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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (5,619)

Search Parameters:
Keywords = melanoma

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 2928 KiB  
Article
A Novel Chimeric Oncolytic Virus Mediates a Multifaceted Cellular Immune Response in a Syngeneic B16 Melanoma Model
by Sonja Glauß, Victoria Neumeyer, Lorenz Hanesch, Janina Marek, Nina Hartmann, Gabriela M. Wiedemann and Jennifer Altomonte
Cancers 2024, 16(19), 3405; https://doi.org/10.3390/cancers16193405 (registering DOI) - 6 Oct 2024
Abstract
Background/Objectives: Oncolytic virotherapy is a promising approach in cancer immunotherapy. We have previously described a recombinant hybrid oncolytic virus (OV), VSV-NDV, which has a favorable safety profile and therapeutic immunogenicity, leading to direct oncolysis, abscopal effects, and prolonged survival in syngeneic in vivo [...] Read more.
Background/Objectives: Oncolytic virotherapy is a promising approach in cancer immunotherapy. We have previously described a recombinant hybrid oncolytic virus (OV), VSV-NDV, which has a favorable safety profile and therapeutic immunogenicity, leading to direct oncolysis, abscopal effects, and prolonged survival in syngeneic in vivo tumor models. While OVs are known to mediate systemic anti-tumor immune responses, the detailed characterization of local and systemic immune responses to fusogenic oncolytic virotherapy remains unexplored. Methods and Results: We analyzed immune cell compartments in the spleen, blood, tumor-draining lymph nodes (TDLNs), and tumors over the course of VSV-NDV therapy in a bilateral syngeneic melanoma mouse model. Our results revealed significant local infiltration and activation of T lymphocytes in tumors and globally in the blood and spleen. Notably, in vivo CD8+ T cell depletion led to complete abrogation of the tumor response, highlighting the crucial role of T cells in promoting the therapeutic effects of oncolytic VSV-NDV. In vitro co-culture experiments enabled the interrogation of human immune cell responses to VSV-NDV-mediated oncolysis. Human peripheral blood mononuclear cells (PBMCs) were efficiently stimulated by exposure to VSV-NDV-infected cancer cells, which recapitulates the in vivo murine findings. Conclusions: Taken together, these data characterize a broad anti-tumor immune cell response to oncolytic VSV-NDV therapy and suggest that CD8+ T cells play a decisive role in therapeutic outcome, which supports the further development of this chimeric vector as a multimechanistic immunotherapy for solid cancers. Full article
(This article belongs to the Special Issue Oncolytic Viruses as an Emerging Aspect of Immune Oncology)
Show Figures

Figure 1

Figure 1
<p>VSV-NDV reduces the tumor mass and mediates the expression of cytokines in murine melanoma. (<b>A</b>) The experimental set-up of the mechanistic endpoint experiment. C57BL6/J mice were implanted with B16OVA cells subcutaneously on contralateral flanks. Seven days later, the mice were randomly distributed into treatment groups (N = 6–8) and injected intratumorally with VSV-NDV (VN) at a dose of 1 × 10<sup>7</sup> TCID50 or PBS in an equal volume of 50 µL on day 0, 3, 6. Tumors were collected on days 2 and 10 post-treatment start. Created in BioRender. Altomonte, J. (2024) <a href="https://BioRender.com/p46u790" target="_blank">https://BioRender.com/p46u790</a>. (<b>B</b>) Tumor weight was measured on day 10. (<b>C</b>) Quantitative real-time RT-PCR was performed to measure intratumoral mRNA expression on day 2 (IL-15, IFN-<math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math>, IFN-α, IFN-β, and CXCL10) or day 10 (T-bet). Relative gene expression was quantified by normalization to GAPDH using the 2<sup>−ΔΔCt</sup> method. Data points indicate values of individual replicates (and group means are presented as bars); * indicates <span class="html-italic">p</span> ˂ 0.05, and ** indicates <span class="html-italic">p</span> ˂ 0.01).</p>
Full article ">Figure 2
<p>VSV-NDV treatment leads to T cell activation in tumors and lymph nodes. Bilateral B16OVA-bearing mice were treated intratumorally with VSV-NDV or PBS and analyzed by flow cytometry. (<b>A</b>) CD8<sup>+</sup>, CD4<sup>+</sup>, and OVA-TCR<sup>+</sup>CD8<sup>+</sup> T cells in the tumor on day 10. (<b>B</b>) VSV-TCR<sup>+</sup>CD8<sup>+</sup> T cells and the correlation of VSV<sup>+</sup>TCR and OVA<sup>+</sup>TCR in the tumor on day 10. (<b>C</b>) CD69 expression on CD8<sup>+</sup> T cells on day 10 in the tumor and PD-1 expression on CD4<sup>+</sup> and CD8<sup>+</sup> T cells on day 10 in the tumor. (<b>D</b>) OVA-TCR<sup>+</sup>CD8<sup>+</sup> or OVA-TCR<sup>+</sup>CD4<sup>+</sup> T cells on day 2 and day 10. (<b>E</b>) CD69 expression on day 2 and PD-1 expression on day 10 of OVA-TCR<sup>+</sup>CD8<sup>+</sup> T cells. N = 5–15 mice; data points indicate values of individual replicates (and group means + SD are presented as bars); * indicates <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01, *** <span class="html-italic">p</span> ˂ 0.001, **** <span class="html-italic">p</span> ˂ 0.0001.</p>
Full article ">Figure 3
<p>Intratumoral VSV-NDV treatment modulates CD44/CD62L expression patterns on T cells in B16 melanoma lesions and tumor-draining lymph nodes (TDLN). Bilateral B16OVA lesions and TDLNs were isolated from female C57Bl/6 mice treated with rVSV-NDV or PBS on day 10 post-treatment and analyzed by flow cytometry. (<b>A</b>) CD44/CD62L expression on CD8<sup>+</sup> T cells and CD122<sup>+</sup>CD44<sup>+</sup>CD62L<sup>+</sup>CD8<sup>+</sup> T cells on day 10 in the tumor. (<b>B</b>) CD44<sup>+</sup>CD62L<sup>−</sup>CD4<sup>+</sup>, CD44<sup>+</sup>CD62L<sup>−</sup>CD8<sup>+</sup>, and OVA-specific CD44<sup>+</sup>CD62L<sup>+</sup>CD8<sup>+</sup> T cells on day 10 in the TDLNs. N = 5–15; data points indicate values of individual replicates (and group means + SD are presented as bars); * indicates <span class="html-italic">p</span> ˂ 0.05, ** indicates <span class="html-italic">p</span> ˂ 0.01.</p>
Full article ">Figure 4
<p>Local VSV-NDV treatment mediates systemic immune-stimulatory effects. C57BL6/J mice were implanted with B16OVA cells subcutaneously on contralateral flanks and, 7 days later, injected intratumorally with VSV-NDV (VN) at a dose of 1 × 10<sup>7</sup> TCID50 or PBS and analyzed by flow cytometry for (<b>A</b>) OVA-TCR<sup>+</sup>CD8<sup>+</sup> and VSV-TCR<sup>+</sup>CD8<sup>+</sup> T cells on day 10, (<b>B</b>) CD69 expression on OVA-TCR<sup>+</sup>CD8<sup>+</sup> T cells in the spleen on day 2 and PD-1<sup>+</sup>OVA<sup>+</sup>CD8<sup>+</sup> T cells in the blood and spleen on day 10, (<b>C</b>) CD44<sup>+</sup>CD62L<sup>−</sup> expression on CD4<sup>+</sup>, CD8<sup>+</sup>, and OVA-TCR<sup>+</sup>CD8<sup>+</sup> T cells on day 10 in blood, or (<b>D</b>) in the spleen. Data points indicate values of individual replicates (and group means ± SD are presented as bars); * indicates <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01, *** <span class="html-italic">p</span> ˂ 0.001, **** <span class="html-italic">p</span> ˂ 0.0001.</p>
Full article ">Figure 5
<p>The intratumoral VSV-NDV treatment of B16 melanoma induces the enrichment and activation of innate immune cells. C57BL6/J mice were implanted with B16OVA cells subcutaneously on contralateral flanks. After seven days, the mice (N = 6–8) were injected intratumorally with VSV-NDV (VN) at a dose of 1 × 10<sup>7</sup> TCID50 or PBS in 50 µL on day 0, 3, 6. TDLNs or tumors were analyzed by flow cytometry for (<b>A</b>) for iNKT cell number and CD69 expression, (<b>B</b>) NK cell number, CD69 expression, IFN-γ expression, and NKG2D expression in the TDLNs and mature NK cells, NKG2D<sup>+</sup> NK cells, and IFN-γ in the spleen, or (<b>C</b>) for DC number, CD86 expression, and MHC-I expression in TDLNs. Data points indicate values of individual replicates (and group means ± SD are presented as bars); * indicates <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01, *** <span class="html-italic">p</span> ˂ 0.001, **** <span class="html-italic">p</span> ˂ 0.0001.</p>
Full article ">Figure 6
<p>In vivo CD8<sup>+</sup> T cell depletion abrogates the therapeutic effects of rVSV-NDV treatment. (<b>A</b>) The experimental set-up of CD8<sup>+</sup> T cell depletion in B16OVA tumor-bearing mice treated with oncolytic rVSV-NDV vectors in vivo. Mice (N = 5–6) received either CD8 depletion antibody or IgG isotype control in 100 µL PBS by intraperitoneal injection 6 days after tumor implantation. One day later, complete depletion was confirmed by flow cytometry, and mice received VSV-NDV (1 × 10<sup>7</sup> TCID50) or PBS in a 50 µL volume by intratumoral injection on day 0, 3, 6, and 14. (<b>B</b>) Tumor growth was monitored daily. (<b>C</b>) Tumor weights on day 7 post-treatment start. Data points indicate values of individual replicates (and group means ± SD are presented as bars); * indicates <span class="html-italic">p</span> &lt; 0.05 and ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Human T cells and NK cells are activated by VSV-NDV-infected cancer cells in vitro. HepG2 cells pre-infected with VSV-NDV were co-cultured with human PBMCs in a 1:1 ratio for 24 h before analysis by flow cytometry. PBMCs that were not co-cultured served as controls. Activation was measured by (<b>A</b>) expression of CD69 on CD4<sup>+</sup> and CD8<sup>+</sup> T cells and CD25 expression on CD8<sup>+</sup> T cells or (<b>B</b>) expression of CD69 and IFN-γ on NK cells (N = 4–5 individual experiments); ** indicates <span class="html-italic">p</span> &lt; 0.01 and **** indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
12 pages, 1708 KiB  
Article
INEAS’s Cost-Effectiveness Analysis of Vemurafenib: Paving the Way for Value-Based Pricing in Tunisia
by Mouna Jameleddine, Nabil Harzallah, Hela Grati, Marie Christine Odabachian Jebali, Jaafar Chemli, Sebastián García Martí, Natalie Soto, Andrés Pichon-Riviere and Chokri Hamouda
J. Mark. Access Health Policy 2024, 12(4), 294-305; https://doi.org/10.3390/jmahp12040023 (registering DOI) - 6 Oct 2024
Abstract
The Tunisian Health Technology Assessment (HTA) body, INEAS, conducted a cost-effectiveness analysis (CEA) of vemurafenib in the treatment of locally advanced or metastatic BRAF V600-mutated melanoma. The objective of this analysis was to enable the use of value-based pricing as a new approach [...] Read more.
The Tunisian Health Technology Assessment (HTA) body, INEAS, conducted a cost-effectiveness analysis (CEA) of vemurafenib in the treatment of locally advanced or metastatic BRAF V600-mutated melanoma. The objective of this analysis was to enable the use of value-based pricing as a new approach to price negotiation. This study was part of a broader HTA report that was prepared in response to a joint request from the regulatory authorities and the CNAM, Tunisia’s compulsory insurance scheme. Our analysis was based on a probabilistic Markov cohort model that calculated the costs and quality-adjusted life years (QALY) associated with vemurafenib compared to the standard of care from a public payer perspective. The CEA indicated that vemurafenib provides a gain of 0.38 life years (1.78 vs. 1.4) for an incremental cost of USD 101,106.62 from the perspective of the main public payer (CNAM). This study revealed an incremental cost-effectiveness ratio (ICER) of 163,311.40 USD/QALY and 163,911.46 USD/QALY, respectively, from the CNAM and public health facilities’ perspectives. Vemurafenib cannot be considered cost-effective in terms of what has normally been considered a reasonable willingness to pay (WTP) in Tunisia. A significant price reduction would be necessary to bring the incremental cost-effectiveness ratio to an acceptable level. Full article
Show Figures

Figure 1

Figure 1
<p>Markov model with three health states.</p>
Full article ">Figure 2
<p>Tornado diagram illustrating the impact on the results of uncertainty in each parameter (CNAM perspective).</p>
Full article ">Figure 3
<p>Probabilistic sensitivity analysis (PHFs perspective).</p>
Full article ">Figure 4
<p>Acceptability curves.</p>
Full article ">
19 pages, 1804 KiB  
Article
Synthesis of Carborane–Thiazole Conjugates as Tyrosinase and 11β-Hydroxysteroid Dehydrogenase Inhibitors: Antiproliferative Activity and Molecular Docking Studies
by Beata Donarska, Joanna Cytarska, Dominika Kołodziej-Sobczak, Renata Studzińska, Daria Kupczyk, Angelika Baranowska-Łączkowska, Karol Jaroch, Paulina Szeliska, Barbara Bojko, Daria Różycka, Agnieszka B. Olejniczak, Wojciech Płaziński and Krzysztof Z. Łączkowski
Molecules 2024, 29(19), 4716; https://doi.org/10.3390/molecules29194716 (registering DOI) - 5 Oct 2024
Viewed by 238
Abstract
The presented study depicts the synthesis of 11 carborane–thiazole conjugates with anticancer activity, as well as an evaluation of their biological activity as inhibitors of two enzymes: tyrosinase and 11β-hydroxysteroid dehydrogenase type 1 (11β-HSD1). The overexpression of tyrosinase results in the intracellular accumulation [...] Read more.
The presented study depicts the synthesis of 11 carborane–thiazole conjugates with anticancer activity, as well as an evaluation of their biological activity as inhibitors of two enzymes: tyrosinase and 11β-hydroxysteroid dehydrogenase type 1 (11β-HSD1). The overexpression of tyrosinase results in the intracellular accumulation of melanin and can be observed in melanoma. The overexpression of 11β-HSD1 results in an elevation of glucocorticoid levels and has been associated with the aggravation of metabolic disorders such as type II diabetes mellitus and obesity. Recently, as the comorbidity of melanomas and metabolic disorders is being recognized as an important issue, the search for new therapeutic options has intensified. This study demonstrates that carborane–thiazole derivatives inhibit both enzymes, exerting beneficial effects. The antiproliferative action of all newly synthesized compounds was evaluated using three cancer cell lines, namely A172 (human brain glioblastoma), B16F10 (murine melanoma) and MDA-MB-231 (human breast adenocarcinoma), as well as a healthy control cell line of HUVEC (human umbilical vein endothelial cells). The results show that 9 out of 11 newly synthesized compounds demonstrated similar antiproliferative action against the B16F10 cell line to the reference drug, and three of these compounds surpassed it. To the best of our knowledge, this study is the first to demonstrate dual inhibitory action of carborane–thiazole derivatives against both tyrosinase and 11β-HSD1. Therefore, it represents the first step towards the simultaneous treatment of melanoma and comorbid diseases such as type II diabetes mellitus. Full article
Show Figures

Figure 1

Figure 1
<p>Lineweaver–Burk plots for tyrosinase inhibition for compounds <b>4f</b> (<b>a</b>) and <b>4h</b> (<b>b</b>).</p>
Full article ">Figure 2
<p>The correlation between the binding energies calculated during docking studies for a series of ligands to the two studied proteins and the experimental values of ln(IC<sub>50</sub>). The vertical bars denote the standard deviation values calculated with respect to various crystal structures of the same protein. The ligands containing the <span class="html-italic">o</span>-carborane moiety are represented by red dots whereas kojic acid and ascorbic acid by green dots. The corresponding results for carbenoxolone are not shown due to magnitude mismatch.</p>
Full article ">Figure 3
<p>(<b>A</b>) The location of the <b>4h</b> ligand molecule bound to the tyrosinase (PDB:2y9x) structure and found in the docking procedure. The ligand molecule is shown as thick sticks, whereas all the closest amino acid residues (of a distance no larger than 0.38 nm) are represented by thin sticks. Copper ions are shown as orange balls. The description of the interaction types is given in the text. (<b>B</b>) The superposition of the most favorable poses of all ligands in the tyrosinase-containing complexes. Compounds <b>4i</b> and <b>4j</b> exhibit the orientation in the binding cavity, which is notably different from those of remaining ligands; compounds <b>4f</b>, <b>4g</b>, and <b>4h</b> are located in the closest vicinity of Cu ions present in the protein structure. (<b>C</b>) The location of the <b>4a</b> ligand molecule bound to the 11β-HSD1 (PDB:3czr) structure and found in the docking procedure. The NADP<sup>+</sup> molecule is highlighted in orange. Other details as in (<b>A</b>). (<b>D</b>) The superposition of the most favorable poses of ligands <b>4a</b>–<b>4g</b> in the 11β-HSD1-containing complexes.</p>
Full article ">Scheme 1
<p>Synthesis of the carborane–thiazoles <b>4a</b>–<b>4k</b>.</p>
Full article ">
12 pages, 764 KiB  
Article
Survival of Patients with Metastatic Melanoma Treated with Ipilimumab after PD-1 Inhibitors: A Single-Center Real-World Study
by Sofia Verkhovskaia, Rosa Falcone, Francesca Romana Di Pietro, Maria Luigia Carbone, Tonia Samela, Marie Perez, Giulia Poti, Maria Francesca Morelli, Albina Rita Zappalà, Zorika Christiana Di Rocco, Roberto Morese, Gabriele Piesco, Paolo Chesi, Paolo Marchetti, Damiano Abeni, Cristina Maria Failla and Federica De Galitiis
Cancers 2024, 16(19), 3397; https://doi.org/10.3390/cancers16193397 - 4 Oct 2024
Viewed by 304
Abstract
Background: When monotherapy with PD-1 inhibitors in metastatic melanoma fails, there are currently no standard second-line choices. In case of the unavailability of clinical trials, ipilimumab represents a possible alternative treatment. Methods: We collected data of 44 patients who received ipilimumab after the [...] Read more.
Background: When monotherapy with PD-1 inhibitors in metastatic melanoma fails, there are currently no standard second-line choices. In case of the unavailability of clinical trials, ipilimumab represents a possible alternative treatment. Methods: We collected data of 44 patients who received ipilimumab after the failure of PD-1 inhibitors from July 2017 to May 2023 at our Institute. Overall survival (OS), progression-free survival (PFS), and post-progression survival (PPS) based on BRAF or NRAS mutation status, sex, and the presence of brain metastases were estimated using the Kaplan–Meier method. Cox regression was used to evaluate independence in multivariate analysis. The objective response rate (ORR) was estimated based on RECIST 1.1. Results: Among the 44 patients enrolled in this study, 28 BRAF-wildtype, 9 BRAF-mutated, and 7 NRAS-mutated patients were identified. OS analysis showed a significant difference between wildtype and BRAF- or NRAS-mutated patients: 23.2 months vs 5.3 and 4.59, respectively, p = 0.017. The presence of brain metastases and BRAF or NRAS mutation were independent factors for mortality in multivariate analysis. Conclusions: In case of failure to enroll patients in innovative clinical trials, second-line ipilimumab still represents an effective therapy in patients with metastatic wildtype melanoma and in the absence of brain metastases. Full article
14 pages, 2060 KiB  
Article
Unraveling the Role of JMJD1B in Genome Stability and the Malignancy of Melanomas
by Perla Cruz, Diego Peña-Lopez, Diego Figueroa, Isidora Riobó, Vincenzo Benedetti, Francisco Saavedra, Claudia Espinoza-Arratia, Thelma M. Escobar, Alvaro Lladser and Alejandra Loyola
Int. J. Mol. Sci. 2024, 25(19), 10689; https://doi.org/10.3390/ijms251910689 - 4 Oct 2024
Viewed by 218
Abstract
Genome instability relies on preserving the chromatin structure, with any histone imbalances threating DNA integrity. Histone synthesis occurs in the cytoplasm, followed by a maturation process before their nuclear translocation. This maturation involves protein folding and the establishment of post-translational modifications. Disruptions in [...] Read more.
Genome instability relies on preserving the chromatin structure, with any histone imbalances threating DNA integrity. Histone synthesis occurs in the cytoplasm, followed by a maturation process before their nuclear translocation. This maturation involves protein folding and the establishment of post-translational modifications. Disruptions in this pathway hinder chromatin assembly and contribute to genome instability. JMJD1B, a histone demethylase, not only regulates gene expression but also ensures a proper supply of histones H3 and H4 for the chromatin assembly. Reduced JMJD1B levels lead to the cytoplasmic accumulation of histones, causing defects in the chromatin assembly and resulting in DNA damage. To investigate the role of JMJD1B in regulating genome stability and the malignancy of melanoma tumors, we used a JMJD1B/KDM3B knockout in B16F10 mouse melanoma cells to perform tumorigenic and genome instability assays. Additionally, we analyzed the transcriptomic data of human cutaneous melanoma tumors. Our results show the enhanced tumorigenic properties of JMJD1B knockout melanoma cells both in vitro and in vivo. The γH2AX staining, Micrococcal Nuclease sensitivity, and comet assays demonstrated increased DNA damage and genome instability. The JMJD1B expression in human melanoma tumors correlates with a lower mutational burden and fewer oncogenic driver mutations. Our findings highlight JMJD1B’s role in maintaining genome integrity by ensuring a proper histone supply to the nucleus, expanding its function beyond gene expression regulation. JMJD1B emerges as a crucial player in preserving genome stability and the development of melanoma, with a potential role as a safeguard against oncogenic mutations. Full article
(This article belongs to the Special Issue Molecular Research on Epigenetic Modifications)
Show Figures

Figure 1

Figure 1
<p><b>Accumulation of histones H3 and H4 and NASP in cytosolic extracts upon JMJD1B depletion.</b> (<b>A</b>) Scheme of the mouse JMJD1B/KDM3B knockout using the CRISPR-Cas9 technique. (<b>B</b>) Western blot analysis of 5 µg of total cell extract from the wild type (WT) and clones obtained from the JMJD1B gene knockout using the CRISPR/Cas9 technique. The total cell extract from the HeLa cells served as a control for the JMJD1B signal. From these clones, we selected three (clones #16, #18, and #19), based on their growth characteristics, phenotypic traits, and lack of JMJD1B expression, as determined by Western blot analyses, for further investigation. Clones #7, #12, #13, #17, and #22 were excluded from the analysis due to poor growth or abnormal cell morphology. (<b>C</b>) Western blot analysis of 5 µg of cytosolic (S100) and nuclear (NE) extracts from the wild type (WT) and clones derived from the B16F10 JMJD1B KO cells, blotted as indicated. The number below the blots represent the fold changes relative to the wild type. The bands were quantified, normalized to β-actin, and then expressed as fold changes compared to the wild type. Of note is that the band below tNASP is only observed in the nuclear extract derived from mouse cells.</p>
Full article ">Figure 2
<p><b>The tumorigenic properties of B16F10 cells are potentiated upon JMJD1B depletion.</b> (<b>A</b>) Proliferation of B16F10 wild-type (WT) and JMJD1B knockout cells measured with the CCK-8 assay in three–four independent experiments. The doubling time for each B16F10 cell is shown in parentheses, calculated with a non-linear fit, giving R squared values of 0.94 (WT), 0.92 (#16), 0.98 (#18), and 0.92 (#19). The significance was determined by a two-way ANOVA. **, <span class="html-italic">p</span> value of &lt;0.01. (<b>B</b>) Soft agar plaque assay. Representative images of the soft agar plaques and graphs displaying pairwise comparisons of the total number of colonies obtained from experiments using B16F10 wild-type (WT) cells and JMJD1B knockout clones #16, #18, and #19 are shown. Images were captured after 17 days and seven to ten biological replicates were used. The significance was determined by a paired <span class="html-italic">t</span>-test. **, <span class="html-italic">p</span> value of &lt;0.01; and ***, <span class="html-italic">p</span> value of &lt;0.001. (<b>C</b>) Wound healing assay. Representative image of three independent assays and graph displaying cell migration expressed as the percentage of migration. Wild-type and JMJD1B knockout B16F10 cells were plated on plastic dishes, wounded with a P200 pipette tip, and photographed at 0 and 8 h. The significance was determined by a <span class="html-italic">t</span>-test. **, <span class="html-italic">p</span> value of &lt;0.01; and ***, <span class="html-italic">p</span> value of &lt;0.001. (<b>D</b>) In vivo tumor growth. RAG1 KO C5BL/6 mice were intradermally injected with 1 × 10<sup>6</sup> wild-type (WT) and JMJD1B knockout clones #18 and #19 B16F10 melanoma cell lines. Tumor growth was evaluated in the subsequent days. The graph presents the average tumor growth for each group, showing the mean ± SEM of the tumor sizes from one experiment, with n = 5 mice per group. The significance was determined by a two-way ANOVA and Bonferroni’s post hoc test. ***, <span class="html-italic">p</span> value ≤ 0.001; and ****, <span class="html-italic">p</span> value ≤ 0.0001.</p>
Full article ">Figure 3
<p><b>Depletion of JMJD1B increases genomic instability in B16F10 cells.</b> (<b>A</b>) Western blot analysis. Western blot analysis of 20 µg of total cell extract derived from wild-type (WT) and JMJD1B knockout clones #16, #18, and #19, blotted as described. (<b>B</b>) Comet assay. Comet assay of B16F10 cells, wild-type (WT) cells, and JMJD1B knockout clones #16, #18, and #19. On the left, there is a representative stained DNA image after electrophoresis of the lysed cells. On the right, a violin graph represents the quantification of DNA found in the tail of the comets. For this assay, between 500 and 1500 cells were quantified per condition from three independent experiments. The data were analyzed with a two-way ANOVA test. **, <span class="html-italic">p</span> value of &lt;0.01; and ****, <span class="html-italic">p</span> value of &lt;0.0001. (<b>C</b>) Micrococcal Nuclease (MNase) digestion. MNase digestion of wild-type (WT) B16F10 cells and JMJD1B knockout clones #16, #18, and #19 was performed for the indicated times. The DNA extracted from the digested cells was analyzed on 1.5% agarose gels and visualized with a GelRed stain and UV illumination. Molecular weight (MW) markers were run between the clones, and their corresponding sizes are indicated on the left side of the gel. The graphs represent densitometric scans of the electrophoretic patterns of the MNase-digested DNA at 0 and 20 min from all the cells. The image is a representative experiment of three independent assays.</p>
Full article ">Figure 4
<p>KDM3B/JMJD1B expression associates with reduced mutational burden in the genome and oncogenic drivers in human melanoma tumors. (<b>A</b>) Spearman’s rank correlation between KDM3B expression and overall mutational load in human melanoma tumors. (<b>B</b>) Volcano plot for the odds ratio and <span class="html-italic">p</span>-value analyzed by logistic regression showing the expression of KDM3B and the probability of SNV mutations for a majority set of tumor suppressor genes and proto-oncogenes. Genes associated with lower and higher mutation risks are identified in the upper left (in green) and upper right (in red) quadrants, respectively. (<b>C</b>) Spearman’s rank correlation between KDM3B and TP53 expression.</p>
Full article ">Figure 5
<p><b>Model of the role of JMJD1B as a regulator of chromosome stability.</b> Through a yet-unknown mechanism, the absence of JMJD1B leads to an increase in the protein levels of the histone chaperone tNASP, which accumulates newly synthesized histones H3/H4 in the cytoplasm, affecting the histone supply to the nucleus and leading to genome instability. This process highlights the potential role of JMJD1B in cancer progression beyond its established function in gene expression regulation: genome instability that correlates with oncogenic driver mutations, including p53.</p>
Full article ">
20 pages, 3133 KiB  
Article
Utilizing Adenovirus Knob Proteins as Carriers in Cancer Gene Therapy Amidst the Presence of Anti-Knob Antibodies
by Naoya Koizumi, Takamasa Hirai, Junpei Kano, Anna Sato, Yurika Suzuki, Arisa Sasaki, Tetsuya Nomura and Naoki Utoguchi
Int. J. Mol. Sci. 2024, 25(19), 10679; https://doi.org/10.3390/ijms251910679 - 3 Oct 2024
Viewed by 375
Abstract
Numerous gene therapy drugs for cancer have received global approval, yet their efficacy against solid tumors remains inadequate. Our previous research indicated that the fiber protein, a component of the adenovirus capsid, can propagate from infected cells to neighboring cells that express the [...] Read more.
Numerous gene therapy drugs for cancer have received global approval, yet their efficacy against solid tumors remains inadequate. Our previous research indicated that the fiber protein, a component of the adenovirus capsid, can propagate from infected cells to neighboring cells that express the adenovirus receptor. We hypothesize that merging this fiber protein with an anti-cancer protein could enable the anti-cancer protein to disseminate around the transfected cells, presenting a novel approach to cancer gene therapy. In our study, we discovered that the knob region of the adenovirus type 5 fiber protein is the smallest unit capable of spreading to adjacent cells in a receptor-specific manner. We also showed that the recombinant knob protein infiltrates cells after dispersing to surrounding cells. To assess the potential of the knob protein to augment gene therapy for solid tumors in mice, we expressed a fusion gene of the A subunit of cytotoxic cholera toxin and the knob region in mouse tumors. We found that this fusion protein only inhibited tumor growth in receptor-expressing mouse melanomas, and this inhibitory effect persisted even in mice with anti-knob antibodies. Our study’s findings propose a novel cancer gene therapy strategy that enhances therapeutic effects by specifically delivering therapeutic proteins, expressed from in vivo administered genes, to target molecules. This outcome offers a fresh perspective on gene therapy for solid cancers, and we anticipate that knob proteins will serve as a platform for this method. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Analysis of adenovirus knob region distribution on 293T Cells. (<b>A</b>,<b>B</b>) Schematic representation of adenovirus (Ad) and Ad fiber protein. (<b>C</b>) Several Ad capsid proteins expressed via plasmid. (<b>D</b>) Flow-cytometry analysis of Ad capsid proteins in 293T cells transfected with plasmids. Twenty-four hours post-transfection, cells were harvested using 2 mM EDTA/PBS and stained with anti-DDDDK monoclonal antibody. (<b>E</b>) Western blot analysis of Ad capsid proteins in 293T cells transfected with plasmids. Twenty-four hours post-transfection, cells were harvested using 2 mM EDTA/PBS and disrupted via ultrasonication. After centrifugation to collect the supernatant, 10 mg of the supernatant was separated on a 15% SDS-polyacrylamide gel, and the Ad capsid proteins were analyzed by Western blotting using an anti-DDDDK monoclonal antibody as outlined in <a href="#sec4-ijms-25-10679" class="html-sec">Section 4</a>.</p>
Full article ">Figure 2
<p>CAR-specific cell-to-cell spread of knob protein (<b>A</b>) Schematic representation of experimental methods. (<b>B</b>–<b>D</b>) Flow-cytometry analysis of type 5 Ad knob protein (Ad5knob) on various co-cultured GFP-expressing cells. Twenty-four hours post-transfection of 293T cells, cells were harvested using 2 mM EDTA/PBS and co-cultured with either CAR-positive or CAR-negative cells (<b>B</b>), CAR-knockdown 293T cells (<b>C</b>), or CAR-expressed SF293 cells (<b>D</b>). After 24 h of co-culture, cells were harvested using 2 mM EDTA/PBS and analyzed by flow-cytometry using an anti-DDDDK monoclonal antibody as outlined in <a href="#sec4-ijms-25-10679" class="html-sec">Section 4</a>. Data are expressed as mean ± S.D. (n = 3).</p>
Full article ">Figure 3
<p>Investigation of cell-to-cell spread of recombinant knob protein. (<b>A</b>) Western blot analysis of recombinant Ad5knob (rAd5knob) protein derived from plasmid-transfected 293T cells. Twenty-four hours post-transfection, cells were harvested using 2 mM EDTA/PBS and disrupted via ultrasonication. Following supernatant collection by centrifugation, rAd5knob was purified using an anti-DDDDK-tag specific gel column. The rAd5knob was then separated on a 15% SDS- or native-polyacrylamide gel and analyzed by Western blotting using an anti-DDDDK monoclonal antibody as outlined in <a href="#sec4-ijms-25-10679" class="html-sec">Section 4</a>. (<b>B</b>) Flow-cytometry analysis of rAd5knob on co-cultured 293T-GFP cells. One hour post-co-culture of rAd5knob-binding 293T cells and 293T-GFP cells, cells were analyzed by flow-cytometry using an anti-DDDDK monoclonal antibody as described in <a href="#sec4-ijms-25-10679" class="html-sec">Section 4</a>. Mean fluorescence intensity data are expressed as mean (n = 3). (<b>C</b>) Immunohistological staining of rAd5knob in 293T cells. One hour post-Ad5knob binding to 293T cells on ice, cells were cultured for 15 or 30 min at 37 degrees. Cells were stained with anti-DDDDK monoclonal antibody and Hoechst 33342 and observed with a confocal laser microscope.</p>
Full article ">Figure 4
<p>Characteristics of cholera toxin A-subunit and knob fusion protein in vitro. (<b>A</b>) Schematic representation of plasmids expressing Ad5knob fusion proteins. The A subunit of cholera toxin (NCTXA) retains its cytotoxic properties. The efficiency of gene transfer for all plasmids can be confirmed by the quantity of GFP co-expressed with IRES sequences. (<b>B</b>) Western blot analysis of Ad knob fusion proteins in 293T cells transfected with plasmids. The Ad knob fusion proteins were analyzed by Western blotting using an anti-DDDDK monoclonal antibody and an anti-cholera toxin monoclonal antibody as outlined in <a href="#sec4-ijms-25-10679" class="html-sec">Section 4</a>. (<b>C</b>) Flow-cytometry analysis of Ad knob fusion proteins on co-cultured cells. Twenty-four hours post-transfection of 293T cells with plasmids, cells were co-cultured with either B16-BL6-hCAR or B16-BL6 cells. After 24 h of co-culture, cells were harvested using 2 mM EDTA/PBS and analyzed by flow-cytometry using an anti-DDDDK monoclonal antibody as described in <a href="#sec4-ijms-25-10679" class="html-sec">Section 4</a>. Data are expressed as mean ± S.D. (n = 3). (<b>D</b>) Cell proliferation analysis of NCTXA and Ad knob fusion proteins in co-cultured cells. Twenty-four hours post-transfection of 293T cells with plasmids, cells were cultured alone or co-cultured with either B16-BL6-hCAR or B16-BL6 cells. After 48 h of co-culture, cells were harvested using 2 mM EDTA/PBS, and the number of cells was counted. Statistical analysis was conducted using two-way analysis of variance and the Student–Newman–Keuls post hoc test (* <span class="html-italic">p</span> &lt; 0.05 vs. control).</p>
Full article ">Figure 5
<p>Inhibition of tumor growth in tumor-bearing mice by transfection of genes encoding NCTXA and Ad knob fusion proteins. (<b>A</b>) Timeline of mouse experiments. (<b>B</b>) Inhibition of tumor growth in tumor-bearing mice by intratumoral administration of gene-transfected B16BL6-hCAR cells. Eight days after subcutaneous transplantation of B16BL6-hCAR cells into mice, gene-transfected B16BL6-hCAR cells were intratumorally administered. Subsequently, cells were administered two more times, and tumor diameter and mouse weight were measured. Statistical analysis was conducted using two-way analysis of variance and the Student–Newman–Keuls post hoc test (* <span class="html-italic">p</span> &lt; 0.05 vs. control). (<b>C</b>) Inhibition of tumor growth in tumor-bearing mice by direct gene transfection via intratumoral administration. Seven days after subcutaneous transplantation of B16BL6-hCAR or B16BL6 cells into mice, transfection reagents and plasmids were intratumorally administered. Subsequently, the reagent and plasmids were administered once more, and tumor diameter and weight were measured. Statistical analysis was conducted using two-way analysis of variance and the Student–Newman–Keuls post hoc test (* <span class="html-italic">p</span> &lt; 0.05 vs. NCTXA). (<b>D</b>) Inhibition of tumor growth in tumor-bearing mice by intratumoral administration of gene-transfected B16BL6-hCAR cells. Seven days after subcutaneous transplantation of B16BL6-hCAR cells into mice, gene-transfected B16BL6-hCAR or B16BL6 cells were intratumorally administered. Subsequently, cells were administered once more, and tumor diameter and weight were measured. Statistical analysis was conducted using two-way analysis of variance and the Student–Newman–Keuls post hoc test (* <span class="html-italic">p</span> &lt; 0.05 vs. NCTXA-Ad5knob expressing B16BL6). (<b>E</b>) Preparation of paraffin sections from tumors. Eight days after subcutaneous transplantation of B16BL6-hCAR cells into mice, gene-transfected B16BL6-hCAR cells were intratumorally administered. Two days later, tumor tissue was recovered. Each section was stained with H&amp;E and anti-caspase-3 monoclonal antibody. Data represent means ± SD of three to six mice.</p>
Full article ">Figure 6
<p>Tumor growth inhibition in Ad knob immunized mice bearing tumors, achieved by transfecting genes encoding NCTXA and Ad knob fusion proteins (<b>A</b>) Timeline for antigen administration in the mouse experiments. (<b>B</b>) Serum antibody levels post-four intraperitoneal administrations of antigens to mice. ELISA was used to measure the antibody levels for Ad5knob or OVA in the serum measured by ELISA. (<b>C</b>) The inhibition of tumor growth in antigen-immunized mice bearing tumors when B16BL6-hCAR cells, transfected with the gene, were administered intratumorally. Eight days post-subcutaneous transplantation of B16BL6-hCAR cells into mice, these gene-transfected cells were administered into the tumor. The tumor diameter and mice weight were then measured. <sup>#</sup> Mice with high anti-Ad5knob antibody titers. Data represent the means ± SD of three to six mice. Statistical analysis was conducted using two-way analysis of variance and the Student–Newman–Keuls post hoc test (* <span class="html-italic">p</span> &lt; 0.05 vs. control).</p>
Full article ">
10 pages, 268 KiB  
Study Protocol
Effect of Telenursing on Supportive Care Needs in Patients with Melanoma and Lung Cancer on Targeted Therapies: A Randomised Controlled Trial Study Protocol
by Aurora De Leo, Gloria Liquori, Alessandro Spano, Nicolò Panattoni, Sara Dionisi, Laura Iacorossi, Noemi Giannetta, Irene Terrenato, Emanuele Di Simone, Marco Di Muzio and Fabrizio Petrone
Methods Protoc. 2024, 7(5), 78; https://doi.org/10.3390/mps7050078 - 3 Oct 2024
Viewed by 278
Abstract
Background: Telenursing comprises a set of tools and interventions enabling nurses to provide remote care. This study aims to assess the impact of telenursing interventions on the supportive care needs of patients with melanoma and lung cancer who are receiving targeted therapies. [...] Read more.
Background: Telenursing comprises a set of tools and interventions enabling nurses to provide remote care. This study aims to assess the impact of telenursing interventions on the supportive care needs of patients with melanoma and lung cancer who are receiving targeted therapies. Methods: This six-month monocentric, double-arm, randomised, controlled trial study protocol will assess the effect of telenursing on the supportive care needs (primary outcome) in 40 patients (20 in each group) after one month. The secondary outcomes will be monitored at baseline, one, three and six months: supportive care needs (at three and six months), therapeutic adherence, quality of life, usability and satisfaction, performance status, patient-reported outcomes and main adverse events. The SPIRIT guidelines will be used for the reporting. Results: The results from this trial will assess the impact of a telenursing intervention on cancer care. Conclusions: This trial could be a starting point for more extensive studies on telenursing interventions to promote nurses’ skills, as well as the quality and safety of care in patients with cancer, highlighting the impact of more outstanding nursing contributions on cancer care. Trial and Protocol Registration: The study protocol was approved by the relevant Italian Ethics Committee Lazio Area 5 (RS1851/23, 2773; 6 September 2023) and was registered on ClinicalTrials.gov (trial registry number NCT06254196). Full article
(This article belongs to the Section Public Health Research)
19 pages, 4379 KiB  
Case Report
Comprehensive Literature Review on Melanoma of Unknown Primary Site Triggered by an Intriguing Case Report
by Eliza-Maria Bordeanu-Diaconescu, Andrei Cretu, Andreea Grosu-Bularda, Mihaela-Cristina Andrei, Florin-Vlad Hodea, Catalina-Stefania Dumitru, Valentin Enache, Cosmin-Antoniu Creanga, Ioan Lascar and Cristian-Sorin Hariga
Diagnostics 2024, 14(19), 2210; https://doi.org/10.3390/diagnostics14192210 - 3 Oct 2024
Viewed by 312
Abstract
Melanoma is one of the most aggressive forms of skin cancer. While most melanomas have a discernible primary site, a small subset, approximately 3.2%, present as a metastatic disease without an identifiable primary origin, a condition known as melanoma of unknown primary (MUP). [...] Read more.
Melanoma is one of the most aggressive forms of skin cancer. While most melanomas have a discernible primary site, a small subset, approximately 3.2%, present as a metastatic disease without an identifiable primary origin, a condition known as melanoma of unknown primary (MUP). Unusual cases of primary melanoma have also been previously reported in the respiratory, gastrointestinal, and urogenital tracts. MUP typically is found in lymph nodes, subcutaneous sites, and visceral organs, with hypotheses about its origin including spontaneous primary tumor regression and ectopic melanocytes. MUP presents unique challenges in diagnosis and treatment due to the absence of a detectable primary tumor. Understanding its genetic and molecular features, epidemiology, prognostic factors, and treatment options is crucial for optimizing patient care and outcomes in this subset of melanoma patients. We conducted an extensive literature review triggered by a case report of a patient with suspected MUP. A 51-year-old woman was transferred from another hospital where an incision was performed for a suspected superinfected hematoma of the left thigh. Since the patient showed high leukocytosis and redness and swelling of the thigh, local debridement, drainage, and excisional biopsy of the tumor mass were performed in our unit in the emergency setting, and the tumor was taken for histopathology evaluation. Intraoperatively, the mass appeared nonspecific. The permanent histopathology report established a diagnosis of melanoma, with tumor proliferation also involving lymphoid tissue, and despite broad clinical and imagistic assessments, the primary melanoma could not be identified. Clinicians must be aware of the varied clinical manifestations of malignant melanoma, especially in cases of occult melanoma where the primary site is not evident. Full article
(This article belongs to the Section Pathology and Molecular Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Aspect of thigh at presentation in the emergency department.</p>
Full article ">Figure 2
<p>CT scan of the lower left limb showing a large mass in the upper thigh.</p>
Full article ">Figure 3
<p>Intraoperative aspect of the mass.</p>
Full article ">Figure 4
<p>Histopathology: showing samples stained with hematoxylin and eosin under magnification of 20× (<b>A</b>), positive immunostaining for SOX10 (<b>B</b>) consistent with malignant melanoma.</p>
Full article ">
16 pages, 756 KiB  
Review
Plaque Radiotherapy for Ocular Melanoma
by George Naveen Thomas, I-Ling Chou and Lingam Gopal
Cancers 2024, 16(19), 3386; https://doi.org/10.3390/cancers16193386 - 3 Oct 2024
Viewed by 259
Abstract
Plaque radiotherapy is an effective treatment modality for medium-sized ocular tumors such as uveal melanoma. The authors review the available literature and concisely summarize the current state of the art of ophthalmic plaque brachytherapy. The choice of radioisotope, which includes Ruthenium-106 and Iodine-125, [...] Read more.
Plaque radiotherapy is an effective treatment modality for medium-sized ocular tumors such as uveal melanoma. The authors review the available literature and concisely summarize the current state of the art of ophthalmic plaque brachytherapy. The choice of radioisotope, which includes Ruthenium-106 and Iodine-125, depends on the intended treatment duration, tumor characteristics, and side effect profiles. Ophthalmic plaques may be customized to allow for the delivery of a precise radiation dose by adjusting seed placement and plaque shape to minimize collateral tissue radiation. High dose rate (HDR) brachytherapy, using beta (e.g., Yttrium-90) and photon-emitting sources (e.g., Ytterbium-169, Selenium-75), allows for rapid radiation dose delivery, which typically lasts minutes, compared to multiple days with low-dose plaque brachytherapy. The efficacy of Ruthenium-106 brachytherapy for uveal melanoma varies widely, with reported local control rates between 59.0% and 98.0%. Factors influencing outcomes include tumor size, thickness, anatomical location, and radiation dose at the tumor apex, with larger and thicker tumors potentially exhibiting poorer response and a higher rate of complications. Plaque brachytherapy is effective for selected tumors, particularly uveal melanoma, providing comparable survival rates to enucleation for medium-sized tumors. The complications of plaque brachytherapy are well described, and many of these are treatable. Full article
(This article belongs to the Special Issue Advances in Brachytherapy in the Treatment of Tumors)
Show Figures

Figure 1

Figure 1
<p>Plaque insertion surgery. (<b>A</b>) Tagging the lateral rectus muscle; (<b>B</b>) disinserting the lateral rectus muscle; (<b>C</b>) passing a traction suture through the stump of the insertion of lateral rectus muscle; (<b>D</b>) passing a scleral suture at the intended location of the anterior eyelet of plaque; (<b>E</b>) placement of the gold plaque with I-125 seeds; (<b>F</b>) anchoring the plaque to the sclera; (<b>G</b>) reattaching the detached lateral rectus; (<b>H</b>) conjunctival closure.</p>
Full article ">
11 pages, 1352 KiB  
Article
Anti-Melanogenic Effects of a Polysaccharide Isolated from Undaria pinnatifida Sporophyll Extracts
by Jae-Hoon Kim, Jeong-Heon Kim, Jae-Hoon Lee, Su-Jin Eom, Nam-Hyouck Lee, Saerom Lee, Tae-Gyu Lim, Kyung-Mo Song and Min-Cheol Kang
Int. J. Mol. Sci. 2024, 25(19), 10624; https://doi.org/10.3390/ijms251910624 - 2 Oct 2024
Viewed by 199
Abstract
Undaria pinnatifida is a temperate brown alga known to exert free radical-scavenging and anti-inflammatory effects. In this study, we investigated the skin-whitening effects of U. pinnatifida sporophyll extracts (UPEs) in α-melanocyte-stimulating hormone (α-MSH)-stimulated B16F10 melanoma cells. The crude polysaccharide fraction (UPF) was obtained [...] Read more.
Undaria pinnatifida is a temperate brown alga known to exert free radical-scavenging and anti-inflammatory effects. In this study, we investigated the skin-whitening effects of U. pinnatifida sporophyll extracts (UPEs) in α-melanocyte-stimulating hormone (α-MSH)-stimulated B16F10 melanoma cells. The crude polysaccharide fraction (UPF) was obtained via ethanol precipitation. Four polysaccharide fractions (UPF1–4) were isolated and purified using ion-exchange column chromatography, and their anti-melanogenic activity was evaluated. UPF3 exhibited the highest anti-melanogenic activity, showing the highest sulfate (39.79%), fucose (143 μg/mg), and galactose (208 μg/mg) contents. UPF3 significantly inhibited intracellular tyrosinase activity in B16F10 cells. We also evaluated the melanogenic signaling pathway to determine the mechanism of action of UPF3 in melanongenesis. UPF3 reduced the expression of tyrosinase-related protein-1 (TRP-1), tyrosinase-related protein-2 (TRP-2), and tyrosinase, which play important roles in melanin production. Therefore, UPF3 has high potential for use in skin-whitening functional pharmaceuticals and cosmetics. Full article
Show Figures

Figure 1

Figure 1
<p>DEAE–cellulose chromatogram of polysaccharides from <span class="html-italic">Undaria pinnatifida</span> sporophyll crude polysaccharides.</p>
Full article ">Figure 2
<p>Cytotoxicity of UPEs on B16F10 cells. * <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 the untreated control group.</p>
Full article ">Figure 3
<p>Melanogenic effects of UPEs on B16F10 cells. (<b>A</b>) Extracellular melanin contents of UPEs. (<b>B</b>) Intracellular melanin contents of UPF1 and UPF3. (<b>C</b>) Cellular tyrosinase activity of UPF1 and UPF3. (<b>D</b>) Extracellular melanin content image of UPF3. (<b>E</b>) Cellular tyrosinase activity determined by L-DOPA staining assay. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared to the untreated control group. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 compared to the α-MSH-treated group.</p>
Full article ">Figure 4
<p>Effects of UPF3 on melanogenesis-related signaling pathways as assessed by Western blot in B16F10 cells. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared to the untreated 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 to the α-MSH-treated group.</p>
Full article ">
18 pages, 5691 KiB  
Article
Evidence of Neutrophils and Neutrophil Extracellular Traps in Human NMSC with Regard to Clinical Risk Factors, Ulceration and CD8+ T Cell Infiltrate
by Linda-Maria Hildegard Moeller, Carsten Weishaupt and Fiona Schedel
Int. J. Mol. Sci. 2024, 25(19), 10620; https://doi.org/10.3390/ijms251910620 - 2 Oct 2024
Viewed by 272
Abstract
Non-melanoma skin cancers (NMSC), including basal cell carcinoma (BCC), cutaneous squamous cell carcinoma (cSCC), and Merkel cell carcinoma (MCC), are increasingly common and present significant healthcare challenges. Neutrophil extracellular traps (NETs), chromatin fibers expulsed by neutrophil granulocytes, can promote immunotherapy resistance via an [...] Read more.
Non-melanoma skin cancers (NMSC), including basal cell carcinoma (BCC), cutaneous squamous cell carcinoma (cSCC), and Merkel cell carcinoma (MCC), are increasingly common and present significant healthcare challenges. Neutrophil extracellular traps (NETs), chromatin fibers expulsed by neutrophil granulocytes, can promote immunotherapy resistance via an impairment of CD8+ T cell-mediated cytotoxicity. Here, to identify a potential therapeutic target, we investigate the expulsion of NETs and their relation to CD8+ T cell infiltration in NMSC. Immunofluorescence staining for neutrophils (CD15) and NETs (H3cit), as well as immunohistochemistry for cytotoxic T cells (CD8+) on human cSCCs (n = 24), BCCs (n = 17) and MCCs (n = 12), revealed a correlation between neutrophil infiltration and ulceration diameter in BCC and MCC, but not in cSCC. In BCC and cSCC, neutrophil infiltration also correlated with the cross-sectional area (CSA). NETs were not associated with established risk factors but with the presence of an ulceration, and, in cSCC, with abscess-like structures. CD8+ T cell infiltration was not reduced in tumors that were NET-positive nor in those with a denser neutrophil infiltration. This study is the first to report and characterize NETs in NMSC. Thus, it gives an incentive for further research in this relevant yet understudied topic. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Exemplary pictures of different neutrophil scores (N0–N3) assessed by immunofluorescence (IF) staining. CD15 (green) was used as a granulocyte marker, and DAPI (blue) served as counterstaining. Scale bars 200 µm. (<b>B</b>) Overview of the analyzed tumor specimens and the detection of neutrophils. (<b>C</b>) Representative H&amp;E as well as respective IF images of a cutaneous squamous cell carcinoma (cSCC), a basal cell carcinoma (BCC), and a Merkel cell carcinoma (MCC) (single neutrophils indicated by arrows). Several of the cSCCs showed a high infiltration, in part with a swarming behavior (dotted circle) which was exclusively present in cSCCs. Scale bars 100 µm.</p>
Full article ">Figure 2
<p>(<b>A</b>) (1) Representative H&amp;E staining of a BCC. Tumors were annotated manually on the scanned slides and the cross-sectional area (CSA) was calculated with the open-source software QuPath 0.4.3. (2) A non-ulcerated MCC. If an intact epidermis (arrowheads) could be distinguished throughout the whole sample, the specimen was labeled as non-ulcerated. (3) and (4) Overview and close-up of an ulcerated cSCC. Ulceration (yellow line) was defined as a discontinuity of the epidermis (start indicated by arrow). Scale bar 1000 µm. (<b>B</b>) In BCC and MCC the neutrophil score given in the immunofluorescence staining correlated significantly with the ulceration diameter. In cSCC, no correlation was found. (<b>C</b>) In cSCC and BCC, but not in MCC, the neutrophil infiltration correlated significantly with the CSA. Bars indicate mean + standard deviation (SD). r = Spearman’s <span class="html-italic">ρ</span>. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>(<b>A</b>) Exemplary pictures of different neutrophil extracellular traps (NET) scores (NET0–NET3) assessed by immunofluorescence staining. H3cit (red) indicates NET formation. DAPI (blue) served as counterstaining. NET0: A non-ulcerated BCC with no NETs. NET1: An ulcerated BCC with some infiltration by NETs. In BCCs, they typically could be distinguished in the scab (arrow) and in the ulceration zone. NET2: An ulcerated cSCC with medium NET infiltration. NET3: An ulcerated cSCC with extensive NET infiltration. Scale bar 200 µm. (<b>B</b>) Overview of the analyzed tumor specimens and the detection of NETs. (<b>C</b>) The presence of NETs was associated with a significantly larger ulceration diameter across all entities. Bars indicate mean + SD. (<b>D</b>) Representative immunofluorescence images of a cSCC with a neutrophil score of N3 and a NET score of NET3. CD15 (green) was used to stain granulocytes. Small images show CD15 (green) and H3cit (red) channels separately. Towards the ulceration, disintegrating granulocytes can be distinguished that, in part, expulse NETs (examples shown by arrowheads). Scale bars 100 µm. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>H&amp;E and the respective immunofluorescence (IF) images of exemplary stained cSCC and MCC metastases. CD15 (green) was used as a granulocyte marker, H3cit (red) visualized NETs, and DAPI (blue) served as counterstaining. Necrotic areas are surrounded by dotted lines. (<b>A</b>) A necrotic cutaneous cSCC metastasis. Scale bar 200 µm. Enlarged section: Neutrophils in the necrotic area expulse NETs (examples indicated by arrowheads). Scale bar 100 µm. (<b>B</b>) A necrotic lymph node, highly infiltrated by neutrophils (green), but without NETs. Scale bar 1 mm.</p>
Full article ">Figure 5
<p>Intra- and peritumoral CD8<sup>+</sup> T cells were counted via immunohistochemistry and the open-source Software QuPath 0.4.3. (<b>A</b>) The tumor (1, green area) was annotated manually and a 300 µm rim (2, red area) was defined as the peritumoral region. Positive cells (3, red dots) were counted with the Positive Cell Detection command. The CD8<sup>+</sup> T-cell density quotient was calculated by dividing the intratumoral CD8<sup>+</sup> T cell density [cells/mm<sup>2</sup>] by the peritumoral CD8<sup>+</sup> T cell density [cells/mm<sup>2</sup>]. Scale bar 200 µm. (<b>B</b>) In BCC and cSCC, significantly more CD8<sup>+</sup> T cells were detected in the peritumoral area than inside the tumor. (<b>C</b>,<b>D</b>) No difference in the CD8<sup>+</sup> T cell density quotient was seen in specimens that contained NETs and those that did not, nor in specimens with different neutrophil infiltration densities. **** indicates <span class="html-italic">p</span> &lt; 0.0001, ns = not significant.</p>
Full article ">
18 pages, 1660 KiB  
Article
Comparative Evaluation of the Phytochemical Composition of Fruits of Ten Haskap Berry (Lonicera caerulea var. kamtschatica Sevast.) Cultivars Grown in Poland
by Natalia Żurek, Stanisław Pluta, Łukasz Seliga, Sabina Lachowicz-Wiśniewska and Ireneusz Tomasz Kapusta
Agriculture 2024, 14(10), 1734; https://doi.org/10.3390/agriculture14101734 - 1 Oct 2024
Viewed by 609
Abstract
The aim of this study was to investigate the qualitative and quantitative fruit profiles of ten cultivars (cvs.) of haskap berry (Lonicera caerulea var. kamtschatica Sevast.) to determine their antioxidant activity (ABTS test, CUPRAC test, ability to capture superoxide (O2˙ [...] Read more.
The aim of this study was to investigate the qualitative and quantitative fruit profiles of ten cultivars (cvs.) of haskap berry (Lonicera caerulea var. kamtschatica Sevast.) to determine their antioxidant activity (ABTS test, CUPRAC test, ability to capture superoxide (O2˙) and hydroxyl radicals (OH˙)), cytotoxic activity (against cancer cell lines breast, MCF-7; colon, HT-29; and melanoma, SK-Mel-28) and physicochemical properties. Most of the selected cultivars had not previously been analyzed for these properties. A total of 19 polyphenolic compounds were identified in the fruits of the tested genotypes, with a quantitative range of 2166.3–3597.0 µg/g. The polyphenol profile was dominated by anthocyanins (90.0–92.4%), and the remaining classes occurred in the following order: phenolic acids > flavonols > flavan-3-ols. The highest concentrations of these polyphenol groups were found in the cultivars ‘Honeybee’, ‘Sinij Uties’ and ‘Usłada’. The fruits of these cultivars were also characterized by the highest antioxidant activity (546.6–683.5 µg/mL for O2˙ and 541.2–652.1 µg/mL for OH˙) and cytotoxic activity (103.6–649.2 µg/mL). The data obtained indicate that the fruits of the new haskap cultivars are a good source of bioactive compounds with possible health-promoting properties. Full article
Show Figures

Figure 1

Figure 1
<p>The content of individual classes of polyphenols (anthocyanins, phenolic acids, flavan-3-ols, flavonols) estimated by the UPLC-PDA-MS/MS method in fruits of ten haskap berry cultivars: <b>1</b>. ‘Boreal Beauty’; <b>2</b>. ‘Boreal Beast’; <b>3</b>. ‘Boreal Blizzard’; <b>4</b>. ‘Aurora’; <b>5</b>. ‘Honeybee’; <b>6</b>. ‘Vostorg’; <b>7</b>. ‘Jugana’; <b>8</b>. ‘Usłada’; <b>9</b>. ‘Lawina’; <b>10</b>. ‘Sinij Uties’. Values are presented as mean and SD. Measurements were made in 3 repetitions. Statistical analysis (a–c) was performed using Duncan’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Pearson’s correlation index for fruits of haskap berry cultivars, showing the strength of the relationship between variables such as antioxidant activity (ABTS, CUPRAC, O<sub>2</sub>˙<sup>−</sup>, OH˙), cytotoxic activity (MCF-7, HT-29, SK-Mel-28), classes of individual polyphenols assessed spectrophotometrically (TPC, TFC, TAC) and chromatographically (anthocyanins, other polyphenols), and compounds identified at the highest concentrations (cyanidin 3-<span class="html-italic">O</span>-glucoside, cyanidin 3,5-<span class="html-italic">O</span>-diglucoside, cyanidin 3-<span class="html-italic">O</span>-rutinoside, chlorogenic acid, quercetin 3-<span class="html-italic">O</span>-rutinoside, procyanidin dimer B-type). A red color indicates a positive correlation and blue indicates a negative correlation. Antioxidant and cytotoxic activity was mainly dependent on the content of anthocyanins, especially cyanidin 3-<span class="html-italic">O</span>-glucoside.</p>
Full article ">Figure 3
<p>Principal component analysis of data regarding the physicochemical properties, content of phenolic compounds and health-promoting activity of haskap berries. The cultivars ‘Honeybee’, ‘Sinij Uties’ and ‘Usłada’ were characterized by the highest values for antioxidant and cytotoxic activity and phenolic compound content.</p>
Full article ">
15 pages, 5885 KiB  
Article
Regulatory Role of IL6 in Immune-Related Adverse Events during Checkpoint Inhibitor Treatment in Melanoma
by Krishna P. Singh, Anuj Singh, Olaf Wolkenhauer and Shailendra Kumar Gupta
Int. J. Mol. Sci. 2024, 25(19), 10600; https://doi.org/10.3390/ijms251910600 - 1 Oct 2024
Viewed by 631
Abstract
The landscape of clinical management for metastatic melanoma (MM) and other solid tumors has been modernized by the advent of immune checkpoint inhibitors (ICI), including programmed cell death-1 (PD-1), programmed cell death-ligand 1 (PD-L1), and cytotoxic T lymphocyte antigen 4 (CTLA-4) inhibitors. While [...] Read more.
The landscape of clinical management for metastatic melanoma (MM) and other solid tumors has been modernized by the advent of immune checkpoint inhibitors (ICI), including programmed cell death-1 (PD-1), programmed cell death-ligand 1 (PD-L1), and cytotoxic T lymphocyte antigen 4 (CTLA-4) inhibitors. While these agents demonstrate efficacy in suppressing tumor growth, they also lead to immune-related adverse events (irAEs), resulting in the exacerbation of autoimmune diseases such as rheumatoid arthritis (RA), ulcerative colitis (UC), and Crohn’s disease (CD). The immune checkpoint inhibitors offer promising advancements in the treatment of melanoma and other cancers, but they also present significant challenges related to irAEs and autoimmune diseases. Ongoing research is crucial to better understand these challenges and develop strategies for mitigating adverse effects while maximizing therapeutic benefits. In this manuscript, we addressed this challenge using network-based approaches by constructing and analyzing the molecular and signaling networks associated with tumor-immune crosstalk. Our analysis revealed that IL6 is the key regulator responsible for irAEs during ICI therapies. Furthermore, we conducted an integrative network and molecular-level analysis, including virtual screening, of drug libraries, such as the Collection of Open Natural Products (COCONUT) and the Zinc15 FDA-approved library, to identify potential IL6 inhibitors. Subsequently, the compound amprenavir was identified as the best molecule that may disrupt essential interactions between IL6 and IL6R, which are responsible for initiating the signaling cascades underlying irAEs in ICI therapies. Full article
(This article belongs to the Special Issue Drug Discovery of Compounds by Structural Design)
Show Figures

Figure 1

Figure 1
<p>This workflow outlines the process for identifying a lead compound for melanoma and autoimmune disease. The methods utilized were enhanced with various filters. Initially, gene-related information for all diseases was obtained using DisGeNET. The common genes identified were then analyzed through a protein–protein interaction (PPI) molecular map using the STRING database. The resulting PPI network was further analyzed in Cytoscape for cluster identification with MCODE. The most promising cluster underwent enrichment analysis, and we used a network-based approach to identify the target. Virtual screening and molecular docking were employed to find the best compound. Finally, the stability of the lead compound (amprenavir) was assessed through a molecular dynamics (MDs) simulation.</p>
Full article ">Figure 2
<p>Venn diagram highlighting the overlapping genes between rheumatoid arthritis (RA), ulcerative colitis (UC), Crohn’s disease (CD), and melanoma metastasis (MM). A total of 132 genes were shared among all the disease phenotypes.</p>
Full article ">Figure 3
<p>A network of the 107 common genes associated with the investigated four disease phenotypes. The network was prepared using the String database, and the connections between the nodes were above the 0.7 confidence score cutoff.</p>
Full article ">Figure 4
<p>A network of top enriched pathways associated with the genes present in the best cluster was identified through the MCODE analysis. The enriched pathways are shown in the green rectangle boxes, the genes are shown as colored ovals, and the disease phenotypes (MM and autoimmune diseases) are shown as circular nodes. The impacts of the genes on the pathways (dashed lines) and their links to melanoma and the autoimmune disease phenotypes (dotted lines) are shown where the pointed arrowheads indicate ‘activation’ and the blunt-end arrowheads indicate ‘suppression’.</p>
Full article ">Figure 5
<p>Two-dimensional representation of the top two compounds (ZINC000003809192 and CNP0003038, respectively) which were extracted after virtual screening and molecular docking using the DS20222.</p>
Full article ">Figure 6
<p>The docked poses obtained from the HDOCK docking tool depicting the interactions between IL6R and IL6 with Amprenavir. In this illustration, IL6R is represented by two different color bases on separate chains (alpha in light green and beta in dark green), while IL6 is shown in blue. The first frame of figure (<b>a</b>) showcases the surface representation of IL6R, the IL6 proteins, and their interactions. Figure (<b>b</b>) showcases a surface representation of IL6R and the IL6 proteins with Amprenavir. Additionally, the frame provides a zoomed-in version of the surface, highlighting the interactions between IL6R, IL6, and Amprenavir.</p>
Full article ">Figure 7
<p>MD simulation analysis of the IL6-Amprenavir complex. (<b>a</b>) Hydrogen bonds and hydrophobic interactions between IL6 and Amprenavir are shown before and after the MD simulation. After the MD simulation, Amprenavir formed two additional hydrogen bonds with IL6 compared to the initial docked pose, while the hydrophobic bonds remained unchanged. The colored arrow indicates the nature of the bonds. All IL6 amino acid residues involved in the bond formation are labelled. (<b>b</b>) Root Mean Square Deviation (RMSD) graph of the IL6 from the docked complex over a simulation period of 50 nanoseconds (ns). (<b>c</b>) Root Mean Square Fluctuation (RMSF) graph of the IL6 interaction site associated with IL6R. The IL6 amino acid residues that directly interacted with IL6R are labelled. (<b>d</b>) Radius of gyration (Rg) graph of IL6 from the IL6-Amprenavir complex. The graph suggests that IL6 attained a more compact structure after binding with the drug.</p>
Full article ">
16 pages, 3844 KiB  
Article
Composition and Activities of Carpesium macrocephalum Franch. & Sav. Essential Oils
by Anna Wajs-Bonikowska, Janusz Malarz, Łukasz Szoka, Paweł Kwiatkowski and Anna Stojakowska
Molecules 2024, 29(19), 4658; https://doi.org/10.3390/molecules29194658 - 30 Sep 2024
Viewed by 254
Abstract
Carpesium macrocephalum, a species native to China, Korea, Japan, and Russia, has been used medicinally in the countries of its origin. Though mono- and sesquiterpenoids are known constituents of C. macrocephalum, the complete analysis of essential oils produced by the roots [...] Read more.
Carpesium macrocephalum, a species native to China, Korea, Japan, and Russia, has been used medicinally in the countries of its origin. Though mono- and sesquiterpenoids are known constituents of C. macrocephalum, the complete analysis of essential oils produced by the roots and aerial parts of the plant has not been published until now. The present study discloses considerable differences in the composition and cytotoxic activity of essential oils distilled from roots and shoots of C. macrocephalum. The GC-MS-FID analyses have led to the identification of 131 compounds in all, of which 114 were found in aerial parts and 110 in the roots of the plants. The essential oil distilled from shoots contained a mixture of nerol and thymol methyl ether (c. 26%), neryl isobutyrate (c. 12%) and linalool (c. 9%) as major constituents, whereas alantolactone (c. 29%), thymol methyl ether (c. 7%) and 2,5-dimethoxy-p-cymene (thymohydroquinone dimethyl ether, c. 7%) predominated in the essential oil obtained from the roots. The oils demonstrated weak antibacterial activity against Staphylococcus aureus and, at concentrations up to 2.08 mg/mL (oil from the aerial parts) and up to 3.38 mg/mL (oil from roots), were inactive against Gram-negative bacteria. The essential oil from the roots of the plant demonstrated strong but not selective cytotoxic activity. Full article
(This article belongs to the Special Issue Plant Bioactive Compounds in Pharmaceuticals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Cell viability after the treatment of human melanoma and normal skin cell lines with the essential oils from <span class="html-italic">Carpesium macrocephalum</span> for 24 h and 48 h: APEO (<b>A</b>) and REO (<b>B</b>). Data are reported as mean ± SD from three independent experiments. * <span class="html-italic">p</span> &lt; 0.05 compared to control group.</p>
Full article ">Figure 2
<p>Apoptosis of melanoma cells after treatment with APEO and REO for 48 h (<b>A</b>). Western blot analysis of caspase-8, caspase-9, Mcl-1, caspase-3, caspase-7, and PARP in melanoma cells after the treatment with APEO and REO for 48 h. Actin served as a control for protein loading (<b>B</b>). Data are presented as mean ± SD from three independent experiments. * <span class="html-italic">p</span> &lt; 0.05 compared to control group.</p>
Full article ">
19 pages, 8433 KiB  
Article
Mushroom against Cancer: Aqueous Extract of Fomitopsis betulina in Fight against Tumors
by Paulina Nowotarska, Maciej Janeczek and Benita Wiatrak
Nutrients 2024, 16(19), 3316; https://doi.org/10.3390/nu16193316 - 30 Sep 2024
Viewed by 365
Abstract
Background/Objectives: This study investigated the anticancer potential of an aqueous extract of the fungus Fomitopsis betulina. Methods: The study assessed the effect of the extract on nine cancer cell lines, including melanoma (LM-MEL-75), lung cancer (A549), and colorectal cancer (HT29, LoVo), and [...] Read more.
Background/Objectives: This study investigated the anticancer potential of an aqueous extract of the fungus Fomitopsis betulina. Methods: The study assessed the effect of the extract on nine cancer cell lines, including melanoma (LM-MEL-75), lung cancer (A549), and colorectal cancer (HT29, LoVo), and four normal cell lines. The cytotoxicity of the extract was evaluated using MTT, sulforhodamine-B (SRB), and clonogenic viability assays. Additionally, the study examined the effect of the extract on plant model organisms, garden cress (Lepidium sativum) and common onion (Allium cepa), to further investigate its biological activity. Results: The assays demonstrated selective cytotoxicity of the extract toward cancer cells, while sparing normal cells. The extract induced significant cytotoxic effects at lower concentrations in lung cancer, melanoma, and colon cancer cells, showing promise as a potential anticancer agent. The results also revealed that the extract inhibited seed germination and root growth, suggesting its potential to disrupt cell cycles and induce apoptosis. Conclusions: This study highlights the therapeutic potential of F. betulina and highlights the need for further research to identify the active ingredients and mechanisms underlying its anticancer effects. Full article
(This article belongs to the Special Issue Anticancer Activities of Dietary Phytochemicals)
Show Figures

Figure 1

Figure 1
<p>Mitochondrial activity after incubation with aqueous extract of <span class="html-italic">F. betulina</span>: (<b>A</b>) Normal cell lines; (<b>B</b>) Cancer cell lines. Data are presented as E/E<sub>0</sub> (where E—the average number of colonies in the tested extract concentration and E<sub>0</sub>—the average number in the control culture and standard deviation (SD); * statistically significant differences in mitochondrial activity (<span class="html-italic">p</span> &lt; 0.05) when comparing treated versus control groups, with error bars representing standard deviation.</p>
Full article ">Figure 2
<p>Cell viability after incubation with an aqueous extract of <span class="html-italic">F. betulina</span>: (<b>A</b>) Normal cell lines; (<b>B</b>) Cancer cell lines. Data are presented as E/E<sub>0</sub> (where E—the average number of colonies in the tested extract concentration and E<sub>0</sub>—the average number in the control culture and standard deviation (SD); * statistically significant differences in cell viability (<span class="html-italic">p</span> &lt; 0.05) when comparing treated versus control groups, with error bars representing standard deviation.</p>
Full article ">Figure 2 Cont.
<p>Cell viability after incubation with an aqueous extract of <span class="html-italic">F. betulina</span>: (<b>A</b>) Normal cell lines; (<b>B</b>) Cancer cell lines. Data are presented as E/E<sub>0</sub> (where E—the average number of colonies in the tested extract concentration and E<sub>0</sub>—the average number in the control culture and standard deviation (SD); * statistically significant differences in cell viability (<span class="html-italic">p</span> &lt; 0.05) when comparing treated versus control groups, with error bars representing standard deviation.</p>
Full article ">Figure 3
<p>Cell colony formation properties after incubation with an aqueous extract of <span class="html-italic">F. betulina</span>: (<b>A</b>) Normal cell lines; (<b>B</b>) Cancer cell lines. Data are presented as E/E<sub>0</sub> (where E—the average number of colonies in the tested extract concentration and E<sub>0</sub>—the average number in the control culture and standard deviation (SD); * statistically significant differences in cell colony formation (<span class="html-italic">p</span> &lt; 0.05) when comparing treated versus control groups, with error bars representing standard deviation.</p>
Full article ">Figure 4
<p>Assessment of the genotoxic effect of an aqueous extract of <span class="html-italic">F. betulina</span> on cress seeds; * statistically significant differences in number of germinated seeds (<span class="html-italic">p</span> &lt; 0.05) when comparing treated versus control groups, with error bars representing standard deviation.</p>
Full article ">Figure 5
<p>Representative photos from a test of cress during growth and exposure to an aqueous extract of <span class="html-italic">F. betulina</span> on cress seeds.</p>
Full article ">Figure 6
<p>Average root length in Modified <span class="html-italic">Allium cepa</span> L. test; * statistically significant differences in average root length (<span class="html-italic">p</span> &lt; 0.05) when comparing treated versus control groups, with error bars representing standard deviation.</p>
Full article ">Figure 7
<p>Representative photos from a Modified <span class="html-italic">Allium cepa</span> L. test during growth and exposure to an aqueous extract of <span class="html-italic">F. betulina</span>.</p>
Full article ">Figure 8
<p><span class="html-italic">F. betulina</span> collected in the Lower Silesian forests and used to prepare the extract came from the same collection.</p>
Full article ">Figure 9
<p>Conducting Modified <span class="html-italic">Allium cepa</span> L. test.</p>
Full article ">Figure 10
<p>Properties of <span class="html-italic">F. betulina</span> aqueous extract in in vitro assays and with model plant organisms.</p>
Full article ">
Back to TopTop