[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (57,900)

Search Parameters:
Keywords = tumor

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 345 KiB  
Review
Synchronous Breast and Colorectal Malignant Tumors—A Systematic Review
by Cristian Iorga, Cristina Raluca Iorga, Alexandru Grigorescu, Iustinian Bengulescu, Traian Constantin and Victor Strambu
Life 2024, 14(8), 1008; https://doi.org/10.3390/life14081008 (registering DOI) - 13 Aug 2024
Abstract
The incidence of breast and colorectal cancers is well established in studies, but the synchronous occurrence of the two types of tumors is a rarity. In general, they are discovered during screening investigations following the diagnosis of an initial tumor. Objective: Our aim [...] Read more.
The incidence of breast and colorectal cancers is well established in studies, but the synchronous occurrence of the two types of tumors is a rarity. In general, they are discovered during screening investigations following the diagnosis of an initial tumor. Objective: Our aim is to describe the main diagnostic and therapeutic challenges for synchronous breast and colorectal tumors. Materials and methods: We performed a systematic review of the literature for cases or case series, using established keywords (synchronous breast and colon tumor and synonyms) for the period of 1970–2023. Five reviewers independently screened the literature, extracted data, and assessed the quality of the included studies. The results were processed according to the PRISMA 2020 guidelines. Results: A total of 15 cases were included in the study, including 2 males (age 50 and 57) and 13 females (median age 60, with range from 40 to 79). In a vast majority of the cases, the diagnosis of synchronous tumor was prompted by the first tumor’s workup. The first diagnosed tumor was colorectal in nine cases and a breast tumor in six cases. The most common histopathological type of breast tumor was invasive ductal carcinoma, and the colon tumors were exclusively adenocarcinomas. All cases had a surgical indication for both breast and colorectal tumor, except one case, in which the breast tumor had multiple metastasis. In four cases, the surgery was performed concomitantly (colectomy and mastectomy). In three cases, surgery was initially carried out for the breast tumor, followed by colon surgery. Oncological treatment was indicated depending on the tumor stage. Conclusions: For the treatment of synchronous tumors, the Tumor Board (T.B) decision is mandatory and must be personalized for each patient. Developing new methods of treatment and investigation may play an important role in the future for understanding synchronous tumor development, incidence, and outcome. Full article
11 pages, 2302 KiB  
Brief Report
Nanoparticle Uptake in the Aging and Oncogenic Drosophila Midgut Measured with Surface-Enhanced Raman Spectroscopy
by Maria Christou, Ayobami Fidelix, Yiorgos Apidianakis and Chrysafis Andreou
Cells 2024, 13(16), 1344; https://doi.org/10.3390/cells13161344 - 13 Aug 2024
Abstract
Colorectal cancer remains a major global health concern. Colonoscopy, the gold-standard colorectal cancer diagnostic, relies on the visual detection of lesions and necessitates invasive biopsies for confirmation. Alternative diagnostic methods, based on nanomedicine, can facilitate early detection of malignancies. Here, we examine the [...] Read more.
Colorectal cancer remains a major global health concern. Colonoscopy, the gold-standard colorectal cancer diagnostic, relies on the visual detection of lesions and necessitates invasive biopsies for confirmation. Alternative diagnostic methods, based on nanomedicine, can facilitate early detection of malignancies. Here, we examine the uptake of surface-enhanced Raman scattering nanoparticles (SERS NPs) as a marker for intestinal tumor detection and imaging using an established Drosophila melanogaster model for gut disease. Young and old Oregon-R and w1118 flies were orally administered SERS NPs and scanned without and upon gut lumen clearance to assess nanoparticle retention as a function of aging. Neither young nor old flies showed significant NP retention in their body after gut lumen clearance. Moreover, tumorigenic flies of the esg-Gal4/UAS-RasV12 genotype were tested for SERS NP retention 2, 4 and 6 days after RasV12 oncogene induction in their midgut progenitor cells. Tumorigenic flies showed a statistically significant NP retention signal at 2 days, well before midgut epithelium impairment. The signal was then visualized in scans of dissected guts revealing areas of NP uptake in the posterior midgut region of high stem cell activity. Full article
Show Figures

Figure 1

Figure 1
<p>SERS intensity at various nanoparticle relative concentrations. (<b>a</b>) The SERS nanoparticles demonstrate a characteristic spectrum that diminishes as they are serially diluted. The characteristic peaks appearing at 1203, 520 and 557 cm<sup>−1</sup> are highlighted. (<b>b</b>) A regression model was applied to the intensity of the 1203 cm<sup>−1</sup> peak, showing a logarithmic response of the nanoparticles as a function of concentration.</p>
Full article ">Figure 2
<p>Examples of (<b>a</b>) OR and (<b>b</b>) <span class="html-italic">w<sup>1118</sup></span> Smurf females with light diffused blue color (<b>left</b>) and non-Smurf (<b>right</b>).</p>
Full article ">Figure 3
<p>SERS NP uptake by young and old females. (<b>a</b>) Raman spectra from young and old flies, <span class="html-italic">w<sup>1118</sup></span> and OR, with and without clearance. The spectra demonstrate many Raman peaks intrinsic to the fly body. Spectra of individual flies are shown in grey and the group average in blue. The shaded bands indicate the areas of the peaks specific to the SERS NPs. (<b>b</b>) Regression (nn-LS) scores indicate moderate SERS signals in the flies fed with SERS NPs, with no statistically significant differences, as indicated by the <span class="html-italic">p</span>-values. Dots are individual fly measurements, the bars show the mean, and the whiskers the standard deviation. For each condition, 5 to 7 flies were scanned.</p>
Full article ">Figure 4
<p>SERS NP uptake by mutant flies. (<b>a</b>) Raman spectra for esg<sup>ts</sup>-Ras<sup>V12</sup> flies on day 2, 4 and 6 of induction of Ras<sup>V12</sup> oncogene in their midgut progenitors. Spectra of individual flies are shown in grey and the group average in blue. The shaded bands indicate the areas of the peaks specific to the SERS NPs. (<b>b</b>) Mean nn-LS scores and <span class="html-italic">p</span>-values as calculated for the different groups, show a small but statistically significant uptake for the 2-day condition, which decreases with disease progression. For each experimental condition, 10 flies were scanned.</p>
Full article ">Figure 5
<p>Raman maps of excised fly midguts. (<b>a</b>) Photograph of excised guts for NP-fed <span class="html-italic">esg<sup>ts</sup>-Ras<sup>V12</sup></span> flies. The superimposed colormap represents the nn-LS scores obtained from the different areas. Scale bar: 10 mm. (<b>b</b>) Detailed view from the sample areas indicated by boxes in (<b>a</b>). The optical microscopy images are shown with and without the colormap superimposed, for comparison. The areas with high nn-LS scores in their posterior midgut region display a reddish metallic texture, indicative of the presence of SERS NPs, particularly evident in (<b>bii</b>). Scale bar: 1 mm. (<b>c</b>) Representative Raman spectra corresponding to the pixels from panel (<b>b</b>) indicated with red crosses (×). Area (<b>bii</b>) features an exceptionally bright signal, whereas other areas display the characteristic peaks at lower intensities. (<b>d</b>–<b>f</b>) Similar to panels (<b>a</b>–<b>c</b>) but from control flies not fed with nanoparticles. Regression signals and Raman spectra have no indication of SERS nanoparticles, as expected.</p>
Full article ">
13 pages, 3243 KiB  
Article
Molecular Iodine Improves the Efficacy and Reduces the Side Effects of Metronomic Cyclophosphamide Treatment against Mammary Cancer Progression
by Evangelina Delgado-González, Ericka de los Ríos-Arellano, Brenda Anguiano and Carmen Aceves
Int. J. Mol. Sci. 2024, 25(16), 8822; https://doi.org/10.3390/ijms25168822 (registering DOI) - 13 Aug 2024
Abstract
Metronomic chemotherapy with cyclophosphamide (Cpp) has shown promising results in cancer protocols. These lower and prolonged doses have antiangiogenic, pro-cytotoxic, and moderate secondary effects. Molecular iodine (I2) reduces the viability of cancer cells and, with chemotherapeutic agents, activates the antitumoral immune [...] Read more.
Metronomic chemotherapy with cyclophosphamide (Cpp) has shown promising results in cancer protocols. These lower and prolonged doses have antiangiogenic, pro-cytotoxic, and moderate secondary effects. Molecular iodine (I2) reduces the viability of cancer cells and, with chemotherapeutic agents, activates the antitumoral immune response and diminishes side effects. The present work evaluates the adjuvant of oral I2 with Cpp using a murine model of mammary cancer. Female Sprague Dawley rats with 7,12-dimethylbenzantracene-induced tumors received Cpp intraperitoneal (50 and 70 mg/kg two times/week, iCpp50 and iCpp70) and oral (0.03%; 50 mg/Kg; oCpp50) doses. I2 (0.05%, 50 mg/100 mL) and oCpp50 were offered in drinking water for three weeks. iCpp70 was the most efficient antitumoral dose but generated severe body weight loss and hemorrhagic cystitis (HC). I2 prevented body weight loss, exhibited adjuvant actions with Cpp, decreasing tumor growth, and canceled HC mechanisms, including decreases in vascular endothelial growth factor (VEGF) and Survivin expression. oCpp50 + I2 diminished angiogenic signals (CD34, vessel-length, and VEGF content) and proinflammatory cytokines (interleukin-10 and tumor necrosis factor-alpha) and increased cytotoxic (lymphocytic infiltration, CD8+ cells, Tbet, and interferon-gamma) and antioxidant markers (nuclear erythroid factor-2 and glutathione peroxidase). I2 enhances the effectiveness of oCpp, making it a compelling candidate for a clinical protocol. Full article
(This article belongs to the Special Issue State-of-the-Art Molecular Oncology in Mexico, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Body weight and tumor growth in rats with mammary cancer. The graphs show the percentage of change in body weight gain in response to intraperitoneal (<b>A</b>) or oral (<b>B</b>) treatments. (<b>C</b>,<b>D</b>) show the percentage of tumor growth. Tumor volume was calculated by the ellipsoid formula: Volume = (major diameter x minor diameter<sup>2</sup>)/2. iCpp, intraperitoneal metronomic chemotherapy; oCpp, oral metronomic chemotherapy. Points represent means ± SD *, statistical differences (<span class="html-italic">p</span> &lt; 0.05) when compared to the control group. **, Statistical differences between iCpp70 and oCpp50 and iCpp70 + I<sub>2</sub> and oCpp50 + I<sub>2</sub>. <span class="html-italic">n</span> = 5/group.</p>
Full article ">Figure 2
<p>Molecular responses after 3 weeks of treatment. Expression (quantitative PCR) in tumoral samples from intraperitoneal (IP) or oral treatment. Apoptosis (<b>A</b>,<b>B</b>), chemoresistance (<b>C</b>,<b>D</b>), and angiogenesis (<b>E</b>,<b>F</b>). Immunodetection of peroxisome proliferator-activated receptor gamma (PPARγ) was analyzed by Western blotting. Expression was normalized by Ponceau staining and analyzed by band densitometry (<b>G</b>,<b>H</b>). Bax, bcl-2-like protein 4; Bcl-2, B-cell lymphoma 2; VEGF, vascular endothelial growth factor. Graphs representing mean ± SD. Data were analyzed using variance analysis (ANOVA) and Tukey post hoc test. Different letters indicate statistical differences (<span class="html-italic">p</span> &lt; 0.05). n = 3–5/group.</p>
Full article ">Figure 3
<p>Vascularity and angiogenesis in tumoral samples of oral treatment. (<b>A</b>): micrographs (20X) stained with anti-clusters of differentiation 34 (CD34; positive cells, black arrows). The insert shows (40X) the lymphocytes (hyperpigmented cells). (<b>B,C</b>): quantification of positive cells and vessel length (um), each value representing the average of three different fields: (<b>D</b>): multiplex protein content of VEGF. Graphs represent means ± SD. ANOVA and Tukey post hoc test were used to analyze data. Different letters indicate statistical differences (<span class="html-italic">p</span> &lt; 0.05). n = 3/group.</p>
Full article ">Figure 4
<p>Antitumoral immune response to oral metronomic chemotherapy (oCCp). (<b>A</b>): hematoxylin-eosin (H&amp;E) staining of tumors to identify lymphocyte infiltration (small round hyperpigmented cells). Micrographs (20X) and magnification (40X) are shown. (<b>B</b>): micrographs of CD8 positive cells (black arrows; 20X). (<b>C</b>): quantification of lymphocytic infiltration by average of three random fields. (<b>D</b>): number of positive cells to CD8, computing average of three random fields (20X). (<b>E</b>,<b>F</b>): immunodetection of Tbox transcription factor 21 (Tbet) and interferon-gamma (IFNγ) by Western blot. Expression was normalized by Ponceau staining and analyzed by band densitometry. Graphs represent means ± SD. ANOVA and Tukey post hoc tests analyzed data. Different letters indicate statistical differences (<span class="html-italic">p</span> &lt; 0.05). n = 3/group.</p>
Full article ">Figure 5
<p>Antioxidant and anti-inflammatory response to oCpp in rat mammary tumors. (<b>A</b>): nuclear erythroid factor 2 (Nrf2); (<b>B</b>): glutathione peroxidase (GPx) protein expression, as indicators of antioxidant mechanism, were evaluated by Western blotting. Expression was normalized by Ponceau staining and analyzed by band densitometry. (<b>C</b>,<b>D</b>), interleukin-10 (IL-10), and tumor necrosis factor (TNF-α) were assessed by multiplex assay. Graphs represent means ± SD. Data were analyzed using ANOVA and Tukey post hoc tests. Different letters indicate statistical differences (<span class="html-italic">p</span> &lt; 0.05). n = 3/group.</p>
Full article ">Figure 6
<p>I<sub>2</sub> supplementation prevents bladder inflammation associated with oCpp. (<b>A</b>): micrographs stained with H&amp;E of bladder epithelium (40X). Black arrows signalized the epithelial thickness, which was associated with inflammation: (<b>B</b>): quantitation of the thickness of the urothelium. The graph represents means ± SD. Data were analyzed using ANOVA and Tukey post hoc tests. Different letters indicate statistical differences (<span class="html-italic">p</span> &lt; 0.05). n = 3/group.</p>
Full article ">
13 pages, 507 KiB  
Review
High-Risk Neuroblastoma Challenges and Opportunities for Antibody-Based Cellular Immunotherapy
by Natasha V. Persaud, Jeong A. Park and Nai Kong V. Cheung
J. Clin. Med. 2024, 13(16), 4765; https://doi.org/10.3390/jcm13164765 - 13 Aug 2024
Abstract
Immunotherapy has emerged as an attractive option for patients with relapsed or refractory high-risk neuroblastoma (HRNB). Neuroblastoma (NB), a sympathetic nervous system cancer arising from an embryonic neural crest cell, is heterogeneous clinically, with outcomes ranging from an isolated abdominal mass that spontaneously [...] Read more.
Immunotherapy has emerged as an attractive option for patients with relapsed or refractory high-risk neuroblastoma (HRNB). Neuroblastoma (NB), a sympathetic nervous system cancer arising from an embryonic neural crest cell, is heterogeneous clinically, with outcomes ranging from an isolated abdominal mass that spontaneously regresses to a widely metastatic disease with cure rates of about 50% despite intensive multimodal treatment. Risk group stratification and stage-adapted therapy to achieve cure with minimal toxicities have accomplished major milestones. Targeted immunotherapeutic approaches including monoclonal antibodies, vaccines, adoptive cellular therapies, their combinations, and their integration into standard of care are attractive therapeutic options, although curative challenges and toxicity concerns remain. In this review, we provide an overview of immune approaches to NB and the tumor microenvironment (TME) within the clinical translational framework. We propose a novel T cell-based therapeutic approach that leverages the unique properties of tumor surface antigens such as ganglioside GD2, incorporating specific monoclonal antibodies and recent advancements in adoptive cell therapy. Full article
(This article belongs to the Special Issue High-Risk Neuroblastoma: New Clinical Insights and Challenges)
14 pages, 703 KiB  
Review
MiRNAs as Regulators of Immune Cells in the Tumor Microenvironment of Ovarian Cancer
by Miłosz Wilczyński, Jacek Wilczyński and Marek Nowak
Cells 2024, 13(16), 1343; https://doi.org/10.3390/cells13161343 - 13 Aug 2024
Abstract
Ovarian cancer is one of the leading causes of cancer deaths among women. There is an ongoing need to develop new biomarkers and targeted therapies to improve patient outcomes. One of the most critical research areas in ovarian cancer is identifying tumor microenvironment [...] Read more.
Ovarian cancer is one of the leading causes of cancer deaths among women. There is an ongoing need to develop new biomarkers and targeted therapies to improve patient outcomes. One of the most critical research areas in ovarian cancer is identifying tumor microenvironment (TME) functions. TME consists of tumor-infiltrating immune cells, matrix, endothelial cells, pericytes, fibroblasts, and other stromal cells. Tumor invasion and growth depend on the multifactorial crosstalk between tumor cells and immune cells belonging to the TME. MiRNAs, which belong to non-coding RNAs that post-transcriptionally control the expression of target genes, regulate immune responses within the TME, shaping the landscape of the intrinsic environment of tumor cells. Aberrant expression of miRNAs may lead to the pathological dysfunction of signaling pathways or cancer cell-regulatory factors. Cell-to-cell communication between infiltrating immune cells and the tumor may depend on exosomes containing multiple miRNAs. MiRNAs may exert both immunosuppressive and immunoreactive responses, which may cause cancer cell elimination or survival. In this review, we highlighted recent advances in the field of miRNAs shaping the landscape of immune cells in the TME. Full article
(This article belongs to the Special Issue Genetic Disorders in Breast and Ovarian Cancer)
Show Figures

Figure 1

Figure 1
<p>miRNAs regulate both TME and tumor, being part of a multilayered and mutual network between cancer and immune cells. Tumor-derived exosomes containing miRNAs modulate the actions of immune cells in the TME. Up- or downregulation of certain miRNAs in immune cells being part of the TME also affects tumor growth and progression. Such phenomena may cause immunotolerance or tumor elimination. Immune cells: natural killer cells—NK cells; myeloid-derived suppressor cells—MDSC; regulatory T cells—Treg; CD8<sup>+</sup> T cells—T cytotoxic; dendritic cells—DC; cancer-associated fibroblasts—CAFs.</p>
Full article ">
11 pages, 4222 KiB  
Article
Design of pH/Redox Co-Triggered Degradable Diselenide-Containing Polyprodrug via a Facile One-Pot Two-Step Approach for Tumor-Specific Chemotherapy
by Yanru Hu and Peng Liu
Molecules 2024, 29(16), 3837; https://doi.org/10.3390/molecules29163837 (registering DOI) - 13 Aug 2024
Abstract
The diselenide bond has attracted intense interest for drug delivery systems (DDSs) for tumor chemotherapy, owing to it possessing higher redox sensitivity than the disulfide one. Various redox-responsive diselenide-containing carriers have been developed for chemotherapeutics delivery. However, the premature drug leakage from these [...] Read more.
The diselenide bond has attracted intense interest for drug delivery systems (DDSs) for tumor chemotherapy, owing to it possessing higher redox sensitivity than the disulfide one. Various redox-responsive diselenide-containing carriers have been developed for chemotherapeutics delivery. However, the premature drug leakage from these DDSs was significant enough to cause toxic side effects on normal cells. Here, a pH/redox co-triggered degradable polyprodrug was designed as a drug self-delivery system (DSDS) by incorporating drug molecules as structural units in the polymer main chains, using a facile one-pot two-step approach. The proposed PDOX could only degrade and release drugs by breaking both the neighboring acid-labile acylhydrazone and the redox-cleavable diselenide conjugations in the drug’s structural units, triggered by the higher acidity and glutathione (GSH) or reactive oxygen species (ROS) levels in the tumor cells. Therefore, a slow solubility-controlled drug release was achieved for tumor-specific chemotherapy, indicating promising potential as a safe and efficient long-acting DSDS for future tumor treatment. Full article
(This article belongs to the Special Issue Exclusive Feature Papers on Molecular Structure)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR spectrum of selenolactone.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra of DOX, D-DOX<sub>ADH</sub> and PDOX.</p>
Full article ">Figure 3
<p>GPC curve of the proposed PDOX.</p>
Full article ">Figure 4
<p>Typical hydrodynamic diameter and distribution of the PDOX nanoparticles by dialysis at different concentrations (<b>a</b>) and the TEM image of the PDOX nanoparticles fabricated at 0.2 mg/mL (<b>b</b>).</p>
Full article ">Figure 5
<p>Drug release profiles of the PDOX nanoparticles in different media.</p>
Full article ">Figure 6
<p>CLSM images of the HepG2 cells after incubation with the PDOX nanoparticles (15 μg/mL) for 48 h: DOX (<b>a</b>), DAPI (<b>b</b>) and merged (<b>c</b>) (scale bar: 20 μm).</p>
Full article ">Figure 7
<p>Cell viability assay in L02 and HepG2 cells of PDOX nanoparticles (<b>a</b>) and DOX (<b>b</b>) with different concentrations for 48 h, respectively. Values are expressed as mean ± SD (<span class="html-italic">n</span> = 6) (* denotes significant difference <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Scheme 1
<p>Synthesis of the pH/redox co-triggered degradable polyprodrug (PDOX).</p>
Full article ">Scheme 2
<p>Synthesis of selenolactone.</p>
Full article ">Scheme 3
<p>pH/Reduction co-triggered degradation of PDOX to release the selenol (DOX-SeH).</p>
Full article ">Scheme 4
<p>pH/Oxidation co-triggered degradation of PDOX to release the seleninic acid (DOX-SeOOH).</p>
Full article ">
27 pages, 2187 KiB  
Review
Contribution of Keratinocytes in Skin Cancer Initiation and Progression
by Océane Dainese-Marque, Virginie Garcia, Nathalie Andrieu-Abadie and Joëlle Riond
Int. J. Mol. Sci. 2024, 25(16), 8813; https://doi.org/10.3390/ijms25168813 (registering DOI) - 13 Aug 2024
Abstract
Keratinocytes are major cellular components of the skin and are strongly involved in its homeostasis. Oncogenic events, starting mainly from excessive sun exposure, lead to the dysregulation of their proliferation and differentiation programs and promote the initiation and progression of non-melanoma skin cancers [...] Read more.
Keratinocytes are major cellular components of the skin and are strongly involved in its homeostasis. Oncogenic events, starting mainly from excessive sun exposure, lead to the dysregulation of their proliferation and differentiation programs and promote the initiation and progression of non-melanoma skin cancers (NMSCs). Primary melanomas, which originate from melanocytes, initiate and develop in close interaction with keratinocytes, whose role in melanoma initiation, progression, and immune escape is currently being explored. Recent studies highlighted, in particular, unexpected modes of communication between melanocytic cells and keratinocytes, which may be of interest as sources of new biomarkers in melanomagenesis or potential therapeutic targets. This review aims at reporting the various contributions of keratinocytes in skin basal cell carcinoma (BCC), cutaneous squamous cell carcinoma (cSCC), and melanoma, with a greater focus on the latter in order to highlight some recent breakthrough findings. The readers are referred to recent reviews when contextual information is needed. Full article
(This article belongs to the Special Issue Recent Advances in Skin Diseases)
Show Figures

Figure 1

Figure 1
<p>Skin architecture. (<b>A</b>) The skin is organized in three superposed layers: the epidermis, the dermis, and the hypo-dermis. (<b>B</b>) The epidermis cohesion is ensured by a large repertoire of adhesion molecules variously expressed among the epidermal layers (<span class="html-italic">Stratum</span>, S.) and forming different specialized adhesion structures, including adherens junctions (AJs), desmosomes (DSMs), tight junctions (TJs), and gap junctions (GJs). (<b>C</b>) Keratinocytes also communicate via Notch signaling.</p>
Full article ">Figure 2
<p>Interactions of keratinocytes with adjacent keratinocytes and melanocytes. Epidermal keratinocytes interact with adjacent keratinocytes (<b>left</b>) and melanocytes (<b>right</b>) through adhesion-mediated interactions and paracrine signaling. These interactions trigger pathways modulating keratinocyte proliferation, differentiation, wound healing and inflammation, and control melanocyte proliferation and functions.</p>
Full article ">Figure 3
<p>Main genetic deregulations in basal cell carcinoma and cutaneous squamous cell carcinoma. (<b>Left</b>) BCC is mostly driven by mutations leading to aberrant activation of the Hedgehog (Hh) pathway. Additional activating mutations in the EGFR and Hippo (Hpo) pathways contribute to the expression of growth and survival target genes. Inactivating TP53 mutation promotes resistance to senescence and cell death. (<b>Right</b>) cSCC presents mainly inactivating mutations in TP53 and CDKN2A promoting resistance to senescence and cell-death signals. Inactivating mutations in Notch1 and TGFBR1/2, activating mutations in HRAS, and increased expression of EGFR promote growth, survival, and migration. Inactivating or activating mutations are indicated with a red or green star, respectively. Increased expression is indicated with a green arrow.</p>
Full article ">Figure 4
<p>Recent advances in the comprehension of keratinocyte contributions to melanoma initiation and progression. Molecular events attributed to keratinocytes and functional consequences are reported in red.</p>
Full article ">
24 pages, 8704 KiB  
Article
Immunomodulatory R848-Loaded Anti-PD-L1-Conjugated Reduced Graphene Oxide Quantum Dots for Photothermal Immunotherapy of Glioblastoma
by Yu-Jen Lu, Reesha Kakkadavath Vayalakkara, Banendu Sunder Dash, Shang-Hsiu Hu, Thejas Pandaraparambil Premji, Chun-Yuan Wu, Yang-Jin Shen and Jyh-Ping Chen
Pharmaceutics 2024, 16(8), 1064; https://doi.org/10.3390/pharmaceutics16081064 (registering DOI) - 13 Aug 2024
Abstract
Glioblastoma multiforme (GBM) is the most severe form of brain cancer and presents unique challenges to developing novel treatments due to its immunosuppressive milieu where receptors like programmed death ligand 1 (PD-L1) are frequently elevated to prevent an effective anti-tumor immune response. To [...] Read more.
Glioblastoma multiforme (GBM) is the most severe form of brain cancer and presents unique challenges to developing novel treatments due to its immunosuppressive milieu where receptors like programmed death ligand 1 (PD-L1) are frequently elevated to prevent an effective anti-tumor immune response. To potentially shift the GBM environment from being immunosuppressive to immune-enhancing, we engineered a novel nanovehicle from reduced graphene oxide quantum dot (rGOQD), which are loaded with the immunomodulatory drug resiquimod (R848) and conjugated with an anti-PD-L1 antibody (aPD-L1). The immunomodulatory rGOQD/R8/aPDL1 nanoparticles can actively target the PD-L1 on the surface of ALTS1C1 murine glioblastoma cells and release R848 to enhance the T-cell-driven anti-tumor response. From in vitro experiments, the PD-L1-mediated intracellular uptake and the rGOQD-induced photothermal response after irradiation with near-infrared laser light led to the death of cancer cells and the release of damage-associated molecular patterns (DAMPs). The combinational effect of R848 and released DAMPs synergistically produces antigens to activate dendritic cells, which can prime T lymphocytes to infiltrate the tumor in vivo. As a result, T cells effectively target and attack the PD-L1-suppressed glioma cells and foster a robust photothermal therapy elicited anti-tumor immune response from a syngeneic mouse model of GBM with subcutaneously implanted ALTS1C1 cells. Full article
(This article belongs to the Special Issue Metal and Carbon Nanomaterials for Pharmaceutical Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustration of the preparation of rGOQD/R8/aPDL1, including reduction of GOQD to rGOQD using polyethylene (PEI), immunoadjuvant drug R848 loading on rGOQD by π–π stacking (rGOQD/R8), and aPD-L1 conjugation on rGO/R8 using amine groups in rGOQD and aldehyde group in activated aPD-L1. (<b>b</b>) The photo-immunotherapy using rGOQD/R8/aPDL1 involves photothermal therapy and immune cell activation to exert an anti-tumor effect by (1) binding to the overexpressed PD-L1 receptors on tumor cell surface; (2) R484 release for activation of adaptive immune response; (3) photothermal-effect-induced cell death; (4) antigen release and antigen-presenting cells (APCs) activation; (5) dendritic cells (DCs) activation; (6) T cells recruitment.</p>
Full article ">Figure 2
<p>(<b>a</b>) The TEM image of rGOQD/R8/aPDL1 (scale bar = 500 nm). (<b>b</b>) The size distribution from dynamic light scattering analysis of GOQD, rGOQD, rGOQD/R8, and rGOQD/R8/aPDL1. (<b>c</b>) The zeta potential values of different nanoparticles (mean ± SD, <span class="html-italic">n</span> = 3). (<b>d</b>) The FTIR spectra of GOQD, rGOQD, and rGOQD/R8. (<b>e</b>) The UV-Vis spectroscopy analysis of GOQD, rGOQD, rGOQD/R8, and rGOQD/R8/aPDL1.</p>
Full article ">Figure 3
<p>The photothermal images (<b>a</b>), and the corresponding temperature profiles (<b>b</b>) by irradiating GOQD or rGOQD/R8 (100 μg/mL) with 808 nm laser (1.5 W/cm<sup>2</sup>) for 5 min. The control is deionized water (DIW). The thermal images (<b>c</b>), and the corresponding temperature profiles (<b>d</b>) by irradiating 25–100 μg/mL rGOQD/R8/aPDL1 with 808 nm laser (1.5 W/cm<sup>2</sup>) for 5 min. (<b>e</b>) The in vitro release of R848 from rGOQD/R8/aPDL1 at pH 5 and 7.4. (<b>f</b>) The stability of rGOQD/R8/aPDL1 in PBS and DMEM cell culture medium by measuring the particle size from DLS. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>The biocompatibility of rGOQD/R8/aPDL1 was tested with 3T3 mouse embryonic fibroblast cells (<b>a</b>) and ALTS1C1 mouse glioma cells (<b>b</b>) with MTS assays at 24, 48, and 72 h. (<b>c</b>) The flow cytometry analysis of intracellular uptake of Cy5.5-tagged rGOQD/R8 and rGOQD/R8/aPDL1 after incubation with ALTS1C1 cells for 4 and 24 h. (<b>d</b>) The corresponding quantified fluorescence intensity from flow cytometry analysis of intracellular uptake of Cy5.5-tagged nanoparticles. * <span class="html-italic">p</span> &lt; 0.05 compared with rGOQD/R8. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>(<b>a</b>) The cell viability of ALTS1C1 cells after incubating cells with rGOQD/R8 (rGOQD/R8+L) or rGOQD/R8/aPDL1 (rGOQD/R8/aPDL1+L) at varying concentrations and irradiated with 808 nm laser at 1.5 W/cm<sup>2</sup> for 5 min. (<b>b</b>) The fluorescence microscopy images of Calcein-AM and PI co-stained cells after incubation with different nanoparticles and 808 nm NIR treatment at 1.5 W/cm<sup>2</sup> for 5 min. Cells treated with PBS and rGOQD/R8 without laser irradiation were used as controls. (<b>c</b>) The CLSM images of ALTS1C1 cells by staining with fluorescein-tagged anti-calreticulin antibody after different treatments. The rGOQD/R8+L and rGOQD/R8/aPDL1+L groups are irradiated with 808 nm NIR laser at 1.5 W/cm<sup>2</sup> for 5 min. (<b>d</b>) The corresponding quantification of calreticulin fluorescence intensity by using the PAX-it software. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3). <sup>α</sup> <span class="html-italic">p</span> &lt; 0.05 compared to PBS, <sup>β</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8, <sup>γ</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8+L.</p>
Full article ">Figure 6
<p>The ex vivo fluorescence images of major organs (<b>a</b>) and the corresponding quantification of nanoparticle distribution in each organ from fluorescence intensity (<b>b</b>) with an in vivo imaging system (IVIS) 4 h after administration of Cy5.5-labelled rGOQD/R8/aPDL1 or rGOQD/R8 to ALTS1C1 tumor-bearing mice through the tail vein. The ex vivo fluorescence images (<b>c</b>) and the corresponding fluorescence intensity (<b>d</b>) of tumors with an IVIS 4 h after administration of Cy5.5-labelled rGOQD/R8/aPDL1 or rGOQD/R8 to ALTS1C1 tumor-bearing mice through the tail vein.</p>
Full article ">Figure 7
<p>The thermal images (<b>a</b>) and the corresponding in vivo peak temperature profiles (<b>b</b>) of ALTS1C1 tumor-bearing mice after intravenous injection of rGOQD/R8/aPDL1 or rGOQD/R8 followed by 808 nm NIR laser irradiation 24 h post injection (mean ± SD, <span class="html-italic">n</span> = 3). The activation of dendritic cells by rGOQD/R8/aPDL1 or rGOQD/R8 was compared with confocal images of lymph nodes 19 days after treatment by immunofluorescence staining of CD11C (antigen-presenting cells) and CD86 (dendritic cells) in red and green, respectively, and counterstaining the nucleus with DAPI in blue (scale bar = 50 μm) (<b>c</b>).</p>
Full article ">Figure 8
<p>The in vivo therapeutic evaluation was studied using a syngeneic mouse model of GBM with subcutaneously implanted ALTSC1 cells. The mice were divided into four groups and the treatment was initiated by injection of samples on day 10, followed by intravenous injection on days 13, 17, and 20. The control group is PBS and the rGOQD/R8+L and rGOQD/R8/PDL1 groups are with 808 nm NIR laser irradiation at 1.5 W/cm<sup>2</sup> for 5 min. The tumor volume change (<b>a</b>), the scattered plot of tumor volume on day 21 (<b>b</b>), and the survival curve of animals (<b>c</b>) of ALTSC1 tumor-bearing mice after different treatments (mean ± SD, <span class="html-italic">n</span> = 3). The sacrificing criteria were when the tumor volume exceeded 1000 mm<sup>3</sup>. <sup>α</sup> <span class="html-italic">p</span> &lt; 0.05 compared to PBS, <sup>β</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8, <sup>γ</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8+L.</p>
Full article ">Figure 9
<p>(<b>a</b>) In vivo immune response after treatment with rGOQD/R8, rGOQD/R8+L, and rGOQD/R8/aPDL1+L. The infiltration of T cells into the tumor after various treatments was measured by immunofluorescence staining with anti-CD4 and anti-CD8 antibodies. The confocal immunofluorescence images and the corresponding quantification of fluorescence intensity of CD4 and CD8 in the tumor tissues 19 days after treatments (scale bar = 50 μm). (<b>b</b>) The immunohistochemistry (IHC) of PD-L1, tumor necrosis factor (TNF-α), and Ki-67, and H&amp;E staining of tumor tissues 19 days after treatments. <sup>α</sup> <span class="html-italic">p</span> &lt; 0.05 compared to PBS, <sup>β</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8, <sup>γ</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8+L.</p>
Full article ">
13 pages, 3012 KiB  
Article
Potentiation by Protein Synthesis Inducers of Translational Readthrough of Pathogenic Premature Termination Codons in PTEN Isoforms
by Leire Torices, Caroline E. Nunes-Xavier and Rafael Pulido
Cancers 2024, 16(16), 2836; https://doi.org/10.3390/cancers16162836 - 13 Aug 2024
Abstract
The PTEN tumor suppressor is frequently targeted in tumors and patients with PTEN hamartoma tumor syndrome (PHTS) through nonsense mutations generating premature termination codons (PTC) that may cause the translation of truncated non-functional PTEN proteins. We have previously described a global analysis of [...] Read more.
The PTEN tumor suppressor is frequently targeted in tumors and patients with PTEN hamartoma tumor syndrome (PHTS) through nonsense mutations generating premature termination codons (PTC) that may cause the translation of truncated non-functional PTEN proteins. We have previously described a global analysis of the readthrough reconstitution of the protein translation and function of the human canonical PTEN isoform by aminoglycosides. Here, we report the efficient functional readthrough reconstitution of the PTEN translational isoform PTEN-L, which displays a minimal number of PTC in its specific N-terminal extension in association with disease. We illustrate the importance of the specific PTC and its nucleotide proximal sequence for optimal readthrough and show that the more frequent human PTEN PTC variants and their mouse PTEN PTC equivalents display similar patterns of readthrough efficiency. The heterogeneous readthrough response of the different PTEN PTC variants was independent of the length of the PTEN protein being reconstituted, and we found a correlation between the amount of PTEN protein being synthesized and the PTEN readthrough efficiency. Furthermore, combination of aminoglycosides and protein synthesis inducers increased the readthrough response of specific PTEN PTC. Our results provide insights with which to improve the functional reconstitution of human-disease-related PTC pathogenic variants from PTEN isoforms by increasing protein synthesis coupled to translational readthrough. Full article
(This article belongs to the Special Issue PTEN: Regulation, Signalling and Targeting in Cancer)
Show Figures

Figure 1

Figure 1
<p>Distribution and translational readthrough of PTEN-L PTC associated with disease. (<b>A</b>) Kernel density plot of the distribution of the PTC on PTEN-L found in the germline of patients (germline PTCome). The potential PTCome includes all the PTC that can be generated via single-nucleotide substitutions, and it is shown as a control for sequence mutability. A schematic of PTEN-L domain composition is shown at the bottom (PTP, protein tyrosine phosphatase domain; C2, C2 domain). Numbers indicate PTEN-L amino acid numbering. (<b>B</b>) PTEN protein expression in the presence of PTEN-L PTC. COS-7 cells were transfected with empty vector (EV) or with plasmids encoding the indicated PTEN/PTEN-L [PTEN-L(Leu)] variants (WT, wild type), and protein expression was monitored via immunoblot using anti-PTEN 6H2.1 mAb, which recognizes the PTEN/PTEN-L common C-terminal region [<a href="#B36-cancers-16-02836" class="html-bibr">36</a>]. Immunoblot with anti-GAPDH antibody is shown as a loading control. The left panel shows a representative experiment, and arrows indicate the migration of PTEN-L and PTEN. The right panel shows the quantification of protein expression (relative to expression of PTEN ± SD) from at least two independent experiments. (<b>C</b>) Translational readthrough of PTC on PTEN-L. COS-7 cells were transfected with empty vector (EV) or with plasmids encoding the indicated PTEN or PTEN-L [PTEN-L(Met)] PTC variants (WT, wild type). Cells were kept untreated (−) or were treated (+) for 24 h with the readthrough inducer G418 (200 μg/mL), and readthrough was monitored via immunoblot with anti-PTEN 6H2.1 mAb. Immunoblot with anti-GAPDH antibody is shown as a loading control. The left panel shows a representative experiment, and arrows indicate the migration of PTEN-L and PTEN. The right panel shows quantification of readthrough, represented as PTC-readthrough (RT) efficiency ± SD (percentage of full-length PTEN or PTEN-L expression from PTC variants with respect to wild-type PTEN or PTEN-L) from at least two independent experiments. (<b>D</b>) Functional reconstitution of PTEN/PTEN-L through readthrough. COS-7 cells were co-transfected with pSG5 HA-AKT1 and empty vector (EV) or plasmids encoding the indicated PTEN/PTEN-L(Met) variants and processed for G418-induced readthrough, as in (<b>C</b>). The left panel shows a representative experiment. The right panel shows quantification of phospho-AKT (pAKT) content, represented as pAKT/AKT ± SD, relative to EV, from at least two independent experiments. The original Western blots are shown in <a href="#app1-cancers-16-02836" class="html-app">Supplementary Material File S1</a>.</p>
Full article ">Figure 2
<p>Nucleotide context and efficiency of PTC readthrough on PTEN-L and PTEN. (<b>A</b>,<b>B</b>) COS-7 cells were transfected with empty vector (EV) or with plasmids encoding PTEN (A) or PTEN-L(Met) (B) variants, as indicated (WT, wild type; +4C indicates a C in position +4, considering the first nucleotide of the PTC as +1), and processed for readthrough and immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. The left panels show representative experiments and arrows indicate the migration of PTEN/PTEN-L. The right panels show quantification of PTC-readthrough (RT) efficiency ± SD, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a> (+4 indicates the nucleotide at position +4). Statistical analysis was performed using an unpaired <span class="html-italic">t</span>-test (* = <span class="html-italic">p</span> &lt; 0.05). The original Western blots are shown in <a href="#app1-cancers-16-02836" class="html-app">Supplementary Material File S1</a>.</p>
Full article ">Figure 3
<p>PTC readthrough of mouse PTEN. (<b>A</b>) Alignment of human and mouse PTEN nucleotide sequences flanking amino acids frequently targeted by PTC in human disease. The residue targeted by PTC is indicated, and the PTC is underlined. The numbers indicate nucleotide numbering according to entries NM_000314 and NM_008960. Differences in nucleotides are marked in grey. (<b>B</b>) Translational readthrough of PTC from mouse PTEN. COS-7 cells were transfected with empty vector (EV) or with plasmids encoding human PTEN (hPTEN) or mouse PTEN (mPTEN) variants, as indicated (WT, wild type), and processed for readthrough and immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. The arrow indicates the migration of PTEN. (<b>C</b>) Comparative PTC-RT efficiency of PTC from human and mouse PTEN. The plot shows relative quantification of PTC-readthrough (RT) efficiency ± SD. The original Western blot is shown in <a href="#app1-cancers-16-02836" class="html-app">Supplementary Material File S1</a>.</p>
Full article ">Figure 4
<p>Readthrough efficiency of PTC variants from PTEN proteoforms synthesized at different levels. (<b>A</b>) COS-7 cells were transfected with plasmids encoding PTEN or PTEN-GFP variants, as indicated (WT, wild type), and processed for readthrough and immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. (<b>B</b>) Quantification of PTC-readthrough (RT) efficiency of PTEN and PTEN-GFP variants, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. (<b>C</b>) COS-7 cells were transfected with empty vector (EV) or with plasmids encoding PTEN, GST-PTEN, or PTEN-GFP variants, as indicated (WT, wild type), and processed for readthrough and immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. The arrows indicate the migration of PTEN, GST-PTEN, and PTEN-GFP. (<b>D</b>) Quantification of relative protein expression of PTEN, GST-PTEN, and PTEN-GFP. Relative protein expression ± SD is shown for wild-type (WT) proteins in the absence of G418 (−) relative to PTEN. (<b>E</b>) Quantification of PTC-readthrough (RT) efficiency of PTEN, GST-PTEN, and PTEN-GFP R130X variants, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. The original Western blots are shown in <a href="#app1-cancers-16-02836" class="html-app">Supplementary Material File S1</a>.</p>
Full article ">Figure 5
<p>Effect of protein synthesis inducer PMA on the PTC-readthrough efficiency of PTEN. (<b>A</b>) COS-7 or HEK293 cells were transfected with empty vector (EV) or with plasmids encoding HA-PTEN, HA-MMADHC, or HA-PTPN11, as indicated. Cells were kept untreated (−) or were treated (+) for 24 h with PMA (100 nM), and protein expression was monitored via immunoblot with anti-HA 12CA5 mAb. (<b>B</b>) COS-7 cells were transfected with the indicated PTEN variants (WT, wild type), and incubated 24 h in the absence (−) or the presence of PMA (100 nM) or G418 (200 μg/μL), as indicated. Cells were processed for immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. The upper panel shows a representative experiment, and the bottom panel shows quantification of PTC-readthrough (RT) efficiency ± SD, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. Statistical analysis was performed using an unpaired <span class="html-italic">t</span>-test (* = <span class="html-italic">p</span> &lt; 0.05). The original Western blots are shown in <a href="#app1-cancers-16-02836" class="html-app">Supplementary Material File S1</a>.</p>
Full article ">Figure 6
<p>Effect of protein synthesis inducer Y-320 on the PTC-readthrough efficiency of PTEN. (<b>A</b>) COS-7 cells were transfected with the indicated PTEN variants (WT, wild-type) and incubated for 24 h in the absence (−) or the presence of Y-320 (1 μM) or G418 (200 μg/μL), as indicated. Cells were processed for immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. (<b>B</b>) Quantification of PTC-readthrough (RT) efficiency of PTEN-L(Met) R130X in the absence (−) or in the presence of Y-320 (1 μM) or G418 (200 μg/μL), as indicated. PTC readthrough was monitored via immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. (<b>C</b>) COS-7 cells were transfected with the indicated PTEN variants and incubated for 24 h or 48 h in the absence (−) or the presence of Y-320 (1 μM), gentamicin (800 μg/mL), or G418 (200 μg/μL), as indicated. Cells were processed for immunoblot, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. (<b>D</b>,<b>E</b>) Quantification of PTC-readthrough (RT) efficiency of selected PTEN PTC variants upon G418 or G418+Y-320 readthrough induction, as in panel A. In (<b>D</b>), readthrough efficiency is shown individually for each variant, as in <a href="#cancers-16-02836-f001" class="html-fig">Figure 1</a>. In (<b>E</b>), readthrough efficiency is shown clustering the variants under the three different conditions. Statistical analysis was performed using a paired <span class="html-italic">t</span>-test (** = <span class="html-italic">p</span> &lt; 0.001). The original Western blots are shown in <a href="#app1-cancers-16-02836" class="html-app">Supplementary Material File S1</a>.</p>
Full article ">
13 pages, 5495 KiB  
Article
Experimental Study: The Development of a Novel Treatment for Chemotherapy-Resistant Tongue Cancer with the Inhibition of the Pathological Periostin Splicing Variant 1-2 with Exon 21
by Shoji Ikebe, Nobutaka Koibuchi, Kana Shibata, Fumihiro Sanada, Hideo Shimizu, Toshihiko Takenobu and Yoshiaki Taniyama
Cells 2024, 13(16), 1341; https://doi.org/10.3390/cells13161341 - 13 Aug 2024
Abstract
Tongue squamous cell carcinoma (TSCC) occurs frequently in the oral cavity, and because of its high proliferative and metastatic potential, it is necessary to develop a novel treatment for it. We have reported the importance of the inhibition of the periostin (POSTN) pathological [...] Read more.
Tongue squamous cell carcinoma (TSCC) occurs frequently in the oral cavity, and because of its high proliferative and metastatic potential, it is necessary to develop a novel treatment for it. We have reported the importance of the inhibition of the periostin (POSTN) pathological splicing variant, including exon 21 (PN1-2), in various malignancies, but its influence is unclear in tongue cancer. In this study, we investigated the potential of POSTN exon 21-specific neutralizing antibody (PN21-Ab) as a novel treatment for TSCC. Human PN2 was transfected into the human TSCC (HSC-3) and cultured under stress, and PN2 was found to increase cell viability. PN2 induced chemotherapy resistance in HSC-3 via the phosphorylation of the cell survival signal Akt. In tissues from human TSCC and primary tumors of an HSC-3 xenograft model, PN1-2 was expressed in the tumor stroma, mainly from fibroblasts. The intensity of PN1-2 mRNA expression was positively correlated with malignancy. In the HSC-3 xenograft model, CDDP and PN21-Ab promoted CDPP’s inhibition of tumor growth. These results suggest that POSTN exon 21 may be a biomarker for tongue cancer and that PN21-Ab may be a novel treatment for chemotherapy-resistant tongue cancer. The treatment points towards important innovations for TSCC, but many more studies are needed to extrapolate the results. Full article
(This article belongs to the Special Issue Advances in the Research of a Key Molecule in Periostin)
Show Figures

Figure 1

Figure 1
<p>The N-terminus of POSTN has four repeat domains (FAS1). The C-terminal region (exons 15–23) centered on PN1-4 undergoes alternative splicing. Pathological POSTN splicing variants include exons 17 and 21 (PN1-3).</p>
Full article ">Figure 2
<p>Hematoxylin–eosin (HE)-stained images and expression patterns of <span class="html-italic">PN1-4</span> and <span class="html-italic">PN1-2</span> mRNA using ISH (RNAscope<sup>TM</sup>, BaseScope<sup>TM</sup>) in each tissue. (<b>A</b>) Normal <span class="html-italic">human</span> tongue tissue. (<b>B</b>) <span class="html-italic">Human</span> TSCC (grade-1) tissue. (<b>C</b>) <span class="html-italic">Human</span> TSCC (grade-2) tissue. In (<b>B</b>,<b>C</b>), the lower images (RNAscope<sup>TM</sup>, BaseScope<sup>TM</sup>) are magnified views of the yellow box in the upper image (HE). The tumor parenchyma is surrounded by yellow dashed lines. <span class="html-italic">PN</span>1-4 and <span class="html-italic">PN</span>1-2 mRNA are stained red (as indicated by the yellow arrows).</p>
Full article ">Figure 2 Cont.
<p>Hematoxylin–eosin (HE)-stained images and expression patterns of <span class="html-italic">PN1-4</span> and <span class="html-italic">PN1-2</span> mRNA using ISH (RNAscope<sup>TM</sup>, BaseScope<sup>TM</sup>) in each tissue. (<b>A</b>) Normal <span class="html-italic">human</span> tongue tissue. (<b>B</b>) <span class="html-italic">Human</span> TSCC (grade-1) tissue. (<b>C</b>) <span class="html-italic">Human</span> TSCC (grade-2) tissue. In (<b>B</b>,<b>C</b>), the lower images (RNAscope<sup>TM</sup>, BaseScope<sup>TM</sup>) are magnified views of the yellow box in the upper image (HE). The tumor parenchyma is surrounded by yellow dashed lines. <span class="html-italic">PN</span>1-4 and <span class="html-italic">PN</span>1-2 mRNA are stained red (as indicated by the yellow arrows).</p>
Full article ">Figure 3
<p>Impact of PN2 on cell death in HSC-3. (<b>A</b>) Number of HSC-3 cells transfected with <span class="html-italic">PN2</span> in serum-free culture (<span class="html-italic">n</span> = 8, **; <span class="html-italic">p</span> &lt; 0.01 vs. control). (<b>B</b>) Number of <span class="html-italic">PN2</span>-transfected HSC-3 cells in serum-free culture with cisplatin treatment (<span class="html-italic">n</span> = 8, *; <span class="html-italic">p</span> &lt; 0.05 vs. control).</p>
Full article ">Figure 4
<p>Protein expression of phospho-Akt in <span class="html-italic">PN2</span> transgenic HSC-3 (serum-free culture). (<b>A</b>) Expression levels of phospho-Akt and total Akt, as measured by Western blotting. (<b>B</b>) Histograms indicating the relative expression levels of phospho-Akt/total Akt (*; <span class="html-italic">p</span> &lt; 0.05 vs. Venus).</p>
Full article ">Figure 5
<p>Protein expression of phospho-Akt in <span class="html-italic">PN2</span>-transfected HSC-3 (with CDDP administration). (<b>A</b>) Expression level of phospho-Akt and total Akt, as detected by Western blotting. (<b>B</b>) Histograms indicating the relative expression levels of phospho-Akt/total Akt. CDDP alone significantly decreased phospho-Akt/total Akt compared to Venus (*; <span class="html-italic">p</span> &lt; 0.05 vs. control), and CDDP with <span class="html-italic">PN2</span> transfection significantly increased phospho-Akt/total Akt compared to CDDP alone (*; <span class="html-italic">p</span> &lt; 0.05 vs. Venus).</p>
Full article ">Figure 6
<p>Expression patterns of <span class="html-italic">PN1-4</span> and <span class="html-italic">PN1-2</span> mRNA using ISH (RNAscope<sup>TM</sup>, BaseScope<sup>TM</sup>) in tissues of HSC-3 xenograft model primary tumors. The lower images are magnified views of the yellow box in the upper images. <span class="html-italic">PN1-4</span> and <span class="html-italic">PN1-2</span> are stained red. Cancer cells (cancer foci) are indicated by yellow arrows. The yellow scale bar indicates 50 µm.</p>
Full article ">Figure 7
<p>(<b>A</b>) Tumor size of primary tumors in an HSC-3 xenograft model treated with PN21-Ab; PN21-Ab significantly decreased tumor size at 14 days (<span class="html-italic">n</span> = 5, **; <span class="html-italic">p</span> &lt; 0.01 vs. control). (<b>B</b>) Tumor size of primary tumor in HSC-3 xenograft model treated with CDDP and PN21-Ab; CDDP significantly decreased tumor size compared to the control, and CDDP with PN21-Ab significantly decreased it compared to CDDP alone (<span class="html-italic">n</span> = 4, **; <span class="html-italic">p</span> &lt; 0.01 vs. control or CDDP with PN21-Ab). Tumor size (mm<sup>3</sup>) was calculated as 1/2 × width (mm) × length (mm).</p>
Full article ">Figure 8
<p>Possible mechanism of PN1-2 expression. PN1-2 is expressed primarily in stromal fibroblasts. Of course, the anticancer effect of CDDP decreases phospho-Akt in tongue cancer cells. However, it is thought that PN1-2 expressed from host stromal fibroblasts actually increases phospho-Akt.</p>
Full article ">
8 pages, 6608 KiB  
Case Report
Superficial Anaplastic Lymphoma Kinase-Rearranged Myxoid Spindle Cell Neoplasm in the Buttock: A Case Report
by Jong-Hyup Kim, In-Chang Koh, Hoon Kim, Soo-Yeon Lim, Joon-Hyuk Choi and Kun-Young Kwon
J. Pers. Med. 2024, 14(8), 858; https://doi.org/10.3390/jpm14080858 (registering DOI) - 13 Aug 2024
Abstract
Anaplastic lymphoma kinase (ALK) is detected in both normal and oncological developmental tissues. Among ALK-related tumors, superficial ALK-rearranged myxoid spindle cell neoplasm (SAMS) is a rare, soft tissue tumor characterized by the immunophenotypical co-expression of CD34 and S100. Here, we describe a patient [...] Read more.
Anaplastic lymphoma kinase (ALK) is detected in both normal and oncological developmental tissues. Among ALK-related tumors, superficial ALK-rearranged myxoid spindle cell neoplasm (SAMS) is a rare, soft tissue tumor characterized by the immunophenotypical co-expression of CD34 and S100. Here, we describe a patient with this rare tumor and outline its clinical and radiological characteristics. A 28-year-old woman with diabetes, hypertension, and panic disorder presented with discomfort caused by a rubbery mass on the left buttock that had persisted for 10 years. Computed tomography revealed a multilobulated hypodense mass with small internal enhancing foci, posing challenges for the exact diagnosis of the lesion. The entire lesion was excised with clear resection margins. An 8.0 × 6.0 cm, well-circumscribed tumor with a lobular growth pattern was observed in the deep subcutaneous tissue. Light microscopy revealed epithelioid, ovoid, and spindle-shaped cells with a reticular cordlike pattern. Immunohistochemistry results were positive for S100, CD34, and vimentin. Break-apart fluorescence in situ hybridization assay results for ALK were also positive. These findings were consistent with those of SAMS. This case suggests that SAMS should be considered when identifying large nonspecific masses during clinical and imaging evaluation. Full article
Show Figures

Figure 1

Figure 1
<p>Abdominal and pelvic computed tomography (CT; (<b>left</b>)) and ultrasound (<b>right</b>) scans obtained 3 years before surgery. CT shows a well-defined, lobulated, mildly enhancing soft-tissue mass (yellow arrow) with skin thickening extending from the skin to the subcutaneous layer of the left buttock. On ultrasonography, the mass (yellow pluses indicating the margins) appears as a subcutaneous hypoechoic lesion with post-acoustic enhancement and is thus considered an epidermoid cyst.</p>
Full article ">Figure 2
<p>Axial (<b>left</b>) and sagittal (<b>right</b>) abdominal and pelvic CT scans obtained immediately before surgery. Both scans show a large, multilobulated, hypodense mass (yellow arrows) with internal enhancing foci in the left buttock.</p>
Full article ">Figure 3
<p>Gross findings. A well-circumscribed, yellowish-white solid tumor with a lobular growth pattern in the subcutaneous area. (<b>a</b>) Surgically excised tumor; (<b>b</b>) Cut section of the tumor.</p>
Full article ">Figure 4
<p>Histopathological findings. (<b>a</b>–<b>c</b>) Tumor cells showing reticular, cord (<b>a</b>), and focally hypercellular ((<b>b</b>), arrows) patterns in the prominent hyalinized stroma, as well as perivascular hyalinization (<b>c</b>). (<b>d</b>) Tumor cells arranged in striking whorls within the myxoid matrix. Hematoxylin and eosin staining; original magnification: (<b>a</b>,<b>b</b>), 100×; (<b>c</b>), 400×; (<b>d</b>), 200×.</p>
Full article ">Figure 5
<p>Immunohistochemical analysis of tumor cells. (<b>a</b>–<b>d</b>) Diffuse positivity for CD34 (<b>a</b>), S100 (<b>b</b>), and anaplastic lymphoma kinase (ALK; (<b>c</b>)), and negativity for CD31 (<b>d</b>). Original magnification (<b>a</b>–<b>d</b>): 100×. (<b>e</b>,<b>f</b>) Break-apart fluorescence in situ hybridization reveals ALK rearrangements.</p>
Full article ">
11 pages, 1462 KiB  
Article
MicroRNA: A Signature for the Clinical Progression of Chronic Lymphocytic Leukemia
by Yuliya A. Veryaskina, Sergei E. Titov, Igor B. Kovynev, Tatiana I. Pospelova, Sofya S. Fyodorova, Yana Yu. Shebunyaeva, Sergei A. Demakov, Pavel S. Demenkov and Igor F. Zhimulev
Lymphatics 2024, 2(3), 157-167; https://doi.org/10.3390/lymphatics2030013 (registering DOI) - 13 Aug 2024
Abstract
Chronic lymphocytic leukemia (CLL) is the most common human leukemia. The disease is caused by abnormal proliferation and development of lymphocytes and their precursors in the blood and bone marrow (BM). Recent studies have shown that the CLL’s clinical course and outcome depend [...] Read more.
Chronic lymphocytic leukemia (CLL) is the most common human leukemia. The disease is caused by abnormal proliferation and development of lymphocytes and their precursors in the blood and bone marrow (BM). Recent studies have shown that the CLL’s clinical course and outcome depend not only on genetic but also epigenetic factors. MicroRNAs (miRNAs) are involved in the development of hematological tumors, including CLL. The aim of this study is to identify the miRNA expression profile in CLL and determine the role of miRNAs in biological pathways associated with leukemogenesis in CLL. The following samples were used in this study: (1) samples obtained by sternal puncture and aspiration biopsy of BM (n = 115). They included samples from 21 CLL patients with anemia and indications for therapy and 45 CLL patients without anemia and with indications for therapy. The control group for the CLL BM samples consisted of patients with non-cancerous blood diseases (n = 35). (2) Lymph node (LN) samples (n = 20) were collected from CLL patients. The control group for the CLL LN samples consisted of patients with lymphadenopathy (n = 37). All cases were patients before treatment. We demonstrated a significant upregulation of miRNA-34a and miRNA-150 in CLL BM samples (p < 0.05) and downregulation of miRNA-451a in CLL LN samples (p < 0.05). We noted a dynamic increase in the levels of miRNA-150 and miRNA-34a in BM at various stages of tumor progression of CLL. We concluded that a dynamic picture of clinical manifestations of CLL closely correlates with changes in epigenetic characteristics of the tumor. Progression of the lymphoproliferative process and indications for cytoreductive therapy are associated with changes in the miRNA profile generated by cancer cells in different sites of clonal expansion. Full article
Show Figures

Figure 1

Figure 1
<p>A comparative analysis of miRNA expression levels in chronic lymphocytic leukemia without indications for therapy (CLL (T−)), with indications for therapy and without anemia (CLL (T+) (A−)), with indications for therapy and anemia (CLL (T+) (A+)), and non-cancerous blood diseases (NCBDs): (<b>A</b>) miRNA-34a; (<b>B</b>) miRNA-150. The median value, upper and lower quartiles, non-outlier range, and outliers indicated by circles are presented in the figure. Horizontal lines connecting boxes in pairs indicate statistically significant differences between the compared subgroups. Asterisks denote statistically significant differences at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Target analysis of miRNA-34a and miRNA-150 using miRnet 2.0. Green diamonds indicate microRNAs, red diamonds indicate their target genes.</p>
Full article ">Figure 3
<p>A schematic representation of the experimental design and workflow of the microRNA analysis. Total RNA was isolated from fine-needle aspiration cytological specimens of bone marrow and paraffin-embedded sections of lymph node tissue. Subsequently, miRNA expression levels were evaluated using real-time PCR. Then, target genes involved in hematopoiesis, lymphocyte differentiation, and cancer pathways were analyzed using bioinformatics methods.</p>
Full article ">
12 pages, 1415 KiB  
Article
Could Flow Cytometry Provide New Prognostic Markers in Colorectal Cancer?
by Vaia Georvasili, Georgios Markopoulos, Evangeli Lampri, Georgios Lianos, George Vartholomatos, Michail Mitsis and Christina Bali
J. Clin. Med. 2024, 13(16), 4753; https://doi.org/10.3390/jcm13164753 - 13 Aug 2024
Abstract
Background/Objectives: Colorectal cancer (CRC) is still accompanied by significant mortality, which poses the necessity of novel markers to predict treatment success and patient survival. This study aims to evaluate the prognostic and survival impact of flowytometry (FC) in CRC patients. Methods: [...] Read more.
Background/Objectives: Colorectal cancer (CRC) is still accompanied by significant mortality, which poses the necessity of novel markers to predict treatment success and patient survival. This study aims to evaluate the prognostic and survival impact of flowytometry (FC) in CRC patients. Methods: In this prospective study, 106 surgically resectable CRC patients were included. Tissue specimens from tumor and normal mucosa were collected and analyzed by FC. DNA and tumor index were calculated. In a subgroup of 46 patients, the CD26 expression on tumor cells was estimated. These parameters were compared with patients’ tumor characteristics as stage, histology data, responsiveness to treatment, metastasis/recurrence, and, finally, patients’ survival to identify possible new biomarkers. Results: The overall survival and the disease-specific survival in our study group was 76% and 72%, respectively, during the 7-year follow up period. Diploid tumors had better median survival than the aneuploid ones. The DNA index had significant correlation to the tumor index and response to neoadjuvant treatment. Similarly, the tumor index was also significantly related to the response to neoadjuvant treatment. Patients with a higher tumor index had worst survival rates. Surprisingly, CD26 levels were not associated with any of the parameters examined and were negatively related to tumor stage and differentiation. Conclusions: FC is a rapid and reliable method of cell analysis. In CRC, it has been used for prognostic and diagnostic purposes. In this study, we have shown that DNA and tumor index could become predictive biomarkers of tumor response to neoadjuvant treatment and survival of resectable CRC patients. Full article
(This article belongs to the Special Issue Recent Advancements and Challenges in Colorectal Surgery)
Show Figures

Figure 1

Figure 1
<p>Panel (<b>A</b>): graph depicting the relationship between DNA index and tumor regression grade (TRG). The plot illustrates the association between the cellular DNA content and the response of tumors to treatment, as determined using the TRG system. Panel (<b>B</b>): graph displaying the correlation between DNA index and mucous element presence within all the tumors in the study group. This plot reveals the association between DNA content variability and the extent of mucous production by tumor cells, highlighting potential prognostic implications. Points labeled with sample numbers indicate cases with values more than two standard deviations from the mean.</p>
Full article ">Figure 2
<p>Median overall and disease-free survival in association with the tumor’s DNA index. Groups &lt;0.95, 0.95–1.05, and &gt;1.05 correspond to hypoploid, diploid (with an acceptable standard deviation of 5%), and hyperploid, as measured by intraoperative flow cytometry.</p>
Full article ">Figure 3
<p>Panel (<b>A</b>) Kaplan–Meier survival curve illustrating overall survival in patients with colorectal tumors, categorized by low and high tumor index. This graph compares the overall survival probabilities over time between the two groups, highlighting differences in patient outcomes based on tumor index categorization. Panel (<b>B</b>): Kaplan–Meier survival curve depicting disease-free survival (DFS) in patients with low- and high-tumor-index colorectal tumors. This curve demonstrates the duration of survival without signs of disease recurrence or progression, contrasting the prognostic impacts of varying tumor index levels.</p>
Full article ">Figure 4
<p>Panel (<b>A</b>): box plot presenting the distribution of tumor index (TI) values among patients undergoing neoadjuvant therapy for colorectal cancer. This graph highlights the statistical correlation between TI levels and the administration of neoadjuvant treatment (<span class="html-italic">p</span> &lt; 0.01), showcasing how TI can vary with treatment application. Panel (<b>B</b>): box plot showing the relationship between tumor index (TI) and tumor regression grade (TRG) following neoadjuvant treatment. Lower TI values are associated with better TRG outcomes, indicating a more effective response to treatment. The plot confirms the statistical significance of this relationship (<span class="html-italic">p</span> &lt; 0.01). Points labeled with sample numbers indicate cases with values more than two standard deviations from the mean.</p>
Full article ">Figure 5
<p>In these histograms, the levels of CD26 expression by tumor cells are shown, categorized into four groups: absent (0), low (1), moderate (2), and high (3). Panel (<b>A</b>) displays the distribution of CD26 expression across different CRC stages (1 = stage I, 2 = stage II, 3 = stage III, 4 = stage IV), while panel (<b>B</b>) illustrates the association between CD26 expression levels and tumor differentiation grades (numbers in axis represent 1 = well differentiated, 2 = moderate differentiation, 3 = poor differentiation). Analysis was performed using intraoperative flow cytometry (iFC). Points labeled with sample numbers indicate cases with values more than two standard deviations from the mean.</p>
Full article ">
19 pages, 4899 KiB  
Article
Antitumoral and Antimetastatic Activity by Mixed Chelate Copper(II) Compounds (Casiopeínas®) on Triple-Negative Breast Cancer, In Vitro and In Vivo Models
by Mauricio M. González-Ballesteros, Luis Sánchez-Sánchez, Adrián Espinoza-Guillén, Jesús Espinal-Enríquez, Carmen Mejía, Enrique Hernández-Lemus and Lena Ruiz-Azuara
Int. J. Mol. Sci. 2024, 25(16), 8803; https://doi.org/10.3390/ijms25168803 (registering DOI) - 13 Aug 2024
Abstract
Triple-negative breast cancer (TNBC), accounting for 15–20% of all breast cancers, has one of the poorest prognoses and survival rates. Metastasis, a critical process in cancer progression, causes most cancer-related deaths, underscoring the need for alternative therapeutic approaches. This study explores the anti-migratory, [...] Read more.
Triple-negative breast cancer (TNBC), accounting for 15–20% of all breast cancers, has one of the poorest prognoses and survival rates. Metastasis, a critical process in cancer progression, causes most cancer-related deaths, underscoring the need for alternative therapeutic approaches. This study explores the anti-migratory, anti-invasive, anti-tumoral, and antimetastatic effects of copper coordination compounds Casiopeína IIIia (CasIIIia) and Casiopeína IIgly (CasIIgly) on MDA-MB-231 and 4T1 breast carcinoma cell lines in vitro and in vivo. These emerging anticancer agents, mixed chelate copper(II) compounds, induce apoptosis by generating reactive oxygen species (ROS) and causing DNA damage. Whole-transcriptome analysis via gene expression arrays indicated that subtoxic concentrations of CasIIIia upregulate genes involved in metal response mechanisms. Casiopeínas® reduced TNBC cell viability dose-dependently and more efficiently than Cisplatin. At subtoxic concentrations (IC20), they inhibited random and chemotactic migration of MDA-MB-231 and 4T1 cells by 50–60%, similar to Cisplatin, as confirmed by transcriptome analysis. In vivo, CasIIIia and Cisplatin significantly reduced tumor growth, volume, and weight in a syngeneic breast cancer model with 4T1 cells. Furthermore, both compounds significantly decreased metastatic foci in treated mice compared to controls. Thus, CasIIIia and CasIIgly are promising chemotherapeutic candidates against TNBC. Full article
(This article belongs to the Special Issue State-of-the-Art Molecular Oncology in Mexico, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of Cisplatin, Casiopeína IIgly (CasIIgly), and Casiopeína IIIia (CasIIIia).</p>
Full article ">Figure 2
<p>Metallodrugs inhibit migration of breast cancer cells: (<b>A</b>) Migration capacity of MDA-MB-231 and 4T1 cells was evaluated by wound-healing assay after treatment with metallodrugs (CasIIIia, CasIIgly, and Cisplatin) for 24 h. Representative photographs of wound-healing assay in the corresponding treatments from the initial time 0 h to the end of treatment 24 h, viewed at 40× are shown. (<b>B</b>) Analysis of wound area and percentage of migration with the respective treatments, in the MDA-MB-231 and 4T1 cells. Bars represent the means ± S.D. of at least three independent experiments. A one-factor ANOVA was performed using Tukey’s test for multiple comparisons. An asterisk (*) indicates statistically significant differences with respect to the untreated group and the vehicle group, with a significance of * <span class="html-italic">p</span> &lt; 0.01. There were no significant differences between the CasIIIia, CasIIgly, and Cisplatin groups.</p>
Full article ">Figure 3
<p>Metallodrugs inhibit chemo-migration of breast cancer cells. Chemo-migration capacity of MDA-MB-231 and 4T1 cells were assessed by transwell assay after exposure to metallodrugs (CasIIIia, CasIIgly, and Cisplatin) for 24 h: (<b>A</b>) Representative photographs of the chemo-migration assay in transwell chambers taken at 24 h with the different treatments (IC<sub>20</sub>). (<b>B</b>) The graphs represent the percentage of cells migrating at 24 h considering the vehicle group (H<sub>2</sub>O) as 100%, CasIIIia, CasIIgly, and Cisplatin. Each bar represents the mean ± S.D., n = 3. A one-factor ANOVA with Tukey’s test for multiple comparisons was performed, and the results were statistically significant. An asterisk (*) indicates statistically significant differences with respect to the untreated group and the vehicle group, with a significance of * <span class="html-italic">p</span> &lt; 0.01. There were no significant differences between the CasIIIia, CasIIgly, and Cisplatin groups.</p>
Full article ">Figure 4
<p>CasIIIia and Cisplatin inhibit tumor growth rate in a breast cancer mouse model (4T1 cells): (<b>A</b>) Tumor growth curve during the period of treatment. A one-factor ANOVA was performed with a Tukey’s test for multiple comparisons. The ANOVA was performed comparing all groups at day 30 (end of treatment). The tumor volume of the CasIIIia- and Cisplatin-treated groups was found to be statistically significant with respect to the untreated group and the glucose solution 5% group. (<b>B</b>) Tumor weight of each treated group at the end of the experiment. (<b>C</b>) The tumor growth velocity was calculated and the treatments with metallodrugs showed a significant effect. Each bar represents the mean ± S.E, n = 10. A one-factor ANOVA with Tukey’s test for multiple comparisons was performed and the results were statistically significant. An asterisk (*) indicates statistically significant differences with respect to the untreated group and the Glucose solution 5% group, with a significance of * <span class="html-italic">p</span> &lt; 0.01. There were no significant differences between the CasIIIia, and Cisplatin groups.</p>
Full article ">Figure 5
<p>Monitoring and % Weight loss in Balb/c mice treated with Metallodrugs: (<b>A</b>) The initial and final weight of BALB/c mice during and after treatment. Mean body weight changes over 21 days after treatment. Each value represents a mean ± SE of n = 10. (<b>B</b>) Toxicity was measured by the percentage weight loss of the mice. The percentage weight loss, as an indicator of toxicity, was calculated for each animal as follows: [(weight on day 21/weight on day 0) − 1] × 100. Unpaired <span class="html-italic">t</span>-test was used to evaluate the statistical significance between groups.</p>
Full article ">Figure 6
<p>The in vivo antimetastatic effect of Metallodrugs: (<b>A</b>) Representative photographs of lungs of the different treatments. (<b>B</b>) Mean number of metastatic <span class="html-italic">foci</span> (Macrometastasis). Each point represents the mean ± S.E, n = 10. A Student’s <span class="html-italic">t</span>-test found that CasIIIia and Cisplatin were statistically different <span class="html-italic">t</span> with respect to the untreated group (the Student’s <span class="html-italic">t</span>-test was performed with standard error). An asterisk (*) indicates statistically significant differences with respect to the untreated group and the Glucose Solution 5% group, with a significance of * <span class="html-italic">p</span> &lt; 0.01. There were no significant differences between the CasIIIia, and Cisplatin groups.</p>
Full article ">Figure 7
<p>Heatmap of the enriched processes based on the differential expression analysis associated with each contrast (rows). Intensity color is proportional to the −log(<span class="html-italic">p</span>-val) of each process (for clarity, the list of enriched processes is provided in the <a href="#app1-ijms-25-08803" class="html-app">Supplementary Material Table S1</a>).</p>
Full article ">Figure 8
<p>Network of enriched processes associated with the differential expression contrasts. Green diamonds represent underexpressed gene sets for the specific contrast. Blue diamonds take account of overexpressed gene sets. Circles show the biological processes linked to each contrast. Color of circles represents the global category of those processes. In <a href="#app1-ijms-25-08803" class="html-app">Supplementary Material</a>, we have provided the same figure with all the names of enriched biological processes.</p>
Full article ">
7 pages, 1635 KiB  
Article
Machine Learning on Ultrasound Texture Analysis Data for Characterizing of Salivary Glandular Tumors: A Feasibility Study
by Li-Jen Liao, Ping-Chia Cheng and Feng-Tsan Chan
Diagnostics 2024, 14(16), 1761; https://doi.org/10.3390/diagnostics14161761 - 13 Aug 2024
Abstract
Background: Objective quantitative texture characteristics may be helpful in salivary glandular tumor differential diagnosis. This study uses machine learning (ML) to explore and validate the performance of ultrasound (US) texture features in diagnosing salivary glandular tumors. Material and methods: 122 patients with salivary [...] Read more.
Background: Objective quantitative texture characteristics may be helpful in salivary glandular tumor differential diagnosis. This study uses machine learning (ML) to explore and validate the performance of ultrasound (US) texture features in diagnosing salivary glandular tumors. Material and methods: 122 patients with salivary glandular tumors, including 71 benign and 51 malignant tumors, are enrolled. Representative brightness mode US pictures are selected for further Gray Level Co-occurrence Matrix (GLCM) texture analysis. We use a t-test to test the significance and use the receiver operating characteristic curve method to find the optimal cut-point for these significant features. After splitting 80% of the data into a training set and 20% data into a testing set, we use five machine learning models, k-nearest Neighbors (kNN), Naïve Bayes, Logistic regression, Artificial Neural Networks (ANNs) and supportive vector machine (SVM), to explore and validate the performance of US GLCM texture features in diagnosing salivary glandular tumors. Results: This study includes 49 female and 73 male patients, with a mean age of 53 years old, ranging from 21 to 93. We find that six GLCM texture features (contrast, inverse difference movement, entropy, dissimilarity, inverse difference and difference entropy) are significantly different between benign and malignant tumors (p < 0.05). In ML, the overall accuracy rates are 74.3% (95%CI: 59.8–88.8%), 94.3% (86.6–100%), 72% (54–89%), 84% (69.5–97.3%) and 73.5% (58.7–88.4%) for kNN, Naïve Bayes, Logistic regression, a one-node ANN and SVM, respectively. Conclusions: US texture analysis with ML has potential as an objective and valuable tool to make a differential diagnosis between benign and malignant salivary gland tumors. Full article
(This article belongs to the Special Issue Current Perspectives and Advances in Ultrasound Imaging)
Show Figures

Figure 1

Figure 1
<p>Overview of the workflow for this study, a maximal rectangle area within the salivary glandular tumor is delineated for the ROI (region of interest).</p>
Full article ">Figure 2
<p>(<b>A</b>) Case 2, the square block is sampled from a right parotid tumor for GLCM texture analysis by Image J. The pathologic report reveals pleomorphic adenoma. (<b>B</b>) The square block is sampled for GLCM texture analysis from another left parotid tumor, and the pathologic report reveals mucoepidermoid carcinoma.</p>
Full article ">Figure 3
<p>With one hidden node with 6 predictors, the accuracy rate of this ANN model is 84.0% (95% CI: 69.5–97.3%).</p>
Full article ">
Back to TopTop