HCT116 WT and BRCA2<sup>−/−</sup> cells after treatment as described in (<b>C</b>). Representative Western blot images and densitometric evaluation are shown (<span class="html-italic">n</span> = 4). Hsp90 served as loading control. All data are given as mean + SEM. 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; 0.01; <span class="html-italic">t</span>-test.</p>
Full article ">Figure A1
<p>(<b>A</b>) Binding modes of selected compounds from virtual screening ((<b>1</b>–<b>4</b>), PDB: 4PJT) to PARP-1. The binding to either G863 or S904 was used as a constraint in docking. All active compounds are able to form this bond and adopt a similar binding mode (compare <a href="#cancers-16-03441-f001" class="html-fig">Figure 1</a>A). The three compounds X17611, X17610, and X17608 (<b>1</b>–<b>3</b>) serve as a verification of the binding mode since the constraints can only be fulfilled with significant losses in the binding free energy due to unfavorable inter- and intramolecular interactions. X17616 (<b>4</b>) is similar to X17618 (<a href="#cancers-16-03441-f001" class="html-fig">Figure 1</a>A), but cannot form a potential hydrogen bond due to the lack of the amide function. (<b>B</b>) Chemical structure of the compounds X17611, X17610, X17608 and X17616.</p>
Full article ">Figure A2
<p>(<b>A</b>,<b>B</b>) Concentration response curve of established PARP inhibitors veliparib and olaparib (<b>A</b>) and eight potential PARP inhibitors (<b>B</b>) with low or no activity in the PARP-1 screening assay kit. All concentrations were tested in duplicates. IC<sub>50</sub> values were derived using a nonlinear regression model in GraphPad Prism 9 (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure A3
<p>(<b>A</b>) Concentration-response curves of 4 potential PARP-1 inhibitors (X17613, X17618, X17620 and X17621) assessed in HCT116 cells as described in <a href="#cancers-16-03441-f002" class="html-fig">Figure 2</a>B (<span class="html-italic">n</span> = 3). IC<sub>50</sub> values were derived using a nonlinear regression model in GraphPad Prism 9. (<b>B</b>) Representative confocal microscopy images at 100× magnification after PAR staining in HCT116 cells treated with the indicated concentrations of X17618 for 2 h with or without subsequent PARP induction by H<sub>2</sub>O<sub>2</sub> treatment for 5 min. Scale bar: 100 µm. (<b>C</b>) Confocal microscopy images at 630× magnification after pan-PAR staining in HCT116 treated as described in B. Scale bar: 20 µm.</p>
Full article ">Figure A4
<p>(<b>A</b>) Western Blot analysis of PARP-1, BRCA2 and ATR expression in HCT116 WT, HCT116 BRCA2<sup>−/−</sup>, HCT116 PARP-1<sup>+/+</sup>, HCT116 PARP-1<sup>−/−</sup>, DLD-1 WT and DLD-1 ATR<sup>−/−</sup> cells. HSP90 served as loading control. (<b>B</b>) Toxicity of PARPi in human colonic epithelial cells (HCEC). Cells were incubated with PARPi for 72 h and viability was assessed using the RRA. Nonlinear regression curve fit was conducted using GraphPad Prism 9 (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Figure A5
<p>Cell viability of (<b>A</b>) HCT116 WT and HCT116 BRCA2<sup>−/−</sup>, (<b>B</b>) HCT116 PARP-1<sup>+/+</sup> and HCT116 PARP-1<sup>−/−</sup>, (<b>C</b>) DLD-1 WT and DLD-1 ATR<sup>s/s</sup>, (<b>D</b>) Caco-2 cells and (<b>E</b>) HCEC after monotreatment with cytostatic drugs IT, 5-FU and OXA for 72 h. Nonlinear regression curve fit was conducted using GraphPad Prism 9 (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Figure A6
<p>(<b>A</b>) Viability of HCT116 PARP-1<sup>−/−</sup> and HCT116 PARP-1<sup>+/+</sup> cells after treatment with PARPi X17618 in combination with chemotherapeutic drugs irinotecan (IT, 0.5 µM), 5-fluorouracil (5-FU, 0.25 µM) and oxaliplatin (OXA, 0.5 µM) for 72 h (<span class="html-italic">n</span> ≥ 3) (<b>B</b>) Viability of HCT116 WT and HCT116 BRCA2<sup>−/−</sup> cells after treatment with PARPi X17618 in combination with chemotherapeutic drugs irinotecan (IT, 0.25 µM), 5-fluorouracil (5-FU, 0.1 µM) and oxaliplatin (OXA, 0.25 µM) for 72 h (<span class="html-italic">n</span> ≥ 3). (<b>C</b>) γH2AX formation in HCT116 WT and BRCA2<sup>−/−</sup> cells after treatment with X17618 (50 µM), IT (0.25 µM) or a combination of both for 24 h. Representative Western blot images and densitometric evaluation are shown (<span class="html-italic">n</span> = 4). All data are presented as mean +/− SEM. 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; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001; <span class="html-italic">t</span>-test.</p>
Full article ">Figure A7
<p>(<b>A</b>) Viability of DLD-1 WT and DLD-1 ATR<sup>s/s</sup> cells after treatment with PARPi olaparib or veliparib in combination with chemotherapeutic drugs (IT, 2.5 µM), 5-fluorouracil (5-FU, 0.1 µM) and oxaliplatin (OXA, 5 µM) for 72 h (<span class="html-italic">n</span> ≥ 3). (<b>B</b>) Viability of DLD-1 WT and DLD-1 ATR<sup>s/s</sup> cells after treatment with PARPi X17613 and X17618 in combination with chemotherapeutic drugs (IT, 2.5 µM), 5-fluorouracil (5-FU, 0.1 µM) and oxaliplatin (OXA, 5 µM) for 72 h (<span class="html-italic">n</span> ≥ 3). Data are shown as mean +/− SEM. * <span class="html-italic">p</span> &lt; 0.01, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001; <span class="html-italic">t</span>-test.</p>
Full article ">
13 pages, 2051 KiB  
Article
Tumorspheres as In Vitro Model for Identifying Predictive Chemoresistance and Tumor Aggressiveness Biomarkers in Breast and Colorectal Cancer
by Toni Martinez-Bernabe, Pere Miquel Morla-Barcelo, Lucas Melguizo-Salom, Margalida Munar-Gelabert, Alba Maroto-Blasco, Margalida Torrens-Mas, Jordi Oliver, Pilar Roca, Mercedes Nadal-Serrano, Daniel Gabriel Pons and Jorge Sastre-Serra
Biology 2024, 13(9), 724; https://doi.org/10.3390/biology13090724 - 15 Sep 2024
Viewed by 703
Abstract
Chemoresistance remains a major challenge in the treatment of breast and colorectal cancer. For this reason, finding reliable predictive biomarkers of response to chemotherapy has become a significant research focus in recent years. However, validating in vitro results may be problematic due to [...] Read more.
Chemoresistance remains a major challenge in the treatment of breast and colorectal cancer. For this reason, finding reliable predictive biomarkers of response to chemotherapy has become a significant research focus in recent years. However, validating in vitro results may be problematic due to the outcome heterogeneity. In this study, we evaluate the use of tumorspheres as an in vitro model for validating biomarkers of chemoresistance in breast and colorectal cancer. Our investigation highlights the crucial role of inflammation-related pathways in modulating the response to chemotherapy. Using in silico approaches, we identified specific markers elevated in responders versus non-responders patients. These markers were consistently higher in three-dimensional (3D) tumorsphere models compared to traditional adherent cell culture models. Furthermore, the number of tumorspheres from breast and colorectal cancer cells increased in response to cisplatin and oxaliplatin treatment, respectively, whereas cell viability decreased in adherent cell culture. This differential response underscores the importance of the 3D tumorsphere model in mimicking the tumor microenvironment more accurately than adherent cell culture. The enhanced chemoresistance observed in the 3D tumorspheres model and their correlation of data with the in silico study suggest that 3D culture models are a better option to approach the in vivo model and also to validate in silico data. Our findings indicate that tumorspheres are an ideal model for validating chemoresistance biomarkers and exploring the interplay between inflammation and chemoresistance in breast and colon cancer. Full article
(This article belongs to the Special Issue Cancer and Signalling: Targeting Cellular Pathways)
Show Figures

Figure 1

Figure 1
<p>Identification of signaling pathways alterations in responders vs. non-responders breast (<b>A</b>) and colorectal (<b>B</b>) cancer patients. Pathways with FDR value ≤ 0.05 are represented. Enrichment plots of matched KEGG pathways are represented, respectively.</p>
Full article ">Figure 2
<p>mRNA expression levels of inflammation-related markers in MCF7 (<b>A</b>) and SW620 (<b>B</b>) tumorspheres compared to a two-dimensional (2D) culture. Statistical significance was analyzed by Student’s <span class="html-italic">t</span>-test and set at * <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 3
<p>Gene expression of inflammation markers in breast cancer (<b>A</b>) and colorectal cancer (<b>B</b>) patients according to response to chemotherapy treatment. Statistical significance was analyzed by Student’s <span class="html-italic">t</span>-test and set at <span class="html-italic">p</span> ≤ 0.05 (highlighted values).</p>
Full article ">Figure 4
<p>Viability and tumorsphere formation efficiency of MCF7 (<b>A</b>–<b>D</b>) and SW620 (<b>E</b>–<b>H</b>) cell lines after Cisplatin/Oxaliplatin treatment, respectively. Statistical significance compared to 0 µM (Cisplatin or Oxaliplatin) was analyzed by Student’s <span class="html-italic">t</span>-test and set at * <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">
24 pages, 2351 KiB  
Review
The Drug Transporter P-Glycoprotein and Its Impact on Ceramide Metabolism—An Unconventional Ally in Cancer Treatment
by Johnson Ung, Miki Kassai, Su-Fern Tan, Thomas P. Loughran, David J. Feith and Myles C. Cabot
Int. J. Mol. Sci. 2024, 25(18), 9825; https://doi.org/10.3390/ijms25189825 - 11 Sep 2024
Viewed by 663
Abstract
The tumor-suppressor sphingolipid ceramide is recognized as a key participant in the cytotoxic mechanism of action of many types of chemotherapy drugs, including anthracyclines, Vinca alkaloids, the podophyllotoxin etoposide, taxanes, and the platinum drug oxaliplatin. These drugs can activate de novo synthesis of [...] Read more.
The tumor-suppressor sphingolipid ceramide is recognized as a key participant in the cytotoxic mechanism of action of many types of chemotherapy drugs, including anthracyclines, Vinca alkaloids, the podophyllotoxin etoposide, taxanes, and the platinum drug oxaliplatin. These drugs can activate de novo synthesis of ceramide or stimulate the production of ceramide via sphingomyelinases to limit cancer cell survival. On the contrary, dysfunctional sphingolipid metabolism, a prominent factor in cancer survival and therapy resistance, blunts the anticancer properties of ceramide-orchestrated cell death pathways, especially apoptosis. Although P-glycoprotein (P-gp) is famous for its role in chemotherapy resistance, herein, we propose alternate interpretations and discuss the capacity of this multidrug transporter as a “ceramide neutralizer”, an unwelcome event, highlighting yet another facet of P-gp’s versatility in drug resistance. We introduce sphingolipid metabolism and its dysfunctional regulation in cancer, present a summary of factors that contribute to chemotherapy resistance, explain how P-gp “neutralizes” ceramide by hastening its glycosylation, and consider therapeutic applications of the P-gp-ceramide connection in the treatment of cancer. Full article
Show Figures

Figure 1

Figure 1
<p>Ceramide metabolism and metabolism of higher sphingolipids. Ceramide is synthesized in the endoplasmic reticulum. The de novo biosynthesis starts with the condensation of serine and palmitoyl-CoA, catalyzed by serine palmitoyltransferase. The product, 3-ketosphinganine, contains 18 carbons and is reduced to sphinganine by 3-ketosphinganine reductase. The next step generates the saturated ceramide precursor, dihydroceramide, via the action of ceramide synthases, of which there are six isoforms (CerS1-6) that ultimately give rise to a multitude of molecular species with distinct roles. Finally, although dihydroceramide is nearly identical in structure to ceramide, it lacks the 4,5-<span class="html-italic">trans</span> double bond, which is inserted by dihydroceramide desaturase to form ceramide. Ceramide can also be produced by the action of specialized phospholipases, known as sphingomyelinases. Sphingomyelinases, which are characterized according to their pH optimum and subcellular locations, cleave sphingomyelin at the phosphodiester bond that is proximal to ceramide, producing ceramide and choline phosphate. Ceramide can also be formed by the action of ceramide-1-phosphate (C1-P) phosphatase and by glucosylcerebrosidase. Once produced, ceramide can be hydrolyzed by ceramidases, glycosylated by glucosylceramide synthase (GCS), used to generate sphingomyelin by sphingomyelin synthases, or phosphorylated by ceramide kinase producing C1-P. Strategic points in de novo synthesis and in subsequent ceramide metabolism can be activated or inhibited, providing useful avenues to define ceramide-regulated events, such as cell fate. Enzyme inhibitors and P-glycoprotein (P-gp) antagonists are often used to amplify the induction of cell death by ceramide. Enzymes that deplete intracellular ceramide either by catabolism (e.g., ceramidase activity) or anabolism (e.g., glycosylation, sphingomyelin synthesis, phosphorylation) can contribute to cancer cell growth.</p>
Full article ">Figure 2
<p>Factors contributing to chemotherapy resistance. The inclusion of dysfunctional sphingolipid metabolism marks an important biology that contributes to chemotherapy resistance. As noted in the text, modifications in sphingolipid metabolism work in concert with other contributors, such as the ABC transporter, P-gp, which directs chemotherapy efflux and also participates in ceramide metabolism. P-gp, P-glycoprotein. Created with BioRender.com.</p>
Full article ">Figure 3
<p>The impact of chemotherapy selection pressure on sphingolipid metabolism. Cancer cells grown to acquire resistance to anticancer agents display prominent elevations in the expression and activity of some of the key enzymes in sphingolipid metabolism, including GCS, AC, and SPHK1 as indicated by the upward arrows. Through glycosylation, GCS converts proapoptotic ceramide to GlcCer, a detoxified product. Likewise, AC hydrolyzes ceramide to disable its cytotoxic impact. SPHK1 utilizes sphingosine, a product of AC, to produce S1-P and elicit mitogenic responses. GCS, glucosylceramide synthase; AC, acid ceramidase; SPHK1, sphingosine kinase 1; So, sphingosine.</p>
Full article ">Figure 4
<p>P-glycoprotein antagonists enhance efficacy of anticancer agents that employ ceramides in mechanism of action. (<b>A</b>) Anticancer drugs (e.g., CNL, daunorubicin, etoposide, cytarabine, fenretinide, imatinib) can increase intracellular ceramide levels via de novo synthesis, ceramide synthase activation, and sphingomyelinase activation. (<b>B</b>) The resulting ceramide deluge converges on mitochondria to elicit the intrinsic pathway of apoptosis (pro-death). Upregulated activities of (<b>C</b>) GCS and (<b>D</b>) P-gp, characteristic in multidrug resistance, hijack this ceramide-governed cell death pathway by hyper-conversion of ceramide to GlcCer. (<b>E</b>) P-gp antagonists block drug efflux to circumvent efflux-mediated resistance mechanisms. (<b>F</b>) GCS inhibitors (e.g., NB-DNJ) and (<b>G</b>) P-gp inhibitors (e.g., zosuquidar, tamoxifen, tariquidar, elacridar) can block ceramide glycosylation directly at GCS or indirectly at Golgi-resident P-gp by blocking GlcCer entry into the Golgi, promoting buildup of GlcCer and product inhibition of GCS. (<b>H</b>) The inhibitory symbol from GlcCer to GCS indicates product inhibition. GCS, glucosylceramide synthase; P-gp, P-glycoprotein; GlcCer, glucosylceramide; NB-DNJ, 1N-Butyldeoxynojirimycin; LacCer, lactosylceramide; CNL, ceramide nanoliposome. Created with BioRender.com.</p>
Full article ">
50 pages, 19245 KiB  
Review
Liposomal Formulations of Metallodrugs for Cancer Therapy
by Eleonora Botter, Isabella Caligiuri, Flavio Rizzolio, Fabiano Visentin and Thomas Scattolin
Int. J. Mol. Sci. 2024, 25(17), 9337; https://doi.org/10.3390/ijms25179337 - 28 Aug 2024
Viewed by 537
Abstract
The search for new antineoplastic agents is imperative, as cancer remains one of the most preeminent causes of death worldwide. Since the discovery of the therapeutic potential of cisplatin, the study of metallodrugs in cancer chemotherapy acquired increasing interest. Starting from cisplatin derivatives, [...] Read more.
The search for new antineoplastic agents is imperative, as cancer remains one of the most preeminent causes of death worldwide. Since the discovery of the therapeutic potential of cisplatin, the study of metallodrugs in cancer chemotherapy acquired increasing interest. Starting from cisplatin derivatives, such as oxaliplatin and carboplatin, in the last years, different compounds were explored, employing different metal centers such as iron, ruthenium, gold, and palladium. Nonetheless, metallodrugs face several drawbacks, such as low water solubility, rapid clearance, and possible side toxicity. Encapsulation has emerged as a promising strategy to overcome these issues, providing both improved biocompatibility and protection of the payload from possible degradation in the biological environment. In this respect, liposomes, which are spherical vesicles characterized by an aqueous core surrounded by lipid bilayers, have proven to be ideal candidates due to their versatility. In fact, they can encapsulate both hydrophilic and hydrophobic drugs, are biocompatible, and their properties can be tuned to improve the selective delivery to tumour sites exploiting both passive and active targeting. In this review, we report the most recent findings on liposomal formulations of metallodrugs, with a focus on encapsulation techniques and the obtained biological results. Full article
Show Figures

Figure 1

Figure 1
<p>Mechanism of lipoplatin internalization in cancer cells and subsequent cisplatin release. Reprinted with permission from Ref. [<a href="#B47-ijms-25-09337" class="html-bibr">47</a>]. 2024, Wiley.</p>
Full article ">Figure 2
<p>Preparation of cisplatin-loaded PLGA–Avastin<sup>®</sup> conjugated liposomes.</p>
Full article ">Figure 3
<p>Synthesis of <b>Pt-4</b> and representation of the liposomal formulation for its encapsulation.</p>
Full article ">Figure 4
<p>Representation of liposomes containing (<b>Pt-6</b>) prodrug.</p>
Full article ">Figure 5
<p>Synthesis of <b>Pt-7</b> and its use in the formulation with aNLG919.</p>
Full article ">Figure 6
<p>Formulation of (<b>Pt-7</b>) with metformin.</p>
Full article ">Figure 7
<p>Representation of the most popular ruthenium complexes tested for cancer therapy.</p>
Full article ">Figure 8
<p>Representation of (<b>A</b>) [Ru(NO)(bpy)(4-pic)](PF<sub>6</sub>)<sub>3</sub> (<b>Ru-1</b>), and (<b>B</b>) [Ru(phen)<sub>2</sub>(dppz)](ClO<sub>4</sub>)<sub>2</sub> (<b>Ru-2</b>).</p>
Full article ">Figure 9
<p>Composition of Ru3Lipo and proposed triggered cell death pathways.</p>
Full article ">Figure 10
<p>Representation of <b>Ru-5</b> and <b>Ru-6</b>.</p>
Full article ">Figure 11
<p>Representation of the ruthenium complexes synthesized by Fandzloch and Jaromin.</p>
Full article ">Figure 12
<p>Representation of (<b>A</b>) <b>Ru-8</b> and (<b>B</b>) the drug decitabine.</p>
Full article ">Figure 13
<p>Encapsulation of <b>Ru-9</b> in PDA-supported liposomes.</p>
Full article ">Figure 14
<p>Liposomes locked-in dendrimers obtained through hydrophobic and hydrophilic loading.</p>
Full article ">Figure 15
<p>Representation of the nucleolipid-based ruthenium complexes named ToThyRu, HoThyRu, DoHuRu, HoThyDansRu, and HoUrRu.</p>
Full article ">Figure 16
<p>Schematic synthesis of nucleolipids and their coordination to ruthenium.</p>
Full article ">Figure 17
<p>Liposomal encapsulation of the Ru(II) complex coordinating curcumin reported by Hong and Kim.</p>
Full article ">Figure 18
<p>Iridium complexes: <b>Ir-20</b>, <b>Ir-21</b>, and <b>Ir-22</b>.</p>
Full article ">Figure 19
<p>Iridium complexes <b>Ir-23</b> and <b>Ir-24</b>.</p>
Full article ">Figure 20
<p>Iridium complexes synthesized by Patra and Patra.</p>
Full article ">Figure 21
<p>Iridium complexes synthesized by Komarnicka and co-workers.</p>
Full article ">Figure 22
<p>Representation of CuPhen and its liposomal formulation prepared by Casini, Soveral, and Gaspar.</p>
Full article ">Figure 23
<p>Representation of (<b>A</b>) the ligand TPZ and (<b>B</b>) the complex <b>Cu-3</b>.</p>
Full article ">Figure 24
<p>Representation of (<b>A</b>) quercetin and (<b>B</b>) curcumin.</p>
Full article ">Figure 25
<p>Representation of (<b>A</b>) CX5461 and (<b>B</b>) Diethyldithiocarbamate.</p>
Full article ">Figure 26
<p>Preparation of the liposomal formulation prepared by Zhang and colleagues to encapsulate Cu(DDC)<sub>2</sub>. Reprinted with permission from Ref. [<a href="#B114-ijms-25-09337" class="html-bibr">114</a>]. 2024, Elsevier.</p>
Full article ">Figure 27
<p>Representation of Triapine, COTI-2, and their copper(II) complexes (<b>Cu-5</b> and <b>Cu-6</b>), together with the techniques used for their encapsulation.</p>
Full article ">Figure 28
<p>Ferrocenyl compounds encapsulated into lipid nanocapsules.</p>
Full article ">Figure 29
<p>Iron(III) complexes bearing trianionic aminobisphenolate ligands.</p>
Full article ">Figure 30
<p>Sn(IV) complex studied by Fernandes and Baptista.</p>
Full article ">Figure 31
<p>Oxovanadium(IV)–curcumin–bipyridine complex (<b>V-1</b>) and its liposomal formulation.</p>
Full article ">Figure 32
<p>Zinc(II) complexes studied by Correia and Gaspar.</p>
Full article ">
12 pages, 52016 KiB  
Article
Changes in AmotL2 Expression in Cells of the Human Enteral Nervous System in Oxaliplatin-Induced Enteric Neuropathy
by Rebeca González-Fernández, Rita Martín-Ramírez, María-del-Carmen Maeso, Alberto Lázaro, Julio Ávila, Pablo Martín-Vasallo and Manuel Morales
Biomedicines 2024, 12(9), 1952; https://doi.org/10.3390/biomedicines12091952 - 26 Aug 2024
Viewed by 3204
Abstract
Gastrointestinal (GI) toxicity is a common side effect in patients undergoing oxaliplatin (OxPt)-based chemotherapy for colorectal cancer (CRC). Frequently, this complication persists in the long term and could affect the efficacy of the treatment and the patient’s life quality. This long-term GI toxicity [...] Read more.
Gastrointestinal (GI) toxicity is a common side effect in patients undergoing oxaliplatin (OxPt)-based chemotherapy for colorectal cancer (CRC). Frequently, this complication persists in the long term and could affect the efficacy of the treatment and the patient’s life quality. This long-term GI toxicity is thought to be related to OxPt-induced enteral neuropathy. AmotL2 is a member of the Angiomotin family of proteins, which play a role in cell survival, neurite outgrowth, synaptic maturation, oxidative stress protection, and inflammation. In order to assess the role of AmotL2 in OxPt-induced enteral neuropathy, we studied the expression of AmotL2 in cells of the enteric nervous system (ENS) of untreated and OxPt-treated CRC patients and its relationship with inflammation, using immunofluorescence confocal microscopy. Our results in human samples show that the total number of neurons and glial cells decreased in OxPt-treated patients, and TNF-α and AmotL2 expression was increased and colocalized in both neurons and glia. AmotL2 differential expression between OxPt-treated and untreated CRC patients shows the involvement of this scaffold protein in the inflammatory component and toxicity by OxPt in the ENS. Full article
(This article belongs to the Special Issue Diabetes and Enteric Nervous System)
Show Figures

Figure 1

Figure 1
<p>Double staining for MAP2 (red) and GFAP (cyan) in human enteral nervous system. Panels (<b>A</b>–<b>C</b>): MAP2+ cells (white arrowheads) and GFAP+ cells (yellow arrowheads) in control with healthy colon. Panels (<b>D</b>–<b>F</b>): healthy colon areas from not-yet-treated CRC patients; most of the cells show both MAP2+ and GFAP+ staining (green arrowheads) and a few MAP2+ cells and GFAP− cells (white arrowheads). Panels (<b>G</b>–<b>I</b>): samples from OxPt-treated CRC patients; MAP2+ and GFAP+ staining (green arrowheads), a few MAP2+/GFAP− cells (white arrowheads) or GFAP+/MAP2- cells (yellow arrowheads). Bar = 20 µm.</p>
Full article ">Figure 2
<p>Oxaliplatin treatment induced a reduction in MAP2- and GFAP-immunoreactive cells within the colon myenteric plexus. Wholemount preparations of the samples containing colon myenteric plexus were labeled with MAP2 and GFAP. Oxaliplatin treatment induced a significant reduction in the density of MAP2- and GFAP-immunoreactive cells within the myenteric plexus when compared with samples obtained previous to OxPt-treated group. Untreated = samples from patients previous to chemotherapy. OxPt = samples from patients after chemotherapy. The data are expressed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 5/group.</p>
Full article ">Figure 3
<p>Staining for TNF-α (green), MAP2 (red), and GFAP (cyan) in human non-tumor (healthy) colon from untreated CRC patients. Panels (<b>A</b>–<b>C</b>): TNF-α staining, MAP2+ (white arrow-heads), and GFAP+ (yellow arrowheads) of ENS surrounding colon crypts. Panels (<b>D</b>–<b>F</b>): DAPI, DAPI/TNF-α staining, and merge of DAPI/TNF-α/MAP2/GFAP. Specific fluorescence for TNF-α was positive only in a few GFAP+ cells (<b>C</b>). Bar = 20 µm.</p>
Full article ">Figure 4
<p>Staining for TNF-α (green), MAP2 (red), and GFAP (cyan) in human non-tumor (healthy) colon from OxPt-treated CRC patients. Panels (<b>A</b>–<b>C</b>): TNFα+/MAP2+ cells (white arrows) and TNFα+/GFAP+ cells (yellow arrows). Panels (<b>D</b>–<b>F</b>): DAPI, DAPI/TNF-α staining, and merge of DAPI/TNF-α/MAP2/GFAP. Bar = 20 µm.</p>
Full article ">Figure 5
<p>Double staining of AmotL2 (green) and TNF-α (cyan) in non-tumor colon areas from untreated and OxPt-treated CRC patients. Panels (<b>A</b>–<b>D</b>): AmotL2+/TNF-α colocalized staining at variable intensities, medium (white arrowheads) or low (yellow arrowheads); AmotL2+/TNF-α/DAPI merge. Panels (<b>E</b>–<b>H</b>): AmotL2+ at high staining level (yellow arrowheads) and TNF-α+ colocalized in some cells (white arrowheads). TNF-α signal intensity varied parallel to that of AmotL2. Panels D and H show AmotL2/TNF-α/DAPI merges. Bar = 20 µm.</p>
Full article ">Figure 6
<p>Double staining of AmotL2 (green) and MAP2 (red) in non-tumor colon areas from untreated and OxPt-treated CRC patients. Panels (<b>A</b>–<b>D</b>): AmotL2+ fluorescence staining at low/medium intensity in MAP2+ cells (white arrowheads) and AmotL2+ fluorescence staining at medium intensity in MAP2- cells (yellow arrowheads). Panels (<b>E</b>–<b>H</b>): high AmotL2 staining intensity in MAP2+ cells (white arrowheads). Bar = 20 µm.</p>
Full article ">Figure 7
<p>Double staining of AmotL2 (green) and GFAP (red) in non-tumor colon areas from untreated and OxPt-treated CRC patients. Panels (<b>A</b>–<b>D</b>): low AmotL2+ fluorescence signal in GFAP+ cells (white arrowheads). Some AmotL2+/GFAP− cells (yellow arrowheads). Panels (<b>E</b>–<b>H</b>): AmotL2+ staining in GFAP+ cells (white arrowheads). Bar = 20 µm.</p>
Full article ">Figure 8
<p>Specific fluorescence signal for AmotL2, GFAP, and TNF-α in neurons and glia from untreated and OxPt-treated CRC patients. Untreated = samples from patients previous to chemotherapy. OxPt = samples from patients after chemotherapy. The data are expressed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 4738 KiB  
Article
A Benzimidazole-Based N-Heterocyclic Carbene Derivative Exhibits Potent Antiproliferative and Apoptotic Effects against Colorectal Cancer
by Sarah Al-Nasser, Maha Hamadien Abdulla, Noura Alhassan, Mansoor-Ali Vaali-Mohammed, Suliman Al-Omar, Naceur Hamdi, Yasser Elnakady, Sabine Matou-Nasri and Lamjed Mansour
Medicina 2024, 60(9), 1379; https://doi.org/10.3390/medicina60091379 - 23 Aug 2024
Viewed by 692
Abstract
Background and Objectives: Colorectal cancer (CRC) remains a major global health issue. Although chemotherapy is the first-line treatment, its effectiveness is limited due to drug resistance developed in CRC. To overcome resistance and improve the prognosis of CRC patients, investigating new therapeutic [...] Read more.
Background and Objectives: Colorectal cancer (CRC) remains a major global health issue. Although chemotherapy is the first-line treatment, its effectiveness is limited due to drug resistance developed in CRC. To overcome resistance and improve the prognosis of CRC patients, investigating new therapeutic approaches is necessary. Materials and Methods: Using human colorectal adenocarcinoma (HT29) and metastatic CRC (SW620) cell lines, the potential anticancer properties of a newly synthesized compound 1-(Isobutyl)-3-(4-methylbenzyl) benzimidazolium chloride (IMBZC) were evaluated by performing MTT cytotoxicity, cell migration, and colony formation assays, as well as by monitoring apoptosis-related protein and gene expression using Western blot and reverse transcription–quantitative polymerase chain reaction technologies. Results: Tested at various concentrations, the half-maximal inhibitory concentrations (IC50) of IMBZC on HT29 and SW620 cell growth were determined to be 22.13 µM (6.97 μg/mL) and 15.53 µM (4.89 μg/mL), respectively. IMBZC did not alter the cell growth of normal HEK293 cell lines. In addition, IMBZC inhibited cell migration and significantly decreased colony formation, suggesting its promising role in suppressing cancer metastasis. Mechanistic analyses revealed that IMBZC treatment increased the expression of pro-apoptotic proteins p53 and Bax, while decreasing the expression of anti-apoptotic proteins Bcl-2 and Bcl-xL, thus indicating the induction of apoptosis in IMBZC-treated CRC cells, compared to untreated cells. Additionally, the addition of IMBZC to conventional chemotherapeutic drugs (i.e., 5-fluorouracil, irinotecan, and oxaliplatin) resulted in an increase in the cytotoxic potential of the drugs. Conclusions: This study suggests that IMBZC has substantial anticancer effects against CRC cells through its ability to induce apoptosis, inhibit cancer cell migration and colony formation, and enhance the cytotoxic effects of conventional chemotherapeutic drugs. These findings indicate that IMBZC could be a promising chemotherapeutic drug for the treatment of CRC. Further research should be conducted using in vivo models to confirm the anti-CRC activities of IMBZC. Full article
(This article belongs to the Section Gastroenterology & Hepatology)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of 1-(Isobutyl)-3-(4-methylbenzyl) benzimidazolium chloride (IMBZC) (MW = 314.85 g/mol).</p>
Full article ">Figure 2
<p>Assessment of the cytotoxic potential of compound IMBZC on colon adenocarcinoma HT29 and mCRC SW620 cells, in comparison with normal HEK293 cells, using MTT assay. Determination of percent viability of (<b>A</b>) HT29, (<b>B</b>) SW620, and (<b>C</b>) HEK923 cells after 24 h of incubation in the absence (i.e., control) or presence of increasing concentrations of (3.97 to 31.76 µM) IMBZC. * <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 vs. control. (ns: not significant).</p>
Full article ">Figure 3
<p>Evaluation of the antiproliferative potential of IMBZC on CRC cell migration using wound healing assay. (<b>A</b>) HT29 and (<b>B</b>) SW620 cells were seeded in 6-well plates and incubated with complete medium until confluence, and then the cell monolayer was scratched with a sterile 200 µL tip and washed with PBS. The medium was replaced with or without IMBZC then incubated for 24 h. Microscopy was used to examine the cells, and digital images were taken.</p>
Full article ">Figure 4
<p>The compound IMBZC inhibits HT29 and SW620 cell-based colony formation. Both HT29 (<b>A</b>) and SW620 (<b>B</b>) cells were incubated for 10–12 days at 37 °C for colony formation, along with untreated HT29 (<b>A</b>.<b>A</b>) and SW620 (<b>B</b>.<b>A</b>) cells (i.e., control). The number of colonies is represented by a bar graph and the data are presented as mean ± SD (N = 3). ** <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; 0.0001 vs. control.</p>
Full article ">Figure 5
<p>Expression of mRNA levels of anti-apoptotic <span class="html-italic">BCL-2</span> and <span class="html-italic">BCL-xL</span> (<b>A</b>,<b>B</b>) and pro-apoptotic <span class="html-italic">BAX</span> and <span class="html-italic">TP53</span> (<b>C</b>,<b>D</b>) genes monitored in colorectal adenocarcinoma HT29 (<b>A</b>,<b>C</b>) and mCRC SW620 (<b>B</b>,<b>D</b>) (cell lines using RT-qPCR. Bar graphs show the relative gene expression levels calculated as a ratio to <span class="html-italic">GAPDH</span>, the internal control, and data are presented as mean ± SD (N = 3) * <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 vs. control.</p>
Full article ">Figure 6
<p>The effect of IMBZC on the anti-apoptotic Bcl-2 and Bcl-xl and pro-apoptotic Bax and p53 protein expression levels. (<b>A</b>) HT29 and (<b>B</b>) SW620 cells were treated for 24 h at different concentrations (7.94, 15.88 and 31.76 µM) of IMBZC. Anti-Bcl-2, Bcl-xl, p53, and Bax antibodies were used to target these proteins in whole cell lysates. The strength of protein bands was semi-quantified relative to β-actin, used as a loading control, and was presented as relative to protein expression, compared to the untreated control. The bar graphs show the mean ± SD of three independent experiments. * <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; 0.0001 vs. control.</p>
Full article ">Figure 7
<p>IMBZC potentiates the cytotoxicity of CRC conventional drugs (i.e., 5-FU, IRI and OXA) in (<b>A</b>) HT29 and (<b>B</b>) SW620 cells. 5-FU, IRI, and OXA drugs were tested at different concentrations (2.5, 5 and 10 µM) for 24 h of incubation. Combinations of conventional drugs with the IMBZC compound on HT29 (<b>C</b>) and SW620 (<b>D</b>) cell lines. The cell viability percentage was determined using MTT assay. The bar graphs show the mean ± SD of three independent experiments. * <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; 0.0001 vs. control.</p>
Full article ">
14 pages, 20836 KiB  
Article
Identification of Clinical Value and Biological Effects of XIRP2 Mutation in Hepatocellular Carcinoma
by Dahuan Li, Xin Bao, Shan Lei, Wenpeng Cao, Zhirui Zeng and Tengxiang Chen
Biology 2024, 13(8), 633; https://doi.org/10.3390/biology13080633 - 19 Aug 2024
Viewed by 768
Abstract
Hepatocellular carcinoma (HCC) is a prevalent malignant digestive tumor. Numerous genetic mutations have been documented in HCC, yet the clinical significance of these mutations remains largely unexplored. The objective of this study is to ascertain the clinical value and biological effects of xin [...] Read more.
Hepatocellular carcinoma (HCC) is a prevalent malignant digestive tumor. Numerous genetic mutations have been documented in HCC, yet the clinical significance of these mutations remains largely unexplored. The objective of this study is to ascertain the clinical value and biological effects of xin actin binding repeat containing 2 (XIRP2) mutation in HCC. The gene mutation landscape of HCC was examined using data from the Cancer Genome Atlas and the International Cancer Genome Consortium databases. The prognostic significance of the XIRP2 mutation was assessed through KM plot analysis. The association between drug sensitivity and the XIRP2 mutation was investigated using the TIDE algorithm and CCK-8 experiments. The biological effects of the XIRP2 mutation were evaluated through qRT-PCR, protein stability experiments, and relevant biological experiments. The XIRP2 mutation is one of the high-frequency mutations in HCC, and is associated with poor prognosis. A total of 72 differentially expressed genes (DEGs) were observed in HCC tissues with the XIRP2 mutation as compared to those with the XIRP2 wildtype, and these DEGs were closely related to ion metabolic processes. The XIRP2 mutation was linked to alterations in the sensitivity of fludarabine, oxaliplatin, WEHI-539, and LCL-161. CCK-8 assays demonstrated that HCC cells carrying the XIRP2 mutation exhibited increased resistance to fludarabine and oxaliplatin, but enhanced sensitivity to WEHI-539 and LCL-161 as compared with those HCC cells with the XIRP2 wildtype. The XIRP2 mutation was found to have no impact on the mRNA levels of XIRP2 in tissues and cells, but it did enhance the stability of the XIRP2 protein. Mechanically, the inhibition of XIRP2 resulted in a significant increase in sensitivity to oxaliplatin through an elevation in zinc ions and a calcium ion overload. In conclusion, the XIRP2 mutation holds potential as a biomarker for predicting the prognosis and drug sensitivity of HCC and serves as a therapeutic target to enhance the efficacy of oxaliplatin. Full article
Show Figures

Figure 1

Figure 1
<p>Analysis of mutations in HCC tissues: (<b>A</b>) Screening of the top 20 mutation genes from 368 HCC tissues in the TCGA. (<b>B</b>) Screening of the top 20 mutation genes from 228 HCC tissues in the ICGC. (<b>C</b>) Exploration of the overlapping gene mutations in both the TCGA and ICGC. (<b>D</b>) The relationship of the overlapping gene mutations and the overall survival times of HCC patients in the TCGA. (<b>E</b>) The relationship of the overlapping gene mutations and the overall survival times of HCC patients in the ICGC. (<b>F</b>) Location of the mutation site of the <span class="html-italic">XIRP2</span> mutation. aa, amino acid. (<b>G</b>) More HCC patients with the <span class="html-italic">XIRP2</span> mutation had high fibrosis Ishak scores. (<b>H</b>) The <span class="html-italic">XIRP2</span> mutation can act as an independent factor for HCC patient survival. * represents <span class="html-italic">p</span> &lt; 0.05; ** represents <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Screening of differentially expressed genes (DEGs) between the <span class="html-italic">XIRP2</span> mutation and <span class="html-italic">XIRP2</span> wildtype in HCC and the enrichment analysis: (<b>A</b>) Volcano plot showing changes of genes between the <span class="html-italic">XIRP2</span> mutation and the <span class="html-italic">XIRP2</span> wildtype in the TCGA. (<b>B</b>) Heatmap showing the DEGs between the <span class="html-italic">XIRP2</span> mutation and the <span class="html-italic">XIRP2</span> wildtype in the TCGA. (<b>C</b>) Volcano plot showing changes of genes between the <span class="html-italic">XIRP2</span> mutation and the <span class="html-italic">XIRP2</span> wildtype in the ICGC. (<b>D</b>) Heatmap showing changes in DEGs between the <span class="html-italic">XIRP2</span> mutation and the <span class="html-italic">XIRP2</span> wildtype in the ICGC. (<b>E</b>) Intersection analysis demonstrated that 33 genes were upregulated in the <span class="html-italic">XIRP2</span> mutation and the <span class="html-italic">XIRP2</span> wildtype in both the TCGA and ICGC. (<b>F</b>) Intersection analysis demonstrated that 39 genes were downregulated in the <span class="html-italic">XIRP2</span> mutation and the <span class="html-italic">XIRP2</span> wildtype in both the TCGA and ICGC. (<b>G</b>) Enrichment analysis of the biological process terms of the 33 overlapping upregulated genes. (<b>H</b>) Enrichment analysis of the biological process terms of the 39 overlapping downregulated genes.</p>
Full article ">Figure 3
<p><span class="html-italic">XIRP2</span> mutation increased the resistance of HCC cells to fludarabine and oxaliplatin but increased their sensitivity to WEHI-539 and LCL 161. (<b>A</b>) The OncoPredict algorithm was used to analyze the change of drug score of 198 drugs between the HCC tissues with the <span class="html-italic">XIRP2</span> wildtype and the <span class="html-italic">XIRP2</span> mutation. (<b>B</b>) Drug score of fludarabine between the HCC tissues with the <span class="html-italic">XIRP2</span> wildtype and the <span class="html-italic">XIRP2</span> mutation. (<b>C</b>) Drug score of oxaliplatin between the HCC tissues with the <span class="html-italic">XIRP2</span> wildtype and the <span class="html-italic">XIRP2</span> mutation. (<b>D</b>) Drug score of WEHI-539 between the HCC tissues with the <span class="html-italic">XIRP2</span> wildtype and the <span class="html-italic">XIRP2</span> mutation. (<b>E</b>) Drug score of LCL-161 between the HCC tissues with the <span class="html-italic">XIRP2</span> wildtype and the <span class="html-italic">XIRP2</span> mutation. SNU475 harbored the <span class="html-italic">XIRP2</span> mutation (I827V), while HepG2, Hep3B, Huh7, Huh1, and SK-Hep1 harbored the <span class="html-italic">XIRP2</span> wildtype. (<b>F</b>) CCK-8 assays were used to detect the IC50 of fludarabine in SNU475, HepG2, Hep3B, Huh7, Huh1, and SK-Hep1 in 48 h. (<b>G</b>) CCK-8 assays was used to detect the IC50 of oxaliplatin in SNU475, HepG2, Hep3B, Huh7, Huh1, and SK-Hep1 in 48 h. (<b>H</b>) CCK-8 assays were used to detect the IC50 of WEHI-539 in SNU475, HepG2, Hep3B, Huh7, Huh1, and SK-Hep1 in 48 h. (<b>I</b>) CCK-8 assays were used to detect the IC50 of LCL-161 in SNU475, HepG2, Hep3B, Huh7, Huh1, and SK-Hep1 in 48 h. * represents <span class="html-italic">p</span> &lt; 0.05; ** represents <span class="html-italic">p</span> &lt; 0.01. <span class="html-italic">n</span> = 3. The control group was used for comparison. Data are shown as mean ± SD.</p>
Full article ">Figure 4
<p>The <span class="html-italic">XIRP2</span> mutation increased the protein stability of XIRP2 protein. (<b>A</b>) Expression of <span class="html-italic">XIRP2</span> mRNA levels in the HCC tissues with the XIRP2 mutation and the XIRP2 wildtype from the TCGA databases. A total of 29 HCC patients with XIRP2-MUT in TCGA cohort (pink dots), while the number of patients with XIRP2-WT was 339 (blue dots). (<b>B</b>) Expression of <span class="html-italic">XIRP2</span> mRNA levels in HCC tissues with the XIRP2 mutation and the XIRP2 wildtype from the ICGC databases. A total of 21 HCC patients with XIRP2-MUT in ICGC cohort (pink dots), while the number of patients with XIRP2-WT was 238 (blue dots). (<b>C</b>) Expression of the <span class="html-italic">XIRP2</span> mRNA levels in SNU475, HepG2, Hep3B, Huh7, Huh1, and SK-Hep1. (<b>D</b>) Expression of XIRP2 protein levels in SNU475, HepG2, Hep3B, Huh7, Huh1, and SK-Hep1 cells. (<b>E</b>,<b>F</b>) Degradation rate of XIRP2 protein in SNU475 and Huh7 cells. * represents <span class="html-italic">p</span> &lt; 0.05; ** represents <span class="html-italic">p</span> &lt; 0.01. <span class="html-italic">n</span> = 3. The control group was used for comparison. Data are shown as mean ± SD. <a href="#app1-biology-13-00633" class="html-app">Figures S1 and S2</a> are Original band for XIRP2 and GAPDH in <a href="#biology-13-00633-f004" class="html-fig">Figure 4</a>D, <a href="#app1-biology-13-00633" class="html-app">Figures S3–S6</a> are Original band for XIRP2 of Huh7, GAPDH of Huh7, XIRP2 of SNU475 and GAPDH of SNU475 in <a href="#biology-13-00633-f004" class="html-fig">Figure 4</a>E.</p>
Full article ">Figure 5
<p>Increased XIRP2 protein in HCC cells with the XIRP2 wildtype reduced their sensitivity to oxaliplatin. (<b>A</b>) XIRP2 plasmids were transfected into Huh7 and Huh1 cells to construct XIRP2 overexpression cells. (<b>B</b>,<b>C</b>) Results of spectrophotometry indicated the levels of zinc ion and calcium ion in Huh7 and Huh1 cells after XIRP2 overexpression and oxaliplatin treatment. (<b>D</b>) Results of immunofluorescence indicated the levels of zinc ion and calcium ion in Huh7 cells after XIRP2 overexpression and oxaliplatin treatment. (<b>E</b>) Cell proliferation in Huh7 and Huh1 cells after XIRP2 overexpression and oxaliplatin treatment. (<b>F</b>) Spherogenesis of Huh7 and Huh1 cells after XIRP2 overexpression and oxaliplatin treatment. (<b>G</b>) TUNEL stain of Huh7 and Huh1 cells after XIRP2 overexpression and oxaliplatin treatment. ** represents <span class="html-italic">p</span> &lt; 0.01. <span class="html-italic">n</span> = 3. The control group was used for comparison. Data are shown as mean ± SD. <a href="#app1-biology-13-00633" class="html-app">Figures S7 and S8</a> are Original band for XIRP2 and GAPDH in <a href="#biology-13-00633-f005" class="html-fig">Figure 5</a>A.</p>
Full article ">Figure 6
<p>Knockdown of XIRP2 in HCC cells with the XIRP2 mutation increased the sensitivity to oxaliplatin via inducing the overload of zinc and calcium ion. (<b>A</b>,<b>B</b>) Targeting XIRP2 siRNAs were transfected into the SNU475 cells to construct XIRP2 knockdown cells, and qRT-PCR and western blotting were used to detect the efficiency of the siRNAs. (<b>C</b>,<b>D</b>) Results of spectrophotometry indicated the levels of zinc ion and calcium ion in the SNU475 cells after XIRP2 knockdown and oxaliplatin treatment. (<b>E</b>) Results of immunofluorescence indicated the levels of zinc ion and calcium ion in the SNU475 cells after XIRP2 knockdown and oxaliplatin treatment. (<b>F</b>) Cell proliferation in the SNU475 cells after XIRP2 knockdown and oxaliplatin treatment or combined with DP-b99 treatment. (<b>G</b>) Spherogenesis of the SNU475 cells after XIRP2 knockdown and oxaliplatin treatment or combined with DP-b99 treatment. (<b>H</b>) TUNEL stain of the SNU475 cells after XIRP2 knockdown and oxaliplatin treatment or combined with DP-b99 treatment. ** represents <span class="html-italic">p</span> &lt; 0.01. <span class="html-italic">n</span> = 3. The control group was used for comparison. Data are shown as mean ± SD. <a href="#app1-biology-13-00633" class="html-app">Figures S9 and S10</a> are Original band for XIRP2 and GAPDH in <a href="#biology-13-00633-f006" class="html-fig">Figure 6</a>B.</p>
Full article ">
13 pages, 552 KiB  
Review
Deciphering the Dilemma: Choosing the Optimal Total Neoadjuvant Treatment Strategy for Locally Advanced Rectal Cancer
by Erik Manriquez, Sebastián Solé, Javiera Silva, Juan Pablo Hermosilla, Rubén Romero and Felipe Quezada-Diaz
Curr. Oncol. 2024, 31(8), 4292-4304; https://doi.org/10.3390/curroncol31080320 - 29 Jul 2024
Viewed by 1048
Abstract
Rectal cancer management has evolved significantly, particularly with neoadjuvant treatment strategies. This narrative review examines the development and effectiveness of these therapies for locally advanced rectal cancer (LARC), highlighting the historical quest that led to current neoadjuvant alternatives. Initially, trials showed the benefits [...] Read more.
Rectal cancer management has evolved significantly, particularly with neoadjuvant treatment strategies. This narrative review examines the development and effectiveness of these therapies for locally advanced rectal cancer (LARC), highlighting the historical quest that led to current neoadjuvant alternatives. Initially, trials showed the benefits of adding radiotherapy (RT) and chemotherapy (CT) to surgery, reducing local recurrence (LR). The addition of oxaliplatin to chemoradiotherapy (CRT) further improved outcomes. TNT integrates chemotherapy and radiotherapy preoperatively to enhance adherence, timing, and systemic control. Key trials, including PRODIGE 23, CAO/ARO/AIO 12, OPRA, RAPIDO, and STELLAR, are analyzed to compare short-course and long-course RT with systemic chemotherapy. The heterogeneity and difficulty in comparing TNT trials due to different designs and outcomes are acknowledged, along with their promising long-term results. On the other hand, it briefly discusses the potential for non-operative management (NOM) in select patients, a strategy gaining traction due to favorable outcomes in specific trials. As a conclusion, this review underscores the complexity of rectal cancer treatment, emphasizing individualized approaches considering patient preferences and healthcare resources. It also highlights the importance of interpreting impressive positive or negative results with caution due to the variability in study designs and patient populations. Full article
(This article belongs to the Section Gastrointestinal Oncology)
Show Figures

Figure 1

Figure 1
<p>Evolution of rectal cancer treatment [<a href="#B8-curroncol-31-00320" class="html-bibr">8</a>,<a href="#B9-curroncol-31-00320" class="html-bibr">9</a>,<a href="#B11-curroncol-31-00320" class="html-bibr">11</a>,<a href="#B12-curroncol-31-00320" class="html-bibr">12</a>,<a href="#B13-curroncol-31-00320" class="html-bibr">13</a>,<a href="#B14-curroncol-31-00320" class="html-bibr">14</a>,<a href="#B15-curroncol-31-00320" class="html-bibr">15</a>,<a href="#B16-curroncol-31-00320" class="html-bibr">16</a>,<a href="#B17-curroncol-31-00320" class="html-bibr">17</a>,<a href="#B18-curroncol-31-00320" class="html-bibr">18</a>,<a href="#B19-curroncol-31-00320" class="html-bibr">19</a>,<a href="#B20-curroncol-31-00320" class="html-bibr">20</a>,<a href="#B21-curroncol-31-00320" class="html-bibr">21</a>,<a href="#B22-curroncol-31-00320" class="html-bibr">22</a>,<a href="#B23-curroncol-31-00320" class="html-bibr">23</a>] CRT: Chemoradiotherapy, CT or Chemo: Chemotherapy, DFS: Disease-free Survival LCRT: Long Course Chemoradiotherapy, SCRT: Short Course Radiotherapy, TME: Total Mesorectal Excision, TNT: Total Neoadjuvant Therapy.</p>
Full article ">
10 pages, 1052 KiB  
Article
Phase IB Study of Oral Selinexor in Combination with Rituximab and Platinum Chemotherapy in Patients with Relapsed/Refractory B-Cell Lymphoma—Final Analysis
by Marie Maerevoet, Olivier Casasnovas, Guillaume Cartron, Franck Morschhauser, Catherine Thieblemont, Kamal Bouabdallah, Pierre Feugier, Vanessa Szablewski, Stephanie Becker and Herve Tilly
Cancers 2024, 16(15), 2672; https://doi.org/10.3390/cancers16152672 - 26 Jul 2024
Viewed by 780
Abstract
Purpose: Selinexor is an oral selective inhibitor of exportine-1 (XPO1) with efficacy as a single agent in heavily pretreated diffuse large B-cell lymphoma (DLBCL). We conducted a study investigating the combination of selinexor with rituximab and platinum-based chemotherapy in B-cell lymphoma. Patients and [...] Read more.
Purpose: Selinexor is an oral selective inhibitor of exportine-1 (XPO1) with efficacy as a single agent in heavily pretreated diffuse large B-cell lymphoma (DLBCL). We conducted a study investigating the combination of selinexor with rituximab and platinum-based chemotherapy in B-cell lymphoma. Patients and methods: We conducted a phase 1b, dose-escalation, and expansion trial, which enrolled patients with relapsed or refractory B-cell non-Hodgkin lymphoma. Patients received oral selinexor according to a 3 + 3 design in combination with rituximab and dexamethasone, high-dose cytarabine, oxaliplatine (DHAOX) or gemcitabine, dexamethasone, and cisplatin (GDP) chemotherapy. Results: A total of 39 patients were enrolled, 27 during the escalation phase and 12 during the expansion phase. Most patients had diffuse large B-cell lymphoma (DLBCL; 77%). Group R-DHAOX was prematurely closed to inclusion due to a recommendation from the French drug agency, independent of this trial. A recommended phase 2 dose (RP2D) of selinexor in association with R-GPD was established at 40 mg on days 1, 8, and 15 of each 21-day cycle. In a population of 18 patients treated at this dose of selinexor, the most frequent grade 3–4 adverse events were hematological. With this regimen, seven obtained a complete metabolic response and five a partial response. The median PFS was 5.8 months. Conclusions: Among the patients with R/R B-cell lymphoma, selinexor at a weekly dose of 40 mg with R-GDP is feasible for outpatients, with a generally acceptable safety profile. Full article
(This article belongs to the Section Cancer Therapy)
Show Figures

Figure 1

Figure 1
<p>Consort diagram: SELINDA study.</p>
Full article ">Figure 2
<p>Outcomes of patients treated at the RP2D of selinexor and R-GDP. (<b>A</b>) Progression-free survival, RP2D cohort. (<b>B</b>) Duration of response, RP2D cohort. (<b>C</b>) Overall survival, RP2D cohort.</p>
Full article ">
14 pages, 3846 KiB  
Article
Activation of p38 and JNK by ROS Contributes to Deoxybouvardin-Mediated Intrinsic Apoptosis in Oxaliplatin-Sensitive and -Resistant Colorectal Cancer Cells
by Si Yeong Seo, Sang Hoon Joo, Seung-On Lee, Goo Yoon, Seung-Sik Cho, Yung Hyun Choi, Jin Woo Park and Jung-Hyun Shim
Antioxidants 2024, 13(7), 866; https://doi.org/10.3390/antiox13070866 - 19 Jul 2024
Cited by 1 | Viewed by 830
Abstract
Colorectal cancer (CRC) remains a global health burden, accounting for almost a million deaths annually. Deoxybouvardin (DB), a non-ribosomal peptide originally isolated from Bouvardia ternifolia, has been reported to possess antitumor activity; however, the detailed mechanisms underlying this anticancer activity have not [...] Read more.
Colorectal cancer (CRC) remains a global health burden, accounting for almost a million deaths annually. Deoxybouvardin (DB), a non-ribosomal peptide originally isolated from Bouvardia ternifolia, has been reported to possess antitumor activity; however, the detailed mechanisms underlying this anticancer activity have not been elucidated. We investigated the anticancer activity of the cyclic hexapeptide, DB, in human CRC HCT116 cells. Cell viability, evaluated by MTT assay, revealed that DB suppressed the growth of both oxaliplatin (Ox)-resistant HCT116 cells (HCT116-OxR) and Ox-sensitive cells in a concentration- and time-dependent manner. Increased reactive oxygen species (ROS) generation was observed in DB-treated CRC cells, and it induced cell cycle arrest at the G2/M phase by regulating p21, p27, cyclin B1, and cdc2 levels. In addition, Western blot analysis revealed that DB activated the phosphorylation of JNK and p38 MAPK in CRC. Furthermore, mitochondrial membrane potential (MMP) was dysregulated by DB, resulting in cytochrome c release and activation of caspases. Taken together, DB exhibited anticancer activity against both Ox-sensitive and Ox-resistant CRC cells by targeting JNK and p38 MAPK, increasing cellular ROS levels, and disrupting MMP. Thus, DB is a potential therapeutic agent for the treatment of Ox-resistant CRC. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Inhibition of growth of CRC cells by DB. (<b>A</b>) Cell viability of CRC cells (HCT116 and HCT116-OxR) and HaCaT treated for 24 (black column) and 48 h (white column) with DB (0, 2, 4, and 6 nM), and Ox (2 µM) as indicated by MTT cell viability assay. Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3). IC<sub>50</sub> values for 48 h incubation. (<b>B</b>–<b>D</b>) Soft agar assay was used to determine the anchorage-independent colony growth in CRC cells (10 days incubation). (<b>B</b>) Micrograph of HCT116 and HCT116-OxR cells at 10 days after treatment. (<b>C</b>,<b>D</b>) colony size and number. * <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 compared to vehicle only.</p>
Full article ">Figure 2
<p>Activation of JNK and p38 MAPK in DB-induced apoptosis. CRC cells HCT116 and HCT116-OxR were analyzed by flow cytometry with annexin V/7-AAD double-staining 48 h after treatment with DB (0, 2, 4, and 6 nM). (<b>A</b>) Flow cytometry plot. (<b>B</b>) Total apoptotic cells. (<b>C</b>) Western blot analysis of cell lysates to detect p-JNK, JNK, p-p38, p38. β-actin was used as the loading control. (<b>D</b>) The ratio of phosphoprotein/total protein signal for JNK and p38. (<b>E</b>,<b>F</b>) Cell viability was assessed by the MTT assay for the CRC cells treated for 48 h with DB, SP600125, SB203580, and Ox. The data are expressed as the mean ± SD from three replicates. * <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 compared with the control group. ### <span class="html-italic">p</span> &lt; 0.001 compared with the DB-alone-treated group.</p>
Full article ">Figure 3
<p>Induction of ROS by DB. CRC cells were treated for 48 h with DB, NAC, and Ox. (<b>A</b>) The cells were analyzed by flow cytometry with Muse<sup>TM</sup> Oxidative Stress Kit. (<b>B</b>) The ratio of ROS-positive cells. (<b>C</b>) Cell viability was assessed using the MTT assay. (<b>D</b>) Western blot analysis to determine the levels of p-JNK, p-p38, and Caspase 3. β-actin was used as the loading control. (<b>E</b>) Graph shows the relative ratio of p-JNK, p-p38, and Casapse-3 over actin in CRC cells treated with DB or NAC. The data are shown as the mean ± standard deviation. (<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, and *** <span class="html-italic">p</span> &lt; 0.001 compared with the control group. # <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 compared with the DB-alone-treated group.</p>
Full article ">Figure 4
<p>Induction of cell cycle arrest at the G2/M phase by DB. CRC cells HCT116 and HCT116-OxR were treated with DB (0, 2, 4, and 6 nM) for 48 h. (<b>A</b>) Flow cytometry analysis with PI staining. (<b>B</b>) The proportion of the cells in the Sub-G1 phase. (<b>C</b>) Cell cycle distribution. Data are shown as mean ± SD from three replicates of three independent experiments. * <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 compared to vehicle only. (<b>D</b>) Western blot analysis of proteins related to cell cycle regulation: p21, p27, cyclin B1, and cdc2. β-actin was used as the loading control. (<b>E</b>) Graph shows the relative level of proteins p21, p27, cyclin B1, and cdc2 in CRC cells treated with DB. The data are shown as the mean ± standard deviation. (<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, and *** <span class="html-italic">p</span> &lt; 0.001 compared with the control group.</p>
Full article ">Figure 5
<p>Dysregulation of mitochondrial membrane by DB. CRC cells HCT116 and HCT116-OxR were treated with DB (0, 2, 4, and 6 nM) for 48 h. (<b>A</b>) Flow cytometry analysis with JC-1 staining. (<b>B</b>) The proportion of the cells with depolarized mitochondrial membrane. Data are shown as mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 compared to vehicle only. (<b>C</b>) Western blot analysis of proteins regulating mitochondrial membrane permeability (Bid, Bax, Bcl-xL, and Bcl-2) and cytochrome c in mitochondrial and cytoplasmic fractions. COX4 was used as the control for mitochondrial fraction, and α-tubulin for cytoplasmic fraction. β-actin was used as the control for proteins from cell lysates. (<b>D</b>) Western blot analysis to determine the level of apoptosis-related proteins Apaf-1, Caspase-3, and PARP. β-actin was used as the control. (<b>E</b>) Graph shows the relative level of proteins Apaf-1, Caspase-3, and PARP in CRC cells treated with DB. The data are shown as the mean ± standard deviation. (<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, and *** <span class="html-italic">p</span> &lt; 0.001 compared with the control group.</p>
Full article ">Figure 6
<p>Caspase activation by DB. CRC cells HCT116 and HCT116-OxR were treated with DB (0, 2, 4, and 6 nM) for 48 h. (<b>A</b>) Flow cytometry analysis with a Muse<sup>®</sup> Multi-caspase Kit. (<b>B</b>) The proportion of the cells with activated caspases was obtained by adding upper right (caspase+ and dead cells) and lower right (caspase+ and live cells). (<b>C</b>) Cell viability assessed by the MTT assay for the CRC cells treated for 48 h with DB, Z-VAD-FMK, and Ox as indicated. Data are presented as mean of three replicates. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 compared to vehicle only. ### <span class="html-italic">p</span> &lt; 0.001 compared with the DB-alone-treated group.</p>
Full article ">
17 pages, 1984 KiB  
Review
Molecular Hydrogen Protects against Various Tissue Injuries from Side Effects of Anticancer Drugs by Reducing Oxidative Stress and Inflammation
by Shin-ichi Hirano and Yoshiyasu Takefuji
Biomedicines 2024, 12(7), 1591; https://doi.org/10.3390/biomedicines12071591 - 17 Jul 2024
Viewed by 1257
Abstract
While drug therapy plays a crucial role in cancer treatment, many anticancer drugs, particularly cytotoxic and molecular-targeted drugs, cause severe side effects, which often limit the dosage of these drugs. Efforts have been made to alleviate these side effects by developing derivatives, analogues, [...] Read more.
While drug therapy plays a crucial role in cancer treatment, many anticancer drugs, particularly cytotoxic and molecular-targeted drugs, cause severe side effects, which often limit the dosage of these drugs. Efforts have been made to alleviate these side effects by developing derivatives, analogues, and liposome formulations of existing anticancer drugs and by combining anticancer drugs with substances that reduce side effects. However, these approaches have not been sufficiently effective in reducing side effects. Molecular hydrogen (H2) has shown promise in this regard. It directly reduces reactive oxygen species, which have very strong oxidative capacity, and indirectly exerts antioxidant, anti-inflammatory, and anti-apoptotic effects by regulating gene expression. Its clinical application in various diseases has been expanded worldwide. Although H2 has been reported to reduce the side effects of anticancer drugs in animal studies and clinical trials, the underlying molecular mechanisms remain unclear. Our comprehensive literature review revealed that H2 protects against tissue injuries induced by cisplatin, oxaliplatin, doxorubicin, bleomycin, and gefitinib. The underlying mechanisms involve reductions in oxidative stress and inflammation. H2 itself exhibits anticancer activity. Therefore, the combination of H2 and anticancer drugs has the potential to reduce the side effects of anticancer drugs and enhance their anticancer activities. This is an exciting prospect for future cancer treatments. Full article
(This article belongs to the Section Molecular and Translational Medicine)
Show Figures

Figure 1

Figure 1
<p>PRISMA flowchart showing the literature search and selection steps.</p>
Full article ">Figure 2
<p>Possible mechanisms underlying protective effects of H<sub>2</sub> against tissue injuries induced by anticancer drugs. Protective effects are mainly categorized as antioxidant, anti-inflammatory, and cell death-regulating effects. AMPK: AMP-dependent protein kinase; Bcl-2: B-cell/CLL lymphoma 2; Bax: Bcl-2-associated x; CAT: catalase; GSH-PX: glutathione peroxidase; H<sub>2</sub>: molecular hydrogen; 4-HNE: 4-hydroxy-2-nonenal; HO-1: heme oxygenase-1; IL: interleukin; 8-iso-PGF2α: 8-iso-prostaglandin F2α; LPS: lipopolysaccharide; LC3: microtubule-associated protein light chain 3; MDA: malondialdehyde; mTOR: mammalian target of rapamycin; Nrf2: nuclear factor erythroid 2-related factor 2; NLRP3: NLR family pyrin domain containing 3; RNS: reactive nitrogen species; ROS: reactive oxygen species; SOD: superoxide dismutase; TNF-α: tumor necrosis factor-α; TLR4: Toll-like receptor 4; TGF-β1: transforming growth factor-β1; TUNEL: terminally deoxynucleotidyl transferase-mediated biotinylated UTP nick end-labeling; ↑: increase; ↓: decrease.</p>
Full article ">Figure 3
<p>Possible mechanism of the anticancer effects of molecular hydrogen (H<sub>2</sub>). H<sub>2</sub> reduces hydroxyl radicals directly, and also exhibits antioxidant, anti-inflammatory, and apoptotic effects via the regulation of gene expression indirectly. Through these effects, H<sub>2</sub> may exhibit anticancer effects. See [<a href="#B27-biomedicines-12-01591" class="html-bibr">27</a>] for the details.↑: increase; ↓: decrease; ↑/↓: increase or decrease.</p>
Full article ">Figure 4
<p>Molecular hydrogen (H<sub>2</sub>) not only has a direct radioprotective effect by reducing hydroxyl radicals, but also indirectly by regulating by gene expression, exhibiting antioxidant, anti-inflammatory and anti-apoptotic effects, which may lead to radioprotective effects. See [<a href="#B25-biomedicines-12-01591" class="html-bibr">25</a>] for the details. ↑: increase; ↓: decrease; ↑/↓: increase or decrease.</p>
Full article ">
55 pages, 15485 KiB  
Article
Anticancer Effect of PtIIPHENSS, PtII5MESS, PtII56MESS and Their Platinum(IV)-Dihydroxy Derivatives against Triple-Negative Breast Cancer and Cisplatin-Resistant Colorectal Cancer
by Maria George Elias, Shadma Fatima, Timothy J. Mann, Shawan Karan, Meena Mikhael, Paul de Souza, Christopher P. Gordon, Kieran F. Scott and Janice R. Aldrich-Wright
Cancers 2024, 16(14), 2544; https://doi.org/10.3390/cancers16142544 - 15 Jul 2024
Cited by 1 | Viewed by 1234
Abstract
Development of resistance to cisplatin, oxaliplatin and carboplatin remains a challenge for their use as chemotherapies, particularly in breast and colorectal cancer. Here, we compare the anticancer effect of novel complexes [Pt(1,10-phenanthroline)(1S,2S-diaminocyclohexane)](NO3)2 (PtIIPHEN [...] Read more.
Development of resistance to cisplatin, oxaliplatin and carboplatin remains a challenge for their use as chemotherapies, particularly in breast and colorectal cancer. Here, we compare the anticancer effect of novel complexes [Pt(1,10-phenanthroline)(1S,2S-diaminocyclohexane)](NO3)2 (PtIIPHENSS), [Pt(5-methyl-1,10-phenanthroline)(1S,2S-diaminocyclohexane)](NO3)2 (PtII5MESS) and [Pt(5,6-dimethyl-1,10-phenanthroline)(1S,2S-diaminocyclohexane)](NO3)2 (PtII56MESS) and their platinum(IV)-dihydroxy derivatives with cisplatin. Complexes are greater than 11-fold more potent than cisplatin in both 2D and 3D cell line cultures with increased selectivity for cancer cells over genetically stable cells. ICP-MS studies showed cellular uptake occurred through an active transport mechanism with considerably altered platinum concentrations found in the cytoskeleton across all complexes after 24 h. Significant reactive oxygen species generation was observed, with reduced mitochondrial membrane potential at 72 h of treatment. Late apoptosis/necrosis was shown by Annexin V-FITC/PI flow cytometry assay, accompanied by increased sub-G0/G1 cells compared with untreated cells. An increase in S and G2+M cells was seen with all complexes. Treatment resulted in significant changes in actin and tubulin staining. Intrinsic and extrinsic apoptosis markers, MAPK/ERK and PI3K/AKT activation markers, together with autophagy markers showed significant activation of these pathways by Western blot. The proteomic profile investigated post-72 h of treatment identified 1597 MDA−MB−231 and 1859 HT29 proteins quantified by mass spectroscopy, with several differentially expressed proteins relative to no treatment. GO enrichment analysis revealed a statistically significant enrichment of RNA/DNA-associated proteins in both the cell lines and specific additional processes for individual drugs. This study shows that these novel agents function as multi-mechanistic chemotherapeutics, offering promising anticancer potential, and thereby supporting further research into their application as cancer therapeutics. Full article
(This article belongs to the Collection Innovations in Cancer Drug Development Research)
Show Figures

Figure 1

Figure 1
<p>Effect of platinum(II) and platinum(IV) complexes on the survival of cancer and normal epithelial cells: (<b>A</b>). MDA−MB−231, (<b>B</b>). MCF−7, (<b>C</b>). HT29, (<b>D</b>). A2780, (<b>E</b>). ADDP, (<b>F</b>). MCF10A. Cells were treated with 3-fold dilutions of the different agents Cisplatin, <b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>, <b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> or <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> starting with a concentration of 150 µM and assayed for cell viability, as described in <a href="#sec2dot3-cancers-16-02544" class="html-sec">Section 2.3</a>. Data points denote mean ± SEM from three independent experiments where samples were run in triplicate.</p>
Full article ">Figure 2
<p>Effect of platinum(II) complexes on the survival of 3D Embedded (<b>A</b>). MDA−MB−231 networks, (<b>B</b>). MDA−MB−231 and (<b>C</b>). HT29 spheroids. Networks and spheroids were treated with 3-fold dilutions of the different agents (Cisplatin, <b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>, <b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>) starting with a concentration of 150 µM and assayed for cell viability, as described in <a href="#sec2dot4-cancers-16-02544" class="html-sec">Section 2.4</a>. Data points denote mean ± SEM from three independent experiments where samples were run in triplicate.</p>
Full article ">Figure 3
<p>Cellular uptake of <b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>, <b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>: ICP-MS analysis of the uptake of platinum in (<b>A</b>). MDA−MB−231 and (<b>B</b>). HT29 cells at 0, 0.5, 1, 3, 6, 12, 24 and 30 h as described in <a href="#sec2dot5-cancers-16-02544" class="html-sec">Section 2.5</a>. <span class="html-italic">n</span> = 3 from three independent experiments where samples were run in triplicate. Data points denote mean <math display="inline"><semantics> <mo>±</mo> </semantics></math> SEM and expressed in nmol/10<sup>6</sup>cells. ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001 in comparison to cisplatin, as measured by two-way ANOVA.</p>
Full article ">Figure 4
<p>Mode of cellular uptake of platinum in MDA−MB−231 and HT29. The intracellular amount of platinum was measured by ICP-MS after incubation at 37 °C or 4 °C, as well as following inhibition of the SLC7A5, transferrin receptor or clathrin-mediated endocytosis as described in <a href="#sec2dot6-cancers-16-02544" class="html-sec">Section 2.6</a>. Data denote mean ± SEM of three independent experiments where samples were run in triplicate. * <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; 0.0001 in comparison to the optimal conditions, as measured by one-way ANOVA.</p>
Full article ">Figure 5
<p>Cellular localisation of platinum in MDA−MB−231 and HT29 post-24 h. The intracellular amount of Pt was measured by ICP-MS after cellular fractionation, as described in <a href="#sec2dot7-cancers-16-02544" class="html-sec">Section 2.7</a>. Data denote mean ± SEM of three independent experiments where samples were run in triplicate. * <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; 0.0001 in comparison to the fractions, as measured by one-way ANOVA.</p>
Full article ">Figure 6
<p>Flow cytometric analysis of cell death mediated by Cisplatin, <b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>, <b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>. MDA−MB−231 and HT29 cells were treated with the complex and analysed at 72 h, as described in <a href="#sec2dot8-cancers-16-02544" class="html-sec">Section 2.8</a>. Representative dot plots: (<b>A</b>). MDA−MB−231, (<b>B</b>). HT29 and Bar graphs (<b>C</b>). MDA−MB−231 and (<b>D</b>). HT29, representing percent viable, apoptotic and necrotic cells. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments where samples were run in triplicate. * <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; 0.0001 compared to control, as measured by one-way ANOVA.</p>
Full article ">Figure 7
<p>Flow cytometric analysis of cell cycle mediated by Cisplatin, <b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>, <b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>. MDA−MB−231 and HT29 cells were treated with IC<sub>30</sub> concentration of each complex and analysed at 72 h, as described in <a href="#sec2dot9-cancers-16-02544" class="html-sec">Section 2.9</a>. Representative histogram plots: (<b>A</b>). MDA−MB−231, (<b>B</b>). HT29 and Bar graphs (<b>C</b>). MDA−MB−231 and (<b>D</b>). HT29, representing percent Sub G1, G0/G1, S and G2+M phases of cell cycle. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments where samples were run in triplicate. * <span class="html-italic">p</span> &lt; 0.05 compared to control, as measured by one-way ANOVA.</p>
Full article ">Figure 8
<p>ROS production upon treatment with platinum(II) and (IV) complexes in MDA−MB−231 and HT29 and at 24, 48 and 72 h, as described in <a href="#sec2dot10-cancers-16-02544" class="html-sec">Section 2.10</a>. <b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>, <b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, Cisplatin and TBHP: t-butyl hydroperoxide. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments where samples were run in triplicate. * <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; 0.0001 compared to control, as measured by one-way ANOVA.</p>
Full article ">Figure 9
<p>MtMP changes upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>) and platinum(IV) (<b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>) complexes, as well as cisplatin in MDA−MB−231 and HT29 cells at 24, 48 and 72 h, as described in <a href="#sec2dot11-cancers-16-02544" class="html-sec">Section 2.11</a>. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments where samples were run in triplicate. ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001 compared with control, as measured by one-way ANOVA.</p>
Full article ">Figure 10
<p><b>Effect of platinum complexes and cisplatin on β−tubulin and F−actin.</b> Immunofluorescence upon treatment with Pt<sup>II</sup> and Pt<sup>IV</sup> complexes, as well as cisplatin in MDA−MB−231, MCF10A and HT29 cells at 72 h, as described in <a href="#sec2dot12-cancers-16-02544" class="html-sec">Section 2.12</a>: (<b>A</b>). MDA−MB−231 airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>B</b>). MDA−MB−231 cell size (µm<sup>2</sup>) (<b>C</b>). Actin expression in MDA−MB−231 (<b>D</b>). Tubulin expression in MDA−MB−231 (<b>E</b>). Edge/Cell ratio of actin expression in MDA−MB−231 (<b>F</b>). Edge/Cell ratio of tubulin expression in MDA−MB−231 (<b>G</b>). Nucleus/Cell ratio of actin expression in MDA−MB−231 (<b>H</b>). Nucleus/Cell ratio of tubulin expression in MDA−MB−231. (<b>I</b>). MCF10A airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>J</b>). MCF10A cell size (µm<sup>2</sup>) (<b>K</b>). Actin expression in MCF10A. (<b>L</b>). Tubulin expression in MCF10A. (<b>M</b>). Edge/Cell ratio of actin expression in MCF10A. (<b>N</b>). Edge/Cell ratio of tubulin expression in MCF10A. (<b>O</b>). Nucleus/Cell ratio of actin expression in MCF10A. (<b>P</b>). Nucleus/Cell ratio of tubulin expression in MCF10A. (<b>Q</b>). HT29 airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>R</b>). HT29 cell size (µm<sup>2</sup>) (<b>S</b>). Actin expression in HT29. (<b>T</b>). Tubulin expression in HT29. (<b>U</b>). Edge/Cell ratio of actin expression in HT29. (<b>V</b>). Edge/Cell ratio of tubulin expression in HT29. (<b>W</b>). Nucleus/Cell ratio of actin expression in HT29. (<b>X</b>). Nucleus/Cell ratio of tubulin expression in HT29. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001, as measured by one-way ANOVA.</p>
Full article ">Figure 10 Cont.
<p><b>Effect of platinum complexes and cisplatin on β−tubulin and F−actin.</b> Immunofluorescence upon treatment with Pt<sup>II</sup> and Pt<sup>IV</sup> complexes, as well as cisplatin in MDA−MB−231, MCF10A and HT29 cells at 72 h, as described in <a href="#sec2dot12-cancers-16-02544" class="html-sec">Section 2.12</a>: (<b>A</b>). MDA−MB−231 airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>B</b>). MDA−MB−231 cell size (µm<sup>2</sup>) (<b>C</b>). Actin expression in MDA−MB−231 (<b>D</b>). Tubulin expression in MDA−MB−231 (<b>E</b>). Edge/Cell ratio of actin expression in MDA−MB−231 (<b>F</b>). Edge/Cell ratio of tubulin expression in MDA−MB−231 (<b>G</b>). Nucleus/Cell ratio of actin expression in MDA−MB−231 (<b>H</b>). Nucleus/Cell ratio of tubulin expression in MDA−MB−231. (<b>I</b>). MCF10A airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>J</b>). MCF10A cell size (µm<sup>2</sup>) (<b>K</b>). Actin expression in MCF10A. (<b>L</b>). Tubulin expression in MCF10A. (<b>M</b>). Edge/Cell ratio of actin expression in MCF10A. (<b>N</b>). Edge/Cell ratio of tubulin expression in MCF10A. (<b>O</b>). Nucleus/Cell ratio of actin expression in MCF10A. (<b>P</b>). Nucleus/Cell ratio of tubulin expression in MCF10A. (<b>Q</b>). HT29 airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>R</b>). HT29 cell size (µm<sup>2</sup>) (<b>S</b>). Actin expression in HT29. (<b>T</b>). Tubulin expression in HT29. (<b>U</b>). Edge/Cell ratio of actin expression in HT29. (<b>V</b>). Edge/Cell ratio of tubulin expression in HT29. (<b>W</b>). Nucleus/Cell ratio of actin expression in HT29. (<b>X</b>). Nucleus/Cell ratio of tubulin expression in HT29. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001, as measured by one-way ANOVA.</p>
Full article ">Figure 10 Cont.
<p><b>Effect of platinum complexes and cisplatin on β−tubulin and F−actin.</b> Immunofluorescence upon treatment with Pt<sup>II</sup> and Pt<sup>IV</sup> complexes, as well as cisplatin in MDA−MB−231, MCF10A and HT29 cells at 72 h, as described in <a href="#sec2dot12-cancers-16-02544" class="html-sec">Section 2.12</a>: (<b>A</b>). MDA−MB−231 airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>B</b>). MDA−MB−231 cell size (µm<sup>2</sup>) (<b>C</b>). Actin expression in MDA−MB−231 (<b>D</b>). Tubulin expression in MDA−MB−231 (<b>E</b>). Edge/Cell ratio of actin expression in MDA−MB−231 (<b>F</b>). Edge/Cell ratio of tubulin expression in MDA−MB−231 (<b>G</b>). Nucleus/Cell ratio of actin expression in MDA−MB−231 (<b>H</b>). Nucleus/Cell ratio of tubulin expression in MDA−MB−231. (<b>I</b>). MCF10A airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>J</b>). MCF10A cell size (µm<sup>2</sup>) (<b>K</b>). Actin expression in MCF10A. (<b>L</b>). Tubulin expression in MCF10A. (<b>M</b>). Edge/Cell ratio of actin expression in MCF10A. (<b>N</b>). Edge/Cell ratio of tubulin expression in MCF10A. (<b>O</b>). Nucleus/Cell ratio of actin expression in MCF10A. (<b>P</b>). Nucleus/Cell ratio of tubulin expression in MCF10A. (<b>Q</b>). HT29 airy scan images at 20<math display="inline"><semantics> <mo>×</mo> </semantics></math>. (<b>R</b>). HT29 cell size (µm<sup>2</sup>) (<b>S</b>). Actin expression in HT29. (<b>T</b>). Tubulin expression in HT29. (<b>U</b>). Edge/Cell ratio of actin expression in HT29. (<b>V</b>). Edge/Cell ratio of tubulin expression in HT29. (<b>W</b>). Nucleus/Cell ratio of actin expression in HT29. (<b>X</b>). Nucleus/Cell ratio of tubulin expression in HT29. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001, as measured by one-way ANOVA.</p>
Full article ">Figure 11
<p>Cell migration (scratch wound healing assay). Percent relative wound density, wound confluence and wound width measured upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>) and platinum(IV) (<b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>) complexes, as well as cisplatin in (<b>A</b>). MDA−MB−231 and (<b>B</b>). HT29 cells up to 72 h, as described in <a href="#sec2dot13-cancers-16-02544" class="html-sec">Section 2.13</a>. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments where samples were run in triplicate. * <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; 0.0001 compared with control, as measured by one-way ANOVA.</p>
Full article ">Figure 12
<p>Protein expression upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b> (lane 3), <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> (lane 4) and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> (lane 5)) and platinum(IV) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 6), <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 7) and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 8)) complexes, as well as cisplatin (lane 2) in MDA−MB−231 and HT29 cells at 72 h, as described in <a href="#sec2dot14-cancers-16-02544" class="html-sec">Section 2.14</a>: (<b>A</b>). MDA−MB−231 microtubule cytoskeleton markers (<b>B</b>). MDA−MB−231 cell proliferation markers (<b>C</b>). MDA−MB−231 intrinsic and extrinsic apoptotic cell death markers (<b>D</b>). MDA−MB−231 autophagy markers (<b>E</b>). HT29 microtubule cytoskeleton markers (<b>F</b>). HT29 cell proliferation markers (<b>G</b>). HT29 intrinsic and extrinsic apoptotic cell death markers (<b>H</b>). HT29 autophagy markers. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001 compared with control (lane 1), as measured by unpaired Student’s <span class="html-italic">t</span>-test. The full uncropped Western blot, with its corresponding molecular markers, is represented in <a href="#app1-cancers-16-02544" class="html-app">Figures S24–S31</a>.</p>
Full article ">Figure 12 Cont.
<p>Protein expression upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b> (lane 3), <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> (lane 4) and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> (lane 5)) and platinum(IV) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 6), <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 7) and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 8)) complexes, as well as cisplatin (lane 2) in MDA−MB−231 and HT29 cells at 72 h, as described in <a href="#sec2dot14-cancers-16-02544" class="html-sec">Section 2.14</a>: (<b>A</b>). MDA−MB−231 microtubule cytoskeleton markers (<b>B</b>). MDA−MB−231 cell proliferation markers (<b>C</b>). MDA−MB−231 intrinsic and extrinsic apoptotic cell death markers (<b>D</b>). MDA−MB−231 autophagy markers (<b>E</b>). HT29 microtubule cytoskeleton markers (<b>F</b>). HT29 cell proliferation markers (<b>G</b>). HT29 intrinsic and extrinsic apoptotic cell death markers (<b>H</b>). HT29 autophagy markers. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001 compared with control (lane 1), as measured by unpaired Student’s <span class="html-italic">t</span>-test. The full uncropped Western blot, with its corresponding molecular markers, is represented in <a href="#app1-cancers-16-02544" class="html-app">Figures S24–S31</a>.</p>
Full article ">Figure 12 Cont.
<p>Protein expression upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b> (lane 3), <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> (lane 4) and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> (lane 5)) and platinum(IV) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 6), <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 7) and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 8)) complexes, as well as cisplatin (lane 2) in MDA−MB−231 and HT29 cells at 72 h, as described in <a href="#sec2dot14-cancers-16-02544" class="html-sec">Section 2.14</a>: (<b>A</b>). MDA−MB−231 microtubule cytoskeleton markers (<b>B</b>). MDA−MB−231 cell proliferation markers (<b>C</b>). MDA−MB−231 intrinsic and extrinsic apoptotic cell death markers (<b>D</b>). MDA−MB−231 autophagy markers (<b>E</b>). HT29 microtubule cytoskeleton markers (<b>F</b>). HT29 cell proliferation markers (<b>G</b>). HT29 intrinsic and extrinsic apoptotic cell death markers (<b>H</b>). HT29 autophagy markers. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001 compared with control (lane 1), as measured by unpaired Student’s <span class="html-italic">t</span>-test. The full uncropped Western blot, with its corresponding molecular markers, is represented in <a href="#app1-cancers-16-02544" class="html-app">Figures S24–S31</a>.</p>
Full article ">Figure 12 Cont.
<p>Protein expression upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b> (lane 3), <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> (lane 4) and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> (lane 5)) and platinum(IV) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 6), <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 7) and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 8)) complexes, as well as cisplatin (lane 2) in MDA−MB−231 and HT29 cells at 72 h, as described in <a href="#sec2dot14-cancers-16-02544" class="html-sec">Section 2.14</a>: (<b>A</b>). MDA−MB−231 microtubule cytoskeleton markers (<b>B</b>). MDA−MB−231 cell proliferation markers (<b>C</b>). MDA−MB−231 intrinsic and extrinsic apoptotic cell death markers (<b>D</b>). MDA−MB−231 autophagy markers (<b>E</b>). HT29 microtubule cytoskeleton markers (<b>F</b>). HT29 cell proliferation markers (<b>G</b>). HT29 intrinsic and extrinsic apoptotic cell death markers (<b>H</b>). HT29 autophagy markers. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001 compared with control (lane 1), as measured by unpaired Student’s <span class="html-italic">t</span>-test. The full uncropped Western blot, with its corresponding molecular markers, is represented in <a href="#app1-cancers-16-02544" class="html-app">Figures S24–S31</a>.</p>
Full article ">Figure 12 Cont.
<p>Protein expression upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b> (lane 3), <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> (lane 4) and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> (lane 5)) and platinum(IV) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 6), <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 7) and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (lane 8)) complexes, as well as cisplatin (lane 2) in MDA−MB−231 and HT29 cells at 72 h, as described in <a href="#sec2dot14-cancers-16-02544" class="html-sec">Section 2.14</a>: (<b>A</b>). MDA−MB−231 microtubule cytoskeleton markers (<b>B</b>). MDA−MB−231 cell proliferation markers (<b>C</b>). MDA−MB−231 intrinsic and extrinsic apoptotic cell death markers (<b>D</b>). MDA−MB−231 autophagy markers (<b>E</b>). HT29 microtubule cytoskeleton markers (<b>F</b>). HT29 cell proliferation markers (<b>G</b>). HT29 intrinsic and extrinsic apoptotic cell death markers (<b>H</b>). HT29 autophagy markers. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. * <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; 0.0001 compared with control (lane 1), as measured by unpaired Student’s <span class="html-italic">t</span>-test. The full uncropped Western blot, with its corresponding molecular markers, is represented in <a href="#app1-cancers-16-02544" class="html-app">Figures S24–S31</a>.</p>
Full article ">Figure 13
<p>Differential proteins upon treatment with platinum(II) (<b>Pt<sup>II</sup>PHEN<span class="html-italic">SS</span></b>, <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> and <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b>) and platinum(IV) (<b>Pt<sup>IV</sup>PHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b>, <b>Pt<sup>IV</sup>5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> and <b>Pt<sup>IV</sup>56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b>) complexes, as well as cisplatin in (<b>A</b>). MDA−MB−231 and (<b>B</b>). HT29 cells at 72 h. UpSet plots summarise the differential protein expression analysis for the prodrugs and ligands. The differential protein expression analysis for the prodrugs and ligands is summarised in UpSet plots. Each figure’s bottom-left horizontal bar graph displays the total number of proteins with variations in log 2-fold change in expression for each complex. The same differentially expressed proteins that are shared by the complexes compared on the left are shown by the black circles connected to the right of these bar graphs. The vertical bar graph at the top quantifies the number of proteins with similar log 2-fold change expression differences in the drug comparisons.</p>
Full article ">Figure 14
<p>Proteomic analysis of MDA−MB−231 and HT29 upon treatment with <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> as described in <a href="#sec2dot15-cancers-16-02544" class="html-sec">Section 2.15</a>: (<b>A</b>). MDA−MB−231 principal component analysis. (<b>B</b>). MDA−MB−231 number of differentially expressed proteins (DEPs). (<b>C</b>). MDA−MB−231 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components, and molecular function in MDA−MB−231, (<b>D</b>). Downregulated and (<b>E</b>). Upregulated proteins. MDA−MB−231 pathway enrichment and gene act network analysis with the most significance of (<b>F</b>). Downregulated pathways; mRNA metabolic process, GTP binding and nucleoside activity and (<b>G</b>). Upregulated pathways; nucleotide binding and pyrophosphatase activity. (<b>H</b>). HT29 principal component analysis. (<b>I</b>). HT29 number of differentially expressed proteins (DEPs). (<b>J</b>). HT29 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components, and molecular function in HT29, (<b>K</b>). Downregulated and (<b>L</b>). Upregulated proteins. HT29 pathway enrichment and gene act network analysis with the most significance of (<b>M</b>). Downregulated pathways; molecular metabolic and catabolic processes and RNA splicing and (<b>N</b>). Upregulated pathways; protein folding, response to toxic substance, detoxification, secretion, and exocytosis activity. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. For clarity, a larger representation is provided in the <a href="#app1-cancers-16-02544" class="html-app">Supplementary Figure S22</a>.</p>
Full article ">Figure 14 Cont.
<p>Proteomic analysis of MDA−MB−231 and HT29 upon treatment with <b>Pt<sup>II</sup>5ME<span class="html-italic">SS</span></b> as described in <a href="#sec2dot15-cancers-16-02544" class="html-sec">Section 2.15</a>: (<b>A</b>). MDA−MB−231 principal component analysis. (<b>B</b>). MDA−MB−231 number of differentially expressed proteins (DEPs). (<b>C</b>). MDA−MB−231 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components, and molecular function in MDA−MB−231, (<b>D</b>). Downregulated and (<b>E</b>). Upregulated proteins. MDA−MB−231 pathway enrichment and gene act network analysis with the most significance of (<b>F</b>). Downregulated pathways; mRNA metabolic process, GTP binding and nucleoside activity and (<b>G</b>). Upregulated pathways; nucleotide binding and pyrophosphatase activity. (<b>H</b>). HT29 principal component analysis. (<b>I</b>). HT29 number of differentially expressed proteins (DEPs). (<b>J</b>). HT29 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components, and molecular function in HT29, (<b>K</b>). Downregulated and (<b>L</b>). Upregulated proteins. HT29 pathway enrichment and gene act network analysis with the most significance of (<b>M</b>). Downregulated pathways; molecular metabolic and catabolic processes and RNA splicing and (<b>N</b>). Upregulated pathways; protein folding, response to toxic substance, detoxification, secretion, and exocytosis activity. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. For clarity, a larger representation is provided in the <a href="#app1-cancers-16-02544" class="html-app">Supplementary Figure S22</a>.</p>
Full article ">Figure 15
<p>Proteomic analysis of MDA−MB−231 and HT29 upon treatment with <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> as described in <a href="#sec2dot15-cancers-16-02544" class="html-sec">Section 2.15</a>: (<b>A</b>). MDA−MB−231 principal component analysis. (<b>B</b>). MDA−MB−231 number of differentially expressed proteins (DEPs). (<b>C</b>). MDA−MB−231 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components and molecular function in MDA−MB−231 (<b>D</b>). Downregulated and (<b>E</b>). Upregulated proteins. MDA−MB−231 pathway enrichment and gene act network analysis with the most significance of (<b>F</b>). Downregulated pathways; translational initiation and mRNA metabolic process and (<b>G</b>). Upregulated pathways; localisation in cell, cellular organisation, protein localisation and exocytosis. (<b>H</b>). HT29 principal component analysis. (<b>I</b>). HT29 number of differentially expressed proteins (DEPs). (<b>J</b>). HT29 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components and molecular function in HT29, (<b>K</b>). Downregulated and (<b>L</b>). Upregulated proteins. HT29 pathway enrichment and gene act network analysis with the most significance of (<b>M</b>). Downregulated pathways; RNA binding and nucleotide binding and (<b>N</b>). Upregulated pathways; ion transmembrane transport, active transport, and oxidoreductase activity. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. For clarity, a larger representation is provided in the <a href="#app1-cancers-16-02544" class="html-app">Supplementary Figure S23</a>.</p>
Full article ">Figure 15 Cont.
<p>Proteomic analysis of MDA−MB−231 and HT29 upon treatment with <b>Pt<sup>II</sup>56ME<span class="html-italic">SS</span></b> as described in <a href="#sec2dot15-cancers-16-02544" class="html-sec">Section 2.15</a>: (<b>A</b>). MDA−MB−231 principal component analysis. (<b>B</b>). MDA−MB−231 number of differentially expressed proteins (DEPs). (<b>C</b>). MDA−MB−231 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components and molecular function in MDA−MB−231 (<b>D</b>). Downregulated and (<b>E</b>). Upregulated proteins. MDA−MB−231 pathway enrichment and gene act network analysis with the most significance of (<b>F</b>). Downregulated pathways; translational initiation and mRNA metabolic process and (<b>G</b>). Upregulated pathways; localisation in cell, cellular organisation, protein localisation and exocytosis. (<b>H</b>). HT29 principal component analysis. (<b>I</b>). HT29 number of differentially expressed proteins (DEPs). (<b>J</b>). HT29 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components and molecular function in HT29, (<b>K</b>). Downregulated and (<b>L</b>). Upregulated proteins. HT29 pathway enrichment and gene act network analysis with the most significance of (<b>M</b>). Downregulated pathways; RNA binding and nucleotide binding and (<b>N</b>). Upregulated pathways; ion transmembrane transport, active transport, and oxidoreductase activity. Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments. For clarity, a larger representation is provided in the <a href="#app1-cancers-16-02544" class="html-app">Supplementary Figure S23</a>.</p>
Full article ">Figure 16
<p>Proteomic analysis of MDA−MB−231 and HT29 upon treatment with cisplatin as described in <a href="#sec2dot15-cancers-16-02544" class="html-sec">Section 2.15</a>: (<b>A</b>). MDA−MB−231 principal component analysis. (<b>B</b>). MDA−MB−231 number of differentially expressed proteins (DEPs). (<b>C</b>). MDA−MB−231 volcano plot of DEPs upregulated (red) and downregulated (green). GO enriched biological processes, cellular components and molecular function in MDA−MB−231, (<b>D</b>). Downregulated and (<b>E</b>). Upregulated proteins. (<b>F</b>). HT29 principal component analysis. (<b>G</b>). HT29 number of differentially expressed proteins (DEPs). (<b>H</b>). HT29 volcano plot of DEPs upregulated (red) and downregulated (green). Data points denote mean ± SEM. <span class="html-italic">n</span> = 3 from three independent experiments.</p>
Full article ">Scheme 1
<p>Chemical structures of <b>PtIIPHEN<span class="html-italic">SS</span></b> (<b>A</b>), <b>PtII5ME<span class="html-italic">SS</span></b> (<b>B</b>), <b>PtII56ME<span class="html-italic">SS</span></b> (<b>C</b>), <b>PtIVPHEN<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (<b>D</b>), <b>PtIV5ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (<b>E</b>) and <b>PtIV56ME<span class="html-italic">SS</span>(OH)<sub>2</sub></b> (<b>F</b>).</p>
Full article ">
26 pages, 4996 KiB  
Review
The Recent Trends of Systemic Treatments and Locoregional Therapies for Cholangiocarcinoma
by Abdullah Esmail, Mohamed Badheeb, Batool Wael Alnahar, Bushray Almiqlash, Yara Sakr, Ebtesam Al-Najjar, Ali Awas, Mohammad Alsayed, Bayan Khasawneh, Mohammed Alkhulaifawi, Amneh Alsaleh, Ala Abudayyeh, Yaser Rayyan and Maen Abdelrahim
Pharmaceuticals 2024, 17(7), 910; https://doi.org/10.3390/ph17070910 - 8 Jul 2024
Cited by 1 | Viewed by 1353
Abstract
Cholangiocarcinoma (CCA) is a hepatic malignancy that has a rapidly increasing incidence. CCA is anatomically classified into intrahepatic (iCCA) and extrahepatic (eCCA), which is further divided into perihilar (pCCA) and distal (dCCA) subtypes, with higher incidence rates in Asia. Despite its rarity, CCA [...] Read more.
Cholangiocarcinoma (CCA) is a hepatic malignancy that has a rapidly increasing incidence. CCA is anatomically classified into intrahepatic (iCCA) and extrahepatic (eCCA), which is further divided into perihilar (pCCA) and distal (dCCA) subtypes, with higher incidence rates in Asia. Despite its rarity, CCA has a low 5-year survival rate and remains the leading cause of primary liver tumor-related death over the past 10–20 years. The systemic therapy section discusses gemcitabine-based regimens as primary treatments, along with oxaliplatin-based options. Second-line therapy is limited but may include short-term infusional fluorouracil (FU) plus leucovorin (LV) and oxaliplatin. The adjuvant therapy section discusses approaches to improve overall survival (OS) post-surgery. However, only a minority of CCA patients qualify for surgical resection. In comparison to adjuvant therapies, neoadjuvant therapy for unresectable cases shows promise. Gemcitabine and cisplatin indicate potential benefits for patients awaiting liver transplantation. The addition of immunotherapies to chemotherapy in combination is discussed. Nivolumab and innovative approaches like CAR-T cells, TRBAs, and oncolytic viruses are explored. We aim in this review to provide a comprehensive report on the systemic and locoregional therapies for CCA. Full article
(This article belongs to the Special Issue Drug Treatment of Cholangiocarcinoma)
Show Figures

Figure 1

Figure 1
<p>Anatomical classification of cholangiocarcinoma (CCA) based on its origin.</p>
Full article ">Figure 2
<p>A comprehensive diagram of cholangiocarcinoma treatment options. IHC—immunohistochemistry, MRI—magnetic resonance imaging, CT—computed tomography, PET—positron emission tomography, ECOG—Eastern Cooperative Oncology Group, 5-FU—fluorouracil, dMMR-deficient DNA mismatch repair, MSI-H—microsatellite instability high.</p>
Full article ">
16 pages, 2872 KiB  
Article
Interactions with DNA Models of the Oxaliplatin Analog (cis-1,3-DACH)PtCl2 
by Alessandra Barbanente, Paride Papadia, Anna Maria Di Cosola, Concetta Pacifico, Giovanni Natile, James D. Hoeschele and Nicola Margiotta
Int. J. Mol. Sci. 2024, 25(13), 7392; https://doi.org/10.3390/ijms25137392 - 5 Jul 2024
Viewed by 1001
Abstract
It is generally accepted that adjacent guanine residues in DNA are the primary target for platinum antitumor drugs and that differences in the conformations of the Pt-DNA adducts can play a role in their antitumor activity. In this study, we investigated the effect [...] Read more.
It is generally accepted that adjacent guanine residues in DNA are the primary target for platinum antitumor drugs and that differences in the conformations of the Pt-DNA adducts can play a role in their antitumor activity. In this study, we investigated the effect of the carrier ligand cis-1,3-diaminocyclohexane (cis-1,3-DACH) upon formation, stability, and stereochemistry of the (cis-1,3-DACH)PtG2 and (cis-1,3-DACH)Pt(d(GpG)) adducts (G = 9-EthlyGuanine, guanosine, 5′- and 3′-guanosine monophosphate; d(GpG) = deoxyguanosil(3′-5′)deoxyguanosine). A peculiar feature of the cis-1,3-DACH carrier ligand is the steric bulk of the diamine, which is asymmetric with respect to the Pt-coordination plane. The (cis-1,3-DACH)Pt(5′GMP)2 and (cis-1,3-DACH)Pt(3′GMP)2 adducts show preference for the ΛHT and ∆HT conformations, respectively (HT stands for Head-to-Tail). Moreover, the increased intensity of the circular dichroism signals in the cis-1,3-DACH derivatives with respect to the analogous cis-(NH3)2 species could be a consequence of the greater bite angle of the cis-1,3-DACH carrier ligand with respect to cis-(NH3)2. Finally, the (cis-1,3-DACH)Pt(d(GpG)) adduct is present in two isomeric forms, each one giving a pair of H8 resonances linked by a NOE cross peak. The two isomers were formed in comparable amounts and had a dominance of the HH conformer but with some contribution of the ΔHT conformer which is related to the HH conformer by having the 3′-G base flipped with respect to the 5′-G residue. Full article
(This article belongs to the Special Issue Nucleic Acid Recognition and Pharmaceutical Ligand Design)
Show Figures

Figure 1

Figure 1
<p>Structure and numbering of atoms of the bis-adduct obtained in the reaction with G = 9-EthylGuanine (9-EtG). (<b>a</b>) <sup>1</sup>H-NMR spectra of the reaction between [Pt(OSO<sub>3</sub>)(OH<sub>2</sub>)(<span class="html-italic">cis</span>-1,3-DACH)] and 9-EtG in D<sub>2</sub>O, pH* 3.00: after 2 h (black spectrum) and after three days (red spectrum) at 37 °C; (<b>b</b>) COSY spectrum of (<span class="html-italic">cis</span>-1,3-DACH)Pt(9-EtG)<sub>2</sub> adduct in D<sub>2</sub>O.</p>
Full article ">Figure 2
<p>COSY spectrum of (<span class="html-italic">cis</span>-1,3-DACH)Pt(Guo)<sub>2</sub> adduct in D<sub>2</sub>O. (<b>a</b>,<b>b</b>) expansions of the boxes in the COSY spectrum.</p>
Full article ">Figure 3
<p>Expansions of the 2D COSY (<b>a</b>,<b>b</b>) and [<sup>1</sup>H-<sup>13</sup>C]-HSQC (<b>c</b>) spectra of (<span class="html-italic">cis</span>-1,3-DACH)Pt(5′GMP)<sub>2</sub> in D<sub>2</sub>O.</p>
Full article ">Figure 4
<p>CD spectra of (<span class="html-italic">cis</span>-1,3-DACH)Pt(5′GMP)<sub>2</sub> (<b>a</b>,<b>b</b>) and (<span class="html-italic">cis</span>-1,3-DACH)Pt(3′GMP)<sub>2</sub> (<b>c</b>,<b>d</b>) in solution: (<b>a</b>,<b>c</b>) pH = 7–3; (<b>b</b>,<b>d</b>) pH = 7–11.</p>
Full article ">Figure 5
<p>Top right: COSY spectrum of (<span class="html-italic">cis</span>-1,3-DACH)Pt(3′GMP)<sub>2</sub> in D<sub>2</sub>O. (<b>a</b>–<b>c</b>) expansions of the COSY spectrum. (<b>d</b>) [<sup>1</sup>H-<sup>13</sup>C]-HSQC of (<span class="html-italic">cis</span>-1,3-DACH)Pt(3′GMP)<sub>2</sub> in D<sub>2</sub>O.</p>
Full article ">Figure 6
<p>(<b>a</b>) <sup>1</sup>H-NMR spectra of the reaction between [Pt(OSO<sub>3</sub>)(OH<sub>2</sub>)(<span class="html-italic">cis</span>-1,3-DACH)] and d(GpG) in D<sub>2</sub>O, pH* 3.00: soon after mixing of the reagents (black) and after one day (green) at 37 °C; (<b>b</b>) Expansion of the NOESY spectrum of (<span class="html-italic">cis</span>-1,3-DACH)Pt(d(GpG)) recorded after completion of the reaction.</p>
Full article ">Scheme 1
<p>Possible conformers (rotamers) of <span class="html-italic">cis</span>-(diam(m)ine)PtG<sub>2</sub> adducts. The arrows represent the G bases, with their tip symbolizing the hydrogen atom in position 8. In the Head-to-Head (HH) arrangement, both G residues have their H8 atoms on the same side of the Pt coordination plane, while in the Head-to-Tail (HT) arrangement, the two Gs residues have their H8 atoms on opposite sides of the Pt coordination plane. In the latter case, the adduct is asymmetric and can have Δ or Λ chirality. Interconversion between conformers is possible via rotation about the Pt–G bond. In the case of fast rotation on the NMR time scale, only one H8G signal will be observed in the case of G = EtG, while two H8G signals will be observed in the case of G = guanosine/guanotide (see following discussion).</p>
Full article ">Scheme 2
<p>Possible HH adducts obtainable for (<span class="html-italic">cis</span>-1,3-DACH)Pt(d(GpG)). The arrows represent the G bases with their tip symbolizing the hydrogen atom in position 8.</p>
Full article ">
Back to TopTop