[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 13, September-1
Previous Issue
Volume 13, August-1
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
cancers-logo

Journal Browser

Journal Browser

Cancers, Volume 13, Issue 16 (August-2 2021) – 327 articles

Cover Story (view full-size image): Aptamers, which are short sequences of oligonucleotide or peptides, have become an increasingly popular tool within the field of oncology. This increase in usage is attributed to an aptamer-specific design which allows for a high degree of target specificity and high binding affinity, thus making aptamers of specific interest for the design of sensing platforms. We present the reader with all the prominent known aptamers and their receptors for the detection of breast cancer. Despite breast cancer being the most commonly occurring cancer in women worldwide, data on these aptamers are lacking and slow progress in a promising area of early detection. With the knowledge presented in this review, researchers will have a hub to explore the advantages and disadvantages of each aptamer and their target, thus allowing for informed project design and increased research in a high-priority disease. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
13 pages, 948 KiB  
Article
Surgical Techniques and Outcomes of Colorectal Anastomosis after Left Hemicolectomy with Low Anterior Rectal Resection for Advanced Ovarian Cancer
by Kyoko Nishikimi, Shinichi Tate, Ayumu Matsuoka, Satoyo Otsuka and Makio Shozu
Cancers 2021, 13(16), 4248; https://doi.org/10.3390/cancers13164248 - 23 Aug 2021
Cited by 2 | Viewed by 3719
Abstract
Extended colon resection is often performed in advanced ovarian cancer. Restoring intestinal continuity and avoiding stoma creation improve patients’ quality of life postoperatively. We tried to minimize the number of anastomoses, restore intestinal continuity, and avoid stoma creation for 295 patients with stage [...] Read more.
Extended colon resection is often performed in advanced ovarian cancer. Restoring intestinal continuity and avoiding stoma creation improve patients’ quality of life postoperatively. We tried to minimize the number of anastomoses, restore intestinal continuity, and avoid stoma creation for 295 patients with stage III/IV ovarian cancer who underwent low anterior rectal resection (LAR) with or without colon resection during cytoreductive surgery. When the remaining colon could not reach the rectal stump after left hemicolectomy with LAR, we used the following techniques for tension-free anastomosis: right colonic transposition, retro-ileal anastomosis through an ileal mesenteric defect, or an additional colic artery division. Rates of stoma creation and rectal anastomotic were 3% (9/295) and 6.6% (19/286), respectively. Among 21 patients in whom the remaining colon did not reach the rectal stump after left hemicolectomy with LAR, 20 underwent tension-free anastomosis, including eight, six, and six patients undergoing right colonic transposition, retro-ileal anastomosis through an ileal mesenteric defect, and an additional colic artery division, respectively. Colorectal anastomosis is feasible for patients with extended colonic resection. Low anastomotic leakage and stoma rates can be achieved with careful attention to colonic mobilization and tension-free anastomosis. Full article
Show Figures

Figure 1

Figure 1
<p>Right colonic transposition for colorectal anastomotic techniques after left hemicolectomy with rectal low anterior resection. The right colon is rotated 180° in the sagittal plane and around the ileocecal pedicle axis or the right colon is rotated 180° in a counterclockwise manner in the frontal plane and around the superior mesenteric vessel axis.</p>
Full article ">Figure 2
<p>Retro-ileal anastomosis through an ileal mesenteric defect for colorectal anastomotic techniques after left hemicolectomy with rectal low anterior resection. The transverse or right colon is passed through a window created in an avascular area of the distal ileal mesentery to reach the rectal stump.</p>
Full article ">Figure 3
<p>Survival. (<b>a</b>) Progression-free survival. (<b>b</b>) Overall survival.</p>
Full article ">
22 pages, 15220 KiB  
Article
Hot Spot Analysis of YAP-TEAD Protein-Protein Interaction Using the Fragment Molecular Orbital Method and Its Application for Inhibitor Discovery
by Jongwan Kim, Hocheol Lim, Sungho Moon, Seon Yeon Cho, Minhye Kim, Jae Hyung Park, Hyun Woo Park and Kyoung Tai No
Cancers 2021, 13(16), 4246; https://doi.org/10.3390/cancers13164246 - 23 Aug 2021
Cited by 24 | Viewed by 4772
Abstract
The Hippo pathway is an important signaling pathway modulating growth control and cancer cell proliferation. Dysregulation of the Hippo pathway is a common feature of several types of cancer cells. The modulation of the interaction between yes-associated protein (YAP) and transcriptional enhancer associated [...] Read more.
The Hippo pathway is an important signaling pathway modulating growth control and cancer cell proliferation. Dysregulation of the Hippo pathway is a common feature of several types of cancer cells. The modulation of the interaction between yes-associated protein (YAP) and transcriptional enhancer associated domain (TEAD) in the Hippo pathway is considered an attractive target for cancer therapeutic development, although the inhibition of PPI is a challenging task. In order to investigate the hot spots of the YAP and TEAD1 interacting complex, an ab initio Fragment Molecular Orbital (FMO) method was introduced. With the hot spots, pharmacophores for the inhibitor design were constructed, then virtual screening was performed to an in-house library. Next, we performed molecular docking simulations and FMO calculations for screening results to study the binding modes and affinities between PPI inhibitors and TEAD1. As a result of the virtual screening, three compounds were selected as virtual hit compounds. In order to confirm their biological activities, cellular (luciferase activity, proximity ligation assay and wound healing assay in A375 cells, qRT-PCR in HEK 293T cells) and biophysical assays (surface plasmon resonance assays) were performed. Based on the findings of the study, we propose a novel PPI inhibitor BY03 and demonstrate a profitable strategy to analyze YAP–TEAD PPI and discover novel PPI inhibitors. Full article
(This article belongs to the Special Issue Hippo Signaling Pathway in Cancers)
Show Figures

Figure 1

Figure 1
<p>PPIs map of YAP-TEAD1. The 3D-SPIE-based interaction maps consist of interface 2 (<b>left</b>) and interface 3 (<b>right</b>) YAP–TEAD1 PPIs. The main interactions in the interface 2 of YAP-TEAD1 are shown by blue lines (<b>left</b>). The orange box represents the LxxLF motif in YAP, and the blue box represents binding partner residues in TEAD1 for the YAP LxxLF motif. The main interactions in interface 3 are represented by green lines (<b>right</b>). Left gray lines indicate interface 1, and right gray lines indicate amino acids excluded from interface 3. All interactions (shown in this map) whose magnitudes are ignored have attractive PIE values more stable than −3.0 kcal/mol.</p>
Full article ">Figure 2
<p>FMO analysis for interface 2 of YAP–TEAD1 PPIs. (<b>A</b>) Hot spot residues of YAP, namely, Asp64, Leu65, Glu66, Leu68, and Phe69, in interface 2 are shown as green sticks, Yellow indicates YAP alpha-helix and dark pink TEAD, and Raspberry color represents the TEAD1 interaction domain for YAP hot spot residues in interface 2. (<b>B</b>) TEAD1 hot spot residues, namely, Ser313, Val318, Tyr346, Lys353, Phe314, Val366, and Asn369, are shown as magenta sticks. The blue box indicates the LxxLF hydrophobic motif in YAP. The detailed PIEs are described in the right panel.</p>
Full article ">Figure 3
<p>FMO analysis for interface 3 of YAP–TEAD1 PPIs. (<b>A</b>) Hot spot residues of YAP, namely, Thr83, Val84, Met86, Arg87, Arg89, Lys90, Leu91, Pro92, Ser94, Phe95, Phe96, Pro98, and Pro99, are shown as green sticks. The green loop indicates YAP, the light greyish-blue TEAD1, and raspberry color indicates the TEAD1 interaction domain for YAP hot spot residues in interface 3. (<b>B</b>) TEAD1 hot spot residues in interface 3 are shown as orange sticks. Details of PIEs are described in the right panel.</p>
Full article ">Figure 4
<p>Pharmacophore models derived from the hot spot at interface 2 of YAP–TEAD PPIs. R, aromatic ring; H, hydrophobic; A, hydrogen bond acceptor. The green helix and sticks represent the YAP region, and the light blue stick represents the TEAD1 region.</p>
Full article ">Figure 5
<p>Structure of the three selected virtual hits from FMO-based virtual screening. (<b>A</b>) The structure of BY01, BY02, BY03. B. Docking poses in the TEAD1 binding pocket of BY01, (<b>B</b>); BY02, (<b>C</b>) and BY03, (<b>D</b>) using Glide in the Schrödinger suite.</p>
Full article ">Figure 6
<p>TEAD reporter luciferase activity observed in A375 cells treated with 10 nM to 10 μM of hit compounds with FA (flufenamic acid) 24 h post-transfection (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>TEAD luciferase activity inhibition observed in A375 cells treated with hit compounds and FA (flufenamic acid) 24 h post-transfection (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 8
<p>SPR analysis of BY03 to TEAD1. Sensorgrams of BY03 binding to the TEAD1. In the SPR experiment, TEAD1 was immobilized onto the vertical channels of a NeutrAvidin-precoated ProteOn GLH sensor chip, and BY03 was injected into the horizontal channels of the sensor chip at different concentrations of 0, 6.25, 12.5, 25, 50, and 100 μM.</p>
Full article ">Figure 9
<p>Fluorescence signals generated by proximity ligation assay demonstrate that endogenous TEAD and YAP are closely located in the nucleus at MSTO-211H. (<b>A</b>) Representative fluorescence images of nuclear YAP-TEAD PLA signals from assay (Scale bars, 10 μm). (<b>B</b>) Stick graph presentation by % of cells with corresponding PLA signal ranges per nucleus.</p>
Full article ">Figure 10
<p>Expression of YAP-TEAD target genes (CYR61 and AMOT) in HEK293T cells incubated in the presence of BY03 at 1 μM, 3 μM, 15 μM, respectively for 24 h. Flufenamic acid (FA) was treated with 15 μM and 100 μM, respectively. mRNAs were measured by qRT-PCR (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 11
<p>FMO results for BY03 in complex with interface 2 in TEAD1. PIEDA results of BY01 and BY02 are shown in <a href="#app1-cancers-13-04246" class="html-app">Table S4</a> and S5, respectively. (<b>A</b>) The structure of BY03 binding to the interface 2 of TEAD1. The ligand is shown in light pink, the protein residues are shown in cyan, and the nitrogen and oxygen atoms are shown in blue and red, respectively. (<b>B</b>) The left bar plot describes the PIEs of the significant residues in the interface2 of TEAD1, while the right bar plot describes the PIEDA of these critical interactions. The electrostatic, exchange repulsion, charge transfer with higher order mixed term, dispersion, and solvation energy terms are shown in yellow, green, red, blue and purple, respectively.</p>
Full article ">Figure 12
<p>Overlay structure of BY03 to YAP at interface 2. The pink color represents BY03 and the green color represents YAP. Grey letters represent TEAD1 binding residues in interface 2 for BY03.</p>
Full article ">Scheme 1
<p>Virtual screening process for the screening of the YAP-TEAD1 inhibitor.</p>
Full article ">
21 pages, 1797 KiB  
Review
Identification and Targeting of Mutant Peptide Neoantigens in Cancer Immunotherapy
by Daniel J. Verdon and Misty R. Jenkins
Cancers 2021, 13(16), 4245; https://doi.org/10.3390/cancers13164245 - 23 Aug 2021
Cited by 18 | Viewed by 7230
Abstract
In recent decades, adoptive cell transfer and checkpoint blockade therapies have revolutionized immunotherapeutic approaches to cancer treatment. Advances in whole exome/genome sequencing and bioinformatic detection of tumour-specific genetic variations and the amino acid sequence alterations they induce have revealed that T cell mediated [...] Read more.
In recent decades, adoptive cell transfer and checkpoint blockade therapies have revolutionized immunotherapeutic approaches to cancer treatment. Advances in whole exome/genome sequencing and bioinformatic detection of tumour-specific genetic variations and the amino acid sequence alterations they induce have revealed that T cell mediated anti-tumour immunity is substantially directed at mutated peptide sequences, and the identification and therapeutic targeting of patient-specific mutated peptide antigens now represents an exciting and rapidly progressing frontier of personalized medicine in the treatment of cancer. This review outlines the historical identification and validation of mutated peptide neoantigens as a target of the immune system, and the technical development of bioinformatic and experimental strategies for detecting, confirming and prioritizing both patient-specific or “private” and frequently occurring, shared “public” neoantigenic targets. Further, we examine the range of therapeutic modalities that have demonstrated preclinical and clinical anti-tumour efficacy through specifically targeting neoantigens, including adoptive T cell transfer, checkpoint blockade and neoantigen vaccination. Full article
Show Figures

Figure 1

Figure 1
<p>Generation and recognition of neoantigenic peptides after mutational or structural changes to somatic DNA. Changes to coding nucleotide sequence can be generated by non-synonymous point mutations, insertion/deletion events leading to reading frameshifts, or larger-scale structural changes such as chromosomal translocation and gene fusion events. When these changes to somatic DNA cause an alteration in amino acid sequence creating a peptide that can be processed and presented in the context of MHC-I/II and induce TCR activation, such a peptide is designated a neoantigen.</p>
Full article ">Figure 2
<p>Overview of neoantigen identification and validation pipelines. Matched whole-genome or whole-exome sequencing data is used to identify sequence differences between healthy and tumour-derived DNA. Concurrently, the eluted tumour MHC-I/II ligandome can be assessed via LC-MS, and tumour RNA expression data is collected to validate the presence and expression level of mutated transcripts. Resulting putative neoantigenic peptide sequences are triaged and prioritized using bioinformatic tools to inform likely proteasomal processing, degree of difference from matched wild-type peptide, and MHC binding affinity and stability based on donor <span class="html-italic">HLA</span> haplotype. Finally, selected candidates are expressed as tandem minigene libraries or synthesized as long peptides and the capacity of patient T cells to recognize transduced or peptide-loaded autologous APC and autologous tumour cells is assessed. Algorithms listed are cited in-text.</p>
Full article ">
10 pages, 1819 KiB  
Article
Comparison of FLASH Proton Entrance and the Spread-Out Bragg Peak Dose Regions in the Sparing of Mouse Intestinal Crypts and in a Pancreatic Tumor Model
by Michele M. Kim, Ioannis I. Verginadis, Denisa Goia, Allison Haertter, Khayrullo Shoniyozov, Wei Zou, Amit Maity, Theresa M. Busch, James M. Metz, Keith A. Cengel, Lei Dong, Costas Koumenis and Eric S. Diffenderfer
Cancers 2021, 13(16), 4244; https://doi.org/10.3390/cancers13164244 - 23 Aug 2021
Cited by 75 | Viewed by 5099
Abstract
Ultra-high dose rate FLASH proton radiotherapy (F-PRT) has been shown to reduce normal tissue toxicity compared to standard dose rate proton radiotherapy (S-PRT) in experiments using the entrance portion of the proton depth dose profile, while proton therapy uses a spread-out Bragg peak [...] Read more.
Ultra-high dose rate FLASH proton radiotherapy (F-PRT) has been shown to reduce normal tissue toxicity compared to standard dose rate proton radiotherapy (S-PRT) in experiments using the entrance portion of the proton depth dose profile, while proton therapy uses a spread-out Bragg peak (SOBP) with unknown effects on FLASH toxicity sparing. To investigate, the biological effects of F-PRT using an SOBP and the entrance region were compared to S-PRT in mouse intestine. In this study, 8–10-week-old C57BL/6J mice underwent 15 Gy (absorbed dose) whole abdomen irradiation in four groups: (1) SOBP F-PRT, (2) SOBP S-PRT, (3) entrance F-PRT, and (4) entrance S-PRT. Mice were injected with EdU 3.5 days after irradiation, and jejunum segments were harvested and preserved. EdU-positive proliferating cells and regenerated intestinal crypts were quantified. The SOBP had a modulation (width) of 2.5 cm from the proximal to distal 90%. Dose rates with a SOBP for F-PRT or S-PRT were 108.2 ± 8.3 Gy/s or 0.82 ± 0.14 Gy/s, respectively. In the entrance region, dose rates were 107.1 ± 15.2 Gy/s and 0.83 ± 0.19 Gy/s, respectively. Both entrance and SOBP F-PRT preserved a significantly higher number of EdU + /crypt cells and percentage of regenerated crypts compared to S-PRT. Moreover, tumor growth studies showed no difference between SOBP and entrance for either of the treatment modalities. Full article
(This article belongs to the Section Cancer Therapy)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic and (<b>b</b>) photo of the beam line design with ridge filter and range shifters to deliver a spread-out Bragg peak (SOBP) in the mouse. (<b>c</b>) Model of ridge filter and (<b>d</b>) photo of the printed ridge filter with protective casing.</p>
Full article ">Figure 2
<p>(<b>a</b>) EBT3 Gafchromic film irradiated at the mid-SOBP with a 2 × 2 cm<sup>2</sup> field size. x and y lateral profiles are shown on the right. (<b>b</b>) Relative depth dose of SOBP used to irradiate mice measured by the multi-layer ionization chamber (MLIC). Mice were placed in a holder in order to be within the shaded blue region. Film (circles) and multi-layer ion chamber measurements (solid line) agreed. The red x indicates the point at which absolute dose measurements were performed for the SOBP studies. (<b>c</b>) Surface plot of relative depth dose measured by a stack of 20 films with 2 to10 mm pieces of solid water plastic between films.</p>
Full article ">Figure 3
<p>Relative depth dose measurement with a ridge filter in the beam line at various angles. The inset plot shows a zoom of the SOBP region. All studies were performed with the ridge filter at 2.5° for a flat SOBP.</p>
Full article ">Figure 4
<p>(<b>a</b>) Representative images of EdU (green) staining in frozen jejunum sections at 3.5 days post 15 Gy of whole abdominal irradiation (10× magnification; scale bar 100 mm) with FLASH proton radiotherapy (F-PRT) or standard proton radiotherapy (S-PRT) dose rate protons in the entrance and SOBP region of the depth dose profile. (<b>b</b>) Quantification of EdU+ cells per crypt. (<b>c</b>) Quantification of the % regenerated crypts. Data presented as mean ± SEM; n.s. = not significant.</p>
Full article ">Figure 5
<p>(<b>a</b>) Tumor growth curves of mice following 18 Gy focal irradiation with protons at standard (S-PRT Entrance, solid red; S-PRT SOBP, dotted red) versus FLASH (F-PRT Entrance, solid green; F-PRT SOBP, dotted green) dose rates. Unirradiated mice served as a control group (NR, blue). Black arrow indicates the time of irradiation. Data expressed as mean ± SEM. (<b>b</b>) Kaplan–Meier survival analysis of the mice from (<b>a</b>).</p>
Full article ">
15 pages, 5864 KiB  
Article
Adhesion of Platelets to Colon Cancer Cells Is Necessary to Promote Tumor Development in Xenograft, Genetic and Inflammation Models
by Marica Cariello, Elena Piccinin, Roberta Zerlotin, Marilidia Piglionica, Claudia Peres, Chiara Divella, Anna Signorile, Gaetano Villani, Giuseppe Ingravallo, Carlo Sabbà and Antonio Moschetta
Cancers 2021, 13(16), 4243; https://doi.org/10.3390/cancers13164243 - 23 Aug 2021
Cited by 7 | Viewed by 2745
Abstract
Platelets represent the linkage between tissue damage and inflammatory response with a putative role in tumorigenesis. Given the importance of the microenvironment in colon cancer development, we elucidated the eventual role of platelets-cancer cells crosstalk in in vivo colon cancer models. To evaluate [...] Read more.
Platelets represent the linkage between tissue damage and inflammatory response with a putative role in tumorigenesis. Given the importance of the microenvironment in colon cancer development, we elucidated the eventual role of platelets-cancer cells crosstalk in in vivo colon cancer models. To evaluate the involvement of platelets in intestinal tumorigenesis, we first analyzed if the ablation of β-integrin P-selectin that drives platelets-cell adhesion, would contribute to platelets-colon cancer cell interaction and drive cancer progression. In a xenograft tumor model, we observed that when tumors are inoculated with platelets, the ablation of P-selectin significantly reduced tumor growth compared to control platelets. Furthermore, in genetic models, as well as in chronic colitis-associated colorectal carcinogenesis, P-selectin ablated mice displayed a significant reduction in tumor number and size compared to control mice. Taken together, our data highlights the importance of platelets in the tumor microenvironment for intestinal tumorigenesis. These results support the hypothesis that a strategy aimed to inhibit platelets adhesion to tumor cells are able to block tumor growth and could represent a novel therapeutic approach to colon cancer treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Platelets-intestinal cells crosstalk in human intestine. (<b>A</b>) Histology was assessed by H&amp;E staining and was observed by light microscopy (magnification, 200×) in normal ileum, colon and colon adenocarcinoma. Representative specimens are shown. (<b>B</b>) The protein expression of P-selectin and CD41 was investigated by immunofluorescence and confocal microscopy analysis in paraffin-embedded sections from normal portions of ileum and colon and colon adenocarcinoma. P-selectin in green and CD41 in red. Representative images are shown. Magnification 630×. (<b>C</b>) CD-41/P-selectin positive platelets were quantified by ImageJ software. Results are expressed as mean ± SEM (* <span class="html-italic">p</span> &lt; 0.05). (<b>D</b>) The protein expression of P-selectin and CD44 was investigated by immunofluorescence and confocal microscopy analysis in paraffin-embedded sections from normal portions of colon and colon adenocarcinoma. P-selectin in green and CD44 in red. Representative images are shown. Magnification 400×.</p>
Full article ">Figure 2
<p>P-selectin KO platelets injection reduced tumor growth in xenograft tumor model. (<b>A</b>) Athymic nu/nu mice were injected subcutaneously with HT29 cells and platelets isolated from P-sel<sup>−/−</sup> and P-sel<sup>+/+</sup> mice were administered directly into the tumor mass every 7 days. Gross morphology of HT-29 cells treated with P-sel<sup>+/+</sup> platelets and HT-29 cells treated with P-sel<sup>−/−</sup>. (<b>B</b>) Tumor growth (%) curves showed a reduction in the expansion of tumors injected with P-sel<sup>−/−</sup> platelets (<span class="html-italic">n</span> = 10 mice per group). (<b>C</b>) Histology was assessed by H&amp;E staining and was observed by light microscopy (magnification, 200×). Representative specimens are shown. (<b>D</b>) Tumor weight (gr) was reported. The results are expressed as mean ± SEM, <span class="html-italic">n</span> = 20 tumors per group; * <span class="html-italic">p</span> ≤ 0.05. (<b>E</b>) Paraffin-embedded tumor specimens from HT-29 cells treated with P-sel<sup>+/+</sup> platelets and HT-29 cells treated with P-sel<sup>−/−</sup> platelets were immunoassayed with CD44 antibody (200× magnification). Representative specimens are shown. (<b>F</b>) The protein expression of P-selectin and CD41 was investigated by immunofluorescence and confocal microscopy analysis in paraffin-embedded sections from tumors injected with P-sel<sup>−/−</sup> or control platelets. P-selectin in green and CD41 in red. Representative images are shown. Magnification 630×.</p>
Full article ">Figure 3
<p>Treatment with P-selectin KO platelets reduced cell cycle progression in xenograft tumor model. Paraffin-embedded tumor specimens from HT-29 cells treated with P-sel<sup>+/+</sup> platelets and HT-29 cells treated with P-sel<sup>−/−</sup> platelets were immunoassayed with (<b>A</b>) cyclin D1 antibody (200× and 300× magnification) and (<b>B</b>) anti-Pcna antibody (200× and 300× magnification). Representative specimens are shown. Cyclin D1 and Pcna staining per field was quantified by ImageJ software and reported as percentage per field. Comparison of HT-29 cells treated with P-sel<sup>+/+</sup> platelets and HT-29 cells treated with P-sel<sup>−/−</sup> platelets (<span class="html-italic">n</span> = 8/group) was performed using the Student’s t-test. Results are expressed as mean ± SEM (* <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Gene expression analysis of Ccne1, Pten and Il6 in HT-29 cells treated with P-sel<sup>+/+</sup> platelets and HT-29 cells treated with P-sel<sup>−/−</sup> platelets (<span class="html-italic">n</span>= 20 tumors per group). Cyclophilin was used as a housekeeping gene to normalize data. The results are expressed as mean ± SEM. Statistical significance (* <span class="html-italic">p</span> &lt; 0.05) was assessed by the Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Role of P-selectin in genetic mouse model of colon cancer. (<b>A</b>) Total number of tumors was counted in APC<sup>Min</sup>/P-sel<sup>−/−</sup> and APC<sup>Min</sup> mice. Results are the average ± standard error of the mean (SEM), <span class="html-italic">n</span> = 10 mice per group; * <span class="html-italic">p</span> ≤ 0.05. (<b>B</b>) Histology was assessed by H&amp;E staining and was observed by light microscopy (magnification, 200×). Representative specimens are shown. Paraffin-embedded tumor specimens from APC<sup>Min</sup>/P-sel<sup>−/−</sup> and APC<sup>Min</sup> mice were immunoassayed with (<b>C</b>) cyclin D1 antibody (100× and 200× magnification) and (<b>D</b>) anti-Pcna antibody (100× and 200× magnification). Representative specimens are shown. (<b>E</b>) Cyclin D1 and (<b>F</b>) Pcna staining per field was quantified by ImageJ software and reported as percentage per field. To perform protein quantification 10 representative images were taken with a 200× magnification for each sample. The percentage of stained area/total area was measured. Values from all consecutive images for each sample were averaged. Comparison of tumor specimens from APC<sup>Min</sup>/P-sel<sup>−/−</sup> and APC<sup>Min</sup> mice was performed using T student’s test (<span class="html-italic">n</span> = 5 mice per group). Results are expressed as mean ± SEM (* <span class="html-italic">p</span> &lt; 0.05). (<b>G</b>) Gene expression analysis of Ccne1, c-myc and Pten in APC<sup>Min</sup>/P-sel<sup>-/-</sup> and APC<sup>Min</sup> mice. Cyclophilin was used as a housekeeping gene to normalize data. The results are expressed as mean ± SEM. Statistical significance (* <span class="html-italic">p</span> &lt; 0.05) was assessed by the Student’s <span class="html-italic">t</span>-test (<span class="html-italic">n</span> = 10 mice per group).</p>
Full article ">Figure 5
<p>P-selectin KO mice are protected from chronic colitis-associated colorectal carcinogenesis. (<b>A</b>) P-sel<sup>-/-</sup> (<span class="html-italic">n</span> = 10) and P-sel<sup>+/+</sup> (wild-type; <span class="html-italic">n</span> = 10) mice were treated for a colitis-associated colorectal carcinogenesis model through one intraperitoneal azoxymethane (AOM) injection and 3 cycles of 2% dextran sulfate sodium (DSS) in drinking water. (<b>B</b>) Total number of tumors was counted. The diameter of each tumor was measured. The number of tumors with a diameter &lt;5 mm and &gt;5 mm was significantly reduced in P-sel<sup>-/-</sup> mice showing that these mice are protected from colon cancer formation. Results are the average ± standard error of the mean (SEM); * <span class="html-italic">p</span> ≤ 0.05 (<b>C</b>) Gene expression analysis of c-myc in P-sel<sup>-/-</sup> and P-sel<sup>+/+</sup> mice. Cyclophilin was used as a housekeeping gene to normalize data. The results are expressed as mean ± SEM, * <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">
32 pages, 4143 KiB  
Review
Emerging Principles in the Transcriptional Control by YAP and TAZ
by Alejandro Lopez-Hernandez, Silvia Sberna and Stefano Campaner
Cancers 2021, 13(16), 4242; https://doi.org/10.3390/cancers13164242 - 23 Aug 2021
Cited by 28 | Viewed by 5481
Abstract
Yes-associated protein (YAP) and TAZ are transcriptional cofactors that sit at the crossroad of several signaling pathways involved in cell growth and differentiation. As such, they play essential functions during embryonic development, regeneration, and, once deregulated, in cancer progression. In this review, we [...] Read more.
Yes-associated protein (YAP) and TAZ are transcriptional cofactors that sit at the crossroad of several signaling pathways involved in cell growth and differentiation. As such, they play essential functions during embryonic development, regeneration, and, once deregulated, in cancer progression. In this review, we will revise the current literature and provide an overview of how YAP/TAZ control transcription. We will focus on data concerning the modulation of the basal transcriptional machinery, their ability to epigenetically remodel the enhancer–promoter landscape, and the mechanisms used to integrate transcriptional cues from multiple pathways. This reveals how YAP/TAZ activation in cancer cells leads to extensive transcriptional control that spans several hallmarks of cancer. The definition of the molecular mechanism of transcriptional control and the identification of the pathways regulated by YAP/TAZ may provide therapeutic opportunities for the effective treatment of YAP/TAZ-driven tumors. Full article
(This article belongs to the Special Issue YAP (Yes-Associated Protein) in Cancer)
Show Figures

Figure 1

Figure 1
<p>YAP/TAZ orchestrate transcriptional control by interacting with other transcription factors (TFs). A schematic overview of the general mechanisms by which YAP/TAZ can integrate with other transcription factors in order to modulate complex transcriptional responses. (<b>a</b>) Regulation of YAP/TAZ targets by <span class="html-italic">cis</span> interactions with other TFs; (<b>b</b>) regulation of other TFs targets by YAP/TAZ; (<b>c</b>) regulation of YAZ/TAZ activity on their targets mediated only by protein–protein chromatin-associated complexes; (<b>d</b>) modulation of other TFs activity, on their respective targets, mediated only by protein–protein chromatin-associated complexes; (<b>e</b>) modulation based on non-nuclear protein–protein interaction. Please note that although DNA regulatory elements have been placed near the transcribed genes, it is implied that these modes of regulation apply to both promoters and enhancers. Created with BioRender.com. (<a href="https://help.biorender.com/en/articles/3619405-how-do-i-cite-biorender" target="_blank">https://help.biorender.com/en/articles/3619405-how-do-i-cite-biorender</a>, accessed on 23 July 2021).</p>
Full article ">Figure 2
<p>Cancer cell programs controlled by YAP and TAZ. This figure summarizes the biological processes controlled by YAP and TAZ in cancer cells. For each program, key downstream genes/pathways are reported according to their documented role as effectors of the indicated biological program. For the sake of simplicity, genes are all represented as regulated by YAP/YAZ, but it is implied that in many cases, the expression of these genes depends on the integration with other transcription factors, as detailed in the text. Created with BioRender.com. (<a href="https://help.biorender.com/en/articles/3619405-how-do-i-cite-biorender" target="_blank">https://help.biorender.com/en/articles/3619405-how-do-i-cite-biorender</a>, accessed on 23 July 2021).</p>
Full article ">
18 pages, 5238 KiB  
Article
Helicobacter pylori CagA Induces Cortactin Y-470 Phosphorylation-Dependent Gastric Epithelial Cell Scattering via Abl, Vav2 and Rac1 Activation
by Nicole Tegtmeyer, Aileen Harrer, Klemens Rottner and Steffen Backert
Cancers 2021, 13(16), 4241; https://doi.org/10.3390/cancers13164241 - 23 Aug 2021
Cited by 12 | Viewed by 2834
Abstract
The pathogen Helicobacter pylori is the first reported bacterial type-1 carcinogen playing a role in the development of human malignancies, including gastric adenocarcinoma. Cancer cell motility is an important process in this scenario, however, the molecular mechanisms are still not fully understood. Here, [...] Read more.
The pathogen Helicobacter pylori is the first reported bacterial type-1 carcinogen playing a role in the development of human malignancies, including gastric adenocarcinoma. Cancer cell motility is an important process in this scenario, however, the molecular mechanisms are still not fully understood. Here, we demonstrate that H. pylori subverts the actin-binding protein cortactin through its type-IV secretion system and injected oncoprotein CagA, e.g., by inducing tyrosine phosphorylation of cortactin at Y-470, which triggers gastric epithelial cell scattering and motility. During infection of AGS cells, cortactin was discovered to undergo tyrosine dephosphorylation at residues Y-421 and Y-486, which is mediated through inactivation of Src kinase. However, H. pylori also profoundly activates tyrosine kinase Abl, which simultaneously phosphorylates cortactin at Y-470. Phosphorylated cortactin interacts with the SH2-domain of Vav2, a guanine nucleotide exchange factor for the Rho-family of GTPases. The cortactin/Vav2 complex then stimulates a previously unrecognized activation cascade including the small GTPase Rac1, to effect actin rearrangements and cell scattering. We hypothesize that injected CagA targets cortactin to locally open the gastric epithelium in order to get access to certain nutrients. This may disturb the cellular barrier functions, likely contributing to the induction of cell motility, which is important in gastric cancer development. Full article
Show Figures

Figure 1

Figure 1
<p>Cortactin undergoes dephosphorylation at tyrosine residues Y-421 and Y-486, and phosphorylation at Y-470 during <span class="html-italic">H. pylori</span> infection. (<b>A</b>) Schematic representation of the cortactin domain structure. Three tyrosine phosphorylation sites in human cortactin are shown on top, their mouse counterparts on the bottom. (<b>B</b>) AGS cells were co-incubated with wt <span class="html-italic">H. pylori</span> or Δ<span class="html-italic">cagA</span> mutant for 6 h. Mock control and infected cells were harvested and separated into cytoplasmic and membrane fractions. The loading control blots for cortactin and integrin-β<sub>1</sub> are shown. (<b>C</b>) Phosphorylation of cortactin at tyrosine residues 421, 470 and 486 was investigated with the indicated phosphotyrosine-specific antibodies. (<b>D</b>) Densitometric quantification of cortactin tyrosine phosphorylation bands of Western blots in panel C. Error bars represent +/− standard deviation (SD). The results are from three independent experiments. <span class="html-italic">p</span>-values of ≤0.001 (***) were considered as statistically significant.</p>
Full article ">Figure 2
<p>Cortactin Y-470 is phosphorylated by activated Abl kinase and forms a physical complex with Vav2 and Rac1, associated with cell motility. (<b>A</b>) AGS cells were infected for 6 h with <span class="html-italic">H. pylori</span> in the presence or absence of the Abl-specific inhibitors imatinib and SKI-DV2-43. Cell lysates were probed with indicated antibodies, demonstrating the activated Abl phosphorylated cortactin at Y-470 (arrows). (<b>B</b>) Densitometric quantification of phospho-band intensities for Abl (Y-412) and cortactin (Y-470). (<b>C</b>) Phase contrast microscopy revealed that <span class="html-italic">H. pylori</span> induces cell scattering, which is abolished by the Abl inhibitor imatinib. 400× magnification, scale bars represent 10 µm (<b>D</b>) AGS cells were co-incubated with wt <span class="html-italic">H. pylori</span> or Δ<span class="html-italic">cagA</span> mutant for 6 h, followed by immunoprecipitation (IP) using cortactin antibodies. Resulting blots were probed with antibodies as indicated. The results show that cortactin phosphorylated at Y-470 interacts with Vav2 and Rac1 in a CagA-dependent manner. (<b>E</b>) Band intensities of Rac1 and Vav2 bound to phosphorylated cortactin were quantified using the Image Lab software version 6.1. Error bars represent +/− SD. The results are from three independent experiments. <span class="html-italic">p</span>-values of ≤0.001 (***) were considered as statistically significant.</p>
Full article ">Figure 3
<p>Vav2 forms a physical complex with Rac1 and cortactin phosphorylated at Y-470. (<b>A</b>) AGS cells were infected for 6 h with wt <span class="html-italic">H. pylori</span> or Δ<span class="html-italic">cagA</span> mutant, followed by immunoprecipitation (IP) using Vav2 antibodies. The corresponding Western blots were probed with antibodies as indicated. The results suggest that Vav2 interacts with Rac1 and cortactin phosphorylated at Y-470 in a CagA-dependent fashion. (<b>B</b>) Band intensities of Rac1 and phosphorylated cortactin bound to precipitated Vav2 were quantified using the Image Lab Software (Bio-Rad). Error bars represent +/− SD. <span class="html-italic">p</span>-value of ≤0.001 (***) was considered as statistically significant.</p>
Full article ">Figure 4
<p>Cortactin phosphorylation at Y-470 temporally correlating with Rac1 activation during <span class="html-italic">H. pylori</span> infection. (<b>A</b>) AGS cells were infected with <span class="html-italic">H. pylori</span> wt or Δ<span class="html-italic">cagA</span> mutant for different times, i.e., 60, 120 and 240 min, followed by probing with α-cortactin-PY-470 (top) or Rac1-GTP pulldown and α-Rac1 Western blot (middle). The α-Rac1 blot of total cell lysates served as loading control (bottom). (<b>B</b>) Quantification of band intensities on Western blots showing cortactin levels phosphorylated at Y-470 and relative Rac1-GTP activity. The results are from three independent experiments. <span class="html-italic">p</span>-values of ≤0.01 (**) and ≤0.001 (***) were considered as statistically significant. The results confirm that the induction of cortactin phosphorylation at Y-470 and activation of Rac1 display a temporal correlation.</p>
Full article ">Figure 5
<p>Vav2 binds via its SH2 domain to cortactin phosphorylated at Y-470. (<b>A</b>) Schematic representation of the Vav2 domain structure.. (<b>B</b>) AGS cells were transfected with GFP-Vav2 wt and various GFP-Vav2 mutant constructs, as indicated, for 48 h, followed by infection with <span class="html-italic">H. pylori</span> for 6 h. Cell lysates were harvested and subjected to IP using α-GFP antibodies to precipitate GFP-Vav2. Similarly, presence of each construct in the IPs was confirmed by α-GFP antibody Western blot (top). This blot was then re-probed with α-cortactin antibody. (<b>C</b>) Densitometric quantification of cortactin band intensities bound to precipitated GFP-Vav2. (<b>D</b>) AGS cells were transfected with GFP-cortactin wt and various GFP-cortactin phospho-mutant constructs, as indicated, for 48 h in the presence or absence of HA-tagged Vav2 plasmid. Total cell lysates were probed with α-GFP antibodies showing similar amounts of GFP-cortactin proteins present in each lane (top). Parallel samples were subjected to IP using α-HA antibodies to precipitate HA-tagged Vav2 (middle). This blot was then re-probed with α-GFP antibodies to detect bound cortactin (bottom). (<b>E</b>) Densitometric quantification of cortactin band intensities bound to precipitated HA-Vav2. The results show that cortactin and Vav2 can bind to each other by a typical phosphotyrosine-SH2 domain interaction. Error bars represent +/− standard deviation (SD). The results are from three independent experiments. <span class="html-italic">p</span>-values of ≤0.05 (*), and ≤0.001 (***) were considered as statistically significant. n.s.: not significant.</p>
Full article ">Figure 6
<p>Control blots for the knockdown of cortactin and Vav2 by siRNAs and quantification of Rac1-GTP levels in AGS cells infected with <span class="html-italic">H. pylori</span>. (<b>A</b>) AGS cells were grown, treated with siRNAs and then infected with <span class="html-italic">H. pylori</span> wt, as described in <a href="#cancers-13-04241-f007" class="html-fig">Figure 7</a>. The resulting cell lysates were probed with the indicated antibodies to demonstrate successful downregulation of cortactin and Vav2 expression, respectively, and equal protein loading in each lane. (<b>B</b>) A second set of samples, as shown in panel A, were subjected to Rac1-GTP pulldowns. The resulting α-Rac1 blot (bottom) was quantified by densitometric evaluation of the bands employing the luminescence imager. The sample showing the strongest Rac1 signal corresponds to 100% Rac1 activity in the respective blot. Inhibition of Rac1-GTP levels by knockdown of cortactin or Vav2 were quantified in three independent experiments. <span class="html-italic">p</span>-values of ≤0.001 (***) were considered to report on statistically significant differences.</p>
Full article ">Figure 7
<p>Importance of cortactin, Vav2 and Rac1 expression as well as cortactin tyrosine phosphorylation for AGS cell migration during infection with <span class="html-italic">H. pylori</span>. (<b>A</b>) Confluent cell monolayers of AGS cells were transfected with siRNAs against cortactin, Vav2 and Rac1, scrambled siRNAs or Cortactin-Y466D, Cortactin-Y466F and vector control constructs. Then, cells were wounded using a pipette tip, followed by infection with <span class="html-italic">H. pylori</span> for 48 h (middle and bottom) or treatment with PBS buffer as control (top). All images at 100× magnification, scale bar represents 100 µm. (<b>B</b>) The areas following wound closure were measured after 48 h of infection. Error bars represent +/− standard deviation (SD). The results are from three independent experiments. <span class="html-italic">p</span>-values of ≤0.001 (***) were considered as statistically highly significant. (<b>C</b>) Proposed signal transduction cascade of injected CagA leading successively to activation of Abl kinase, phosphorylation of cortactin at Y-470, Vav2 binding, Rac1 GTPase activation and cell motility.</p>
Full article ">
17 pages, 921 KiB  
Review
The Role of Circular RNAs in Keratinocyte Carcinomas
by Thomas Meyer, Michael Sand, Lutz Schmitz and Eggert Stockfleth
Cancers 2021, 13(16), 4240; https://doi.org/10.3390/cancers13164240 - 23 Aug 2021
Cited by 3 | Viewed by 2842
Abstract
Keratinocyte carcinomas (KC) include basal cell carcinomas (BCC) and cutaneous squamous cell carcinomas (cSCC) and represents the most common cancer in Europe and North America. Both entities are characterized by a very high mutational burden, mainly UV signature mutations. Predominately mutated genes in [...] Read more.
Keratinocyte carcinomas (KC) include basal cell carcinomas (BCC) and cutaneous squamous cell carcinomas (cSCC) and represents the most common cancer in Europe and North America. Both entities are characterized by a very high mutational burden, mainly UV signature mutations. Predominately mutated genes in BCC belong to the sonic hedgehog pathway, whereas, in cSCC, TP53, CDKN2A, NOTCH1/2 and others are most frequently mutated. In addition, the dysregulation of factors associated with epithelial to mesenchymal transition (EMT) was shown in invasive cSCC. The expression of factors associated with tumorigenesis can be controlled in several ways and include non-coding RNA molecules, such as micro RNAs (miRNA) long noncoding RNAs (lncRNA) and circular RNAs (circRNA). To update findings on circRNA in KC, we reviewed 13 papers published since 2016, identified in a PubMed search. In both BCC and cSCC, numerous circRNAs were identified that were differently expressed compared to healthy skin. Some of them were shown to target miRNAs that are also dysregulated in KC. Moreover, some studies confirmed the biological functions of individual circRNAs involved in cancer development. Thus, circRNAs may be used as biomarkers of disease and disease progression and represent potential targets of new therapeutic approaches for KC. Full article
(This article belongs to the Special Issue Circular RNAs: New Insights into the Molecular Biology of Cancer)
Show Figures

Figure 1

Figure 1
<p>micro-RNA–circular RNA network to control mRNA expression in tumourigenesis. Expression of tumour suppressor genes and oncogenes is regulated in multiple ways, including canonical/alternative splicing of primary transcripts, resulting in mRNA and circular RNA (circRNA), respectively (1). The expression of mRNA is also regulated in multiple ways including, inhibition by micro-RNAs (miRNA) with complementary seed sequences (2). miRNAs may be inhibited by circRNAs, with specific MRE sequences sponging these miRNAs (3). They may derive from the same gene encoding the mRNA targeted by miRNAs, but also from other genes.</p>
Full article ">
10 pages, 821 KiB  
Article
CD200 Baseline Serum Levels Predict Prognosis of Chronic Lymphocytic Leukemia
by Giovanni D’Arena, Candida Vitale, Marta Coscia, Daniela Lamorte, Giuseppe Pietrantuono, Francesca Perutelli, Fiorella D’Auria, Teodora Statuto, Luciana Valvano, Annamaria Tomasso, Valentina Griggio, Rebecca Jones, Giovanna Mansueto, Oreste Villani, Simona D’Agostino, Vito Viglioglia, Vincenzo De Feo, Fabrizio Calapai, Carmen Mannucci, Alessandro Sgambato, Dimitar G. Efremov and Luca Laurentiadd Show full author list remove Hide full author list
Cancers 2021, 13(16), 4239; https://doi.org/10.3390/cancers13164239 - 23 Aug 2021
Cited by 4 | Viewed by 2894
Abstract
Membrane-bound CD200 is overexpressed in chronic lymphocytic leukemia (CLL), and there is some evidence that its soluble ectodomain (sCD200) could also be involved in the pathophysiology and the disease. However, very little is known about sCD200’s prognostic significance. sCD200 was tested at diagnosis [...] Read more.
Membrane-bound CD200 is overexpressed in chronic lymphocytic leukemia (CLL), and there is some evidence that its soluble ectodomain (sCD200) could also be involved in the pathophysiology and the disease. However, very little is known about sCD200’s prognostic significance. sCD200 was tested at diagnosis in 272 patients with CLL and in 78 age- and sex-matched healthy subjects using a specific human CD200 (OX-2 membrane glycoprotein) ELISA kit. A significantly higher concentration of sCD200 was found in CLL patients compared to controls. In our cohort, sCD200 was significantly higher in patients who were older than 66 years, with Binet stage C, unmutated IgVH and unfavorable (del11q or del17p) FISH. Time-to-first treatment and overall survival were significantly shorter in patients with higher sCD200 concentration, using as a cut-off 1281 pg/mL, the median value for sCD200 concentration in the whole CLL cohort. However, the prognostic impact of sCD200 was not confirmed in multivariate analysis. Baseline sCD200 values appeared to have an impact on the response to chemotherapy or chemo-immunotherapy, but not to targeted agents. Collectively, our data show that sCD200 serum levels correlate with more aggressive clinical and biological features and are able to predict a worse prognosis. This work supports the relevant role of CD200 not only as a diagnostic tool but also as a prognostic indicator and a potential therapeutic target in CLL. Full article
(This article belongs to the Special Issue Therapeutic Targets in Chronic Lymphocytic Leukemia)
Show Figures

Figure 1

Figure 1
<p>sCD200 in patients with CLL and in HD. Each dot represents the value for the sCD200 serum concentration from a single patient sample. Lines represent median values and interquartile ranges. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Time-to-first treatment for patients with sCD200 <sup>low</sup> vs. sCD200 <sup>high</sup>.</p>
Full article ">Figure 3
<p>Overall survival for patients with sCD200 <sup>low</sup> vs. sCD200 <sup>high</sup>.</p>
Full article ">Figure 4
<p>Pre- and post-treatment sCD200. Each dot represents the value for the sCD200 serum concentration from a single patient, assessed before and after chemo-immunotherapy treatment. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
14 pages, 3375 KiB  
Article
Adipsin-Dependent Secretion of Hepatocyte Growth Factor Regulates the Adipocyte-Cancer Stem Cell Interaction
by Masahiro Mizuno, Behnoush Khaledian, Masao Maeda, Takanori Hayashi, Seiya Mizuno, Eiji Munetsuna, Takashi Watanabe, Seishi Kono, Seiji Okada, Motoshi Suzuki, Shintaro Takao, Hironobu Minami, Naoya Asai, Fumihiro Sugiyama, Satoru Takahashi and Yohei Shimono
Cancers 2021, 13(16), 4238; https://doi.org/10.3390/cancers13164238 - 23 Aug 2021
Cited by 11 | Viewed by 4271
Abstract
Adipose tissue is a component of the tumor microenvironment and is involved in tumor progression. We have previously shown that adipokine adipsin (CFD) functions as an enhancer of tumor proliferation and cancer stem cell (CSC) properties in breast cancers. We established the Cfd-knockout [...] Read more.
Adipose tissue is a component of the tumor microenvironment and is involved in tumor progression. We have previously shown that adipokine adipsin (CFD) functions as an enhancer of tumor proliferation and cancer stem cell (CSC) properties in breast cancers. We established the Cfd-knockout (KO) mice and the mammary adipose tissue-derived stem cells (mADSCs) from them. Cfd-KO in mADSCs significantly reduced their ability to enhance tumorsphere formation of breast cancer patient-derived xenograft (PDX) cells, which was restored by the addition of Cfd in the culture medium. Hepatocyte growth factor (HGF) was expressed and secreted from mADSCs in a Cfd-dependent manner. HGF rescued the reduced ability of Cfd-KO mADSCs to promote tumorsphere formation in vitro and tumor formation in vivo by breast cancer PDX cells. These results suggest that HGF is a downstream effector of Cfd in mADSCs that enhances the CSC properties in breast cancers. Full article
(This article belongs to the Special Issue From Progression to Metastasis of Solid Cancer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Generation of Cfd-null mice. (<b>A</b>) Schematic representation of the <span class="html-italic">Cfd</span> locus in the mouse genome. <span class="html-italic">Cfd</span> gene is located on chromosome 10. Two sgRNAs were designed which targeted the upstream and downstream of the full coding sequence of the <span class="html-italic">Cfd</span> gene. PAM sequences are marked in bold. F1, R1, R2, and R3 represent the position of the primers used to genotype wild type (WT) and KO alleles. (<b>B</b>) PCR analysis of the genomic DNA obtained from the mouse tail at 3 weeks of age. The band for WT allele (primers used: F1 and R2, 576 bp) and that for KO allele (primers used: F1 and R3, 577 bp) were determined. Deletion of the Cfd sequence was confirmed by sequencing.</p>
Full article ">Figure 2
<p>Establishment of Cfd-KO mADSCs. (<b>A</b>) Microscopic appearance of WT and Cfd-KO mADSCs. Scale bar: 50 μm. (<b>B</b>) The expression level of <span class="html-italic">Cfd</span> mRNA in mADSCs. <span class="html-italic">Cfd</span> mRNA was undetectable in Cfd KO mADSCs. <span class="html-italic">GAPDH</span> was used as a control. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) The expression level of Cfd protein in differentiated mADSCs. Murine Cfd protein is composed of a mixture of glycosylated isoforms with two prominent bands observed between 30 and 50 kDa [<a href="#B27-cancers-13-04238" class="html-bibr">27</a>]. Cfd protein was undetectable in the cell lysate of Cfd-KO mADSC. Actin was used as a control. (<b>D</b>) Expression levels of cell surface markers in WT and KO mADSCs. Expression levels were analyzed using a flow cytometer. The percentage of cells positive for cell surface marker expression is presented.</p>
Full article ">Figure 3
<p>Cfd-KO in mADSCs suppressed the tumorsphere formation of cocultured breast cancer PDX cells. (<b>A</b>) Representative images of the tumorspheres formed by coculture of breast cancer PDX cells (PDX KUB06, KUB11, or KUB22) with WT or Cfd-KO mADSCs. Scale bar, 100 μm. The bar graph shows the number of tumorspheres (&gt;100 μm in diameter) formed by PDX cells cocultured with mADSCs. * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Cfd rescued the reduced ability of Cfd-KO mADSCs to induce sphere formation of breast cancer PDX cells. Representative images of the tumorspheres formed by coculture of breast cancer PDX cells (PDX KUB06, or KUB11) with Cfd-KO mADSCs with or without Cfd (9.5 μg/mL) in the culture medium were presented. Scale bar, 100 μm. The bar graph shows the number of tumorspheres (&gt;100 μm in diameter) formed by PDX cells cocultured with Cfd-KO mADSCs. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>HGF rescues the reduction of tumorsphere formation by Cfd-KO mADSCs. (<b>A</b>) Profiling of adipokines secreted from WT- and Cfd-KO mADSCs. Culture medium of mADSCs was analyzed using cytokine arrays. The intensity of the spot for Cfd in WT mADSCs was set at 1.0. (<b>B</b>) Reduction of the HGF mRNA levels in the Cfd-KO mADSCs. The addition of Cfd in the culture medium (Cfd, 9.5 μg/mL) upregulated the expression level of Cfd mRNA in the Cfd-KO mADSCs. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) HGF rescued the reduced ability of Cfd-KO ADSCs to induce sphere formation of breast cancer PDX cells. HGF (50 ng/mL) was added to the culture medium. The number of PDX tumorspheres (&gt;100 μm in diameter) were presented. Scale bar, 100 μm. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>HGF enhanced the tumor formation by breast cancer PDX cells co-injected with Cfd-KO mADSCs. (<b>A</b>) Schematic presentation of experimental procedures. Breast cancer PDX cells and mADSCs were cocultured for three days with or without HGF (50 ng/mL) and xenotransplanted to the mammary fat pad regions of the immunodeficient mice. (<b>B</b>) HGF enhanced the tumor formation by breast cancer PDX cells co-injected with Cfd-KO mADSCs. Four hundred thousand breast cancer PDX cells (KUB06) and 8 × 10<sup>4</sup> Cfd-KO mADSCs were cocultured with or without HGF (50 ng/mL) for 3 days in vitro and xenotransplanted in the mammary fat pad regions of the immunodeficient mice (<span class="html-italic">n</span> = 5). HGF or control PBS was subcutaneously injected twice a week. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) The appearance of the xenograft tumor. Scale bar, 10 mm. The bar graph shows the weight of tumors at day 28. * <span class="html-italic">p</span> &lt; 0.05. (<b>D</b>) Schematic illustration of the molecular functions of adipsin (Cfd) in the adipocyte-breast cancer cell interaction. 1. Anaphylatoxin C3a, the product of Cfd in the alternative complement pathway, functions as an activator of the CSC properties of breast cancer PDX cells [<a href="#B11-cancers-13-04238" class="html-bibr">11</a>]. 2. HGF is a downstream effector of Cfd secreted from mADSCs and enhances the CSC properties of breast cancer PDX cells.</p>
Full article ">
16 pages, 775 KiB  
Review
Epigenetic Changes Affecting the Development of Hepatocellular Carcinoma
by Ewa Wolinska and Maciej Skrzypczak
Cancers 2021, 13(16), 4237; https://doi.org/10.3390/cancers13164237 - 23 Aug 2021
Cited by 15 | Viewed by 3987
Abstract
Hepatocellular carcinoma (HCC) remains a serious oncologic issue with still a dismal prognosis. So far, no key molecular mechanism that underlies its pathogenesis has been identified. Recently, by specific molecular approaches, many genetic and epigenetic changes arising during HCC pathogenesis were detected. Epigenetic [...] Read more.
Hepatocellular carcinoma (HCC) remains a serious oncologic issue with still a dismal prognosis. So far, no key molecular mechanism that underlies its pathogenesis has been identified. Recently, by specific molecular approaches, many genetic and epigenetic changes arising during HCC pathogenesis were detected. Epigenetic studies revealed modified methylation patterns in HCC tumors, dysfunction of enzymes engaged in the DNA methylation process, and a set of histone modifications that influence gene expression. HCC cells are also influenced by the disrupted function of non-coding RNAs, such as micro RNAs and long non-coding RNAs. Moreover, a role of liver cancer stem cells in HCC development is becoming evident. The reversibility of epigenetic changes offers the possibility of influencing them and regulating their undesirable effects. All these data can be used not only to identify new therapeutic targets but also to predict treatment response. This review focuses on epigenetic changes in hepatocellular carcinoma and their possible implications in HCC therapy. Full article
(This article belongs to the Special Issue Molecular Mechanisms Involved in Hepatocarcinogenesis)
Show Figures

Figure 1

Figure 1
<p>Epigenetic processes implicated in hepatocellular carcinoma development. Dnmt DNA methyltransferase, HDAC histone deacetylase, siRNA small interfering RNA, lncRNAs long noncoding RNAs, and miRNAs micro RNAs.</p>
Full article ">
16 pages, 483 KiB  
Study Protocol
Prospective Evaluation of Radiotherapy-Induced Immunologic and Genetic Effects in Colorectal Cancer Oligo-Metastatic Patients with Lung-Limited Disease: The PRELUDE-1 Study
by Alessandro Ottaiano, Angela Petito, Mariachiara Santorsola, Valerio Gigantino, Maurizio Capuozzo, Daniela Fontanella, Rossella Di Franco, Valentina Borzillo, Sergio Buonopane, Vincenzo Ravo, Esmeralda Scipilliti, Giuseppe Totaro, Marcello Serra, Gianluca Ametrano, Roberta Penta, Fabiana Tatangelo, Giosuè Scognamiglio, Annabella Di Mauro, Maurizio Di Bonito, Maria Napolitano, Stefania Scala, Giuseppina Rea, Sara Santagata, Angela Lombardi, Anna Grimaldi, Carlo Caputo, Anna Crispo, Egidio Celentano, Gianfranco De Feo, Luisa Circelli, Giovanni Savarese, Raffaella Ruggiero, Francesco Perri, Vincenza Granata, Gerardo Botti, Michele Caraglia, Guglielmo Nasti and Paolo Mutoadd Show full author list remove Hide full author list
Cancers 2021, 13(16), 4236; https://doi.org/10.3390/cancers13164236 - 23 Aug 2021
Cited by 8 | Viewed by 2961
Abstract
Background: in recent years, the management of advanced colorectal cancer (CRC) has been greatly improved with integrated strategies including stereotactic radiation therapy (SRT). The administration of SRT has been demonstrated, particularly in oligo-metastatic (om) CRC, to be a safe and effective option. Interestingly, [...] Read more.
Background: in recent years, the management of advanced colorectal cancer (CRC) has been greatly improved with integrated strategies including stereotactic radiation therapy (SRT). The administration of SRT has been demonstrated, particularly in oligo-metastatic (om) CRC, to be a safe and effective option. Interestingly, it has been demonstrated that SRT can induce regression of tumors in non-irradiated regions (“abscopal effect”) through stimulation of anti-tumor immune effects (“radiation-induced immunity”). We have recently shown that lung-limited omCRC is characterized by regression of tumor clones bearing specific key driver gene mutations. Aims: to assess the genetic evolution on tumor cancer cells induced by SRT in lung-limited omCRC. Secondary objectives included descriptions of the abscopal effect, responses’ duration, toxicity, and progression-free survival. A translational research will be performed to evaluate tumor genetic evolution (through liquid biopsies and Next Generation Sequencing), HLA class I repertoire, peripheral immune cells, and cytokine dynamics. Methods: PRELUDE-1 is a prospective translational study. SRT will be administered only to the largest nodule (with a maximum diameter ≤ 25 mm) in omCRC with two or three radiologically evident lesions. The sample size is based on the innovative hypothesis that radiation-induced immunity could induce regression of tumor clones bearing KRAS oncogene mutations. According to the binomial test, considering the frequency of KRAS mutations and assuming a probability of mutant KRAS→wild type KRAS of p0 = 0.0077, with α = 0.05 and 1-β = 0.60, the final sample size is 25 patients. Full article
(This article belongs to the Special Issue Clinical Perspective and Translational Oncology of Liquid Biopsy)
Show Figures

Figure 1

Figure 1
<p>Timeline of PRELUDE-1 study.</p>
Full article ">
18 pages, 6362 KiB  
Review
Challenges and Future Perspectives of Immunotherapy in Pancreatic Cancer
by Anna Maxi Wandmacher, Anne Letsch and Susanne Sebens
Cancers 2021, 13(16), 4235; https://doi.org/10.3390/cancers13164235 - 23 Aug 2021
Cited by 19 | Viewed by 4368
Abstract
To date, extensive efforts to harness immunotherapeutic strategies for the treatment of pancreatic ductal adenocarcinoma (PDAC) have yielded disappointing results in clinical trials. These strategies mainly focused on cancer vaccines and immune checkpoint inhibitors alone or in combination with chemotherapeutic or targeted agents. [...] Read more.
To date, extensive efforts to harness immunotherapeutic strategies for the treatment of pancreatic ductal adenocarcinoma (PDAC) have yielded disappointing results in clinical trials. These strategies mainly focused on cancer vaccines and immune checkpoint inhibitors alone or in combination with chemotherapeutic or targeted agents. However, the growing preclinical and clinical data sets from these efforts have established valuable insights into the immunological characteristics of PDAC biology. Most notable are the immunosuppressive role of the tumour microenvironment (TME) and PDAC’s characteristically poor immunogenicity resulting from tumour intrinsic features. Moreover, PDAC tumour heterogeneity has been increasingly well characterized and may additionally limit a “one-fits-all” immunotherapeutic strategy. In this review, we first outline mechanisms of immunosuppression and immune evasion in PDAC. Secondly, we summarize recently published data on preclinical and clinical efforts to establish immunotherapeutic strategies for the treatment of PDAC including diverse combinatorial treatment approaches aiming at overcoming this resistance towards immunotherapeutic strategies. Particularly, these combinatorial treatment approaches seek to concomitantly increase PDAC antigenicity, boost PDAC directed T-cell responses, and impair the immunosuppressive character of the TME in order to allow immunotherapeutic agents to unleash their full potential. Eventually, the thorough understanding of the currently available data on immunotherapeutic treatment strategies of PDAC will enable researchers and clinicians to develop improved treatment regimens and to design innovative clinical trials to overcome the pronounced immunosuppression of PDAC. Full article
Show Figures

Figure 1

Figure 1
<p>Therapeutic strategies to target immunosuppressive components of the TME in PDAC. Abbreviations: Extracellular matrix (ECM); cancer associated fibroblasts (CAF); regulatory T-cells (Treg); tumour associated macrophages (TAM); dendritic cells (DC). Figure created with BioRender.com.</p>
Full article ">
18 pages, 27519 KiB  
Article
Quantitative Tyrosine Phosphoproteome Profiling of AXL Receptor Tyrosine Kinase Signaling Network
by Xinyan Wu, Li Wang, Nicole A. Pearson, Santosh Renuse, Ran Cheng, Ye Liang, Dong-Gi Mun, Anil K. Madugundu, Yaoyu Xu, Parkash S. Gill and Akhilesh Pandey
Cancers 2021, 13(16), 4234; https://doi.org/10.3390/cancers13164234 - 23 Aug 2021
Cited by 5 | Viewed by 4405
Abstract
Overexpression and amplification of AXL receptor tyrosine kinase (RTK) has been found in several hematologic and solid malignancies. Activation of AXL can enhance tumor-promoting processes such as cancer cell proliferation, migration, invasion and survival. Despite the important role of AXL in cancer development, [...] Read more.
Overexpression and amplification of AXL receptor tyrosine kinase (RTK) has been found in several hematologic and solid malignancies. Activation of AXL can enhance tumor-promoting processes such as cancer cell proliferation, migration, invasion and survival. Despite the important role of AXL in cancer development, a deep and quantitative mapping of its temporal dynamic signaling transduction has not yet been reported. Here, we used a TMT labeling-based quantitative proteomics approach to characterize the temporal dynamics of the phosphotyrosine proteome induced by AXL activation. We identified >1100 phosphotyrosine sites and observed a widespread upregulation of tyrosine phosphorylation induced by GAS6 stimulation. We also detected several tyrosine sites whose phosphorylation levels were reduced upon AXL activation. Gene set enrichment-based pathway analysis indicated the activation of several cancer-promoting and cell migration/invasion-related signaling pathways, including RAS, EGFR, focal adhesion, VEGFR and cytoskeletal rearrangement pathways. We also observed a rapid induction of phosphorylation of protein tyrosine phosphatases, including PTPN11 and PTPRA, upon GAS6 stimulation. The novel molecules downstream of AXL identified in this study along with the detailed global quantitative map elucidating the temporal dynamics of AXL activation should not only help understand the oncogenic role of AXL, but also aid in developing therapeutic options to effectively target AXL. Full article
(This article belongs to the Special Issue Proteomics in Cancer)
Show Figures

Figure 1

Figure 1
<p>AXL expression and GAS6 stimulation in breast cancer cell lines. (<b>A</b>). Scatter plot showing AXL expression levels based on RNA-Seq data and its protein expression based on proteomics data (both from CCLA) as indicated. The color of the dots represents the subtype of breast cancer cell lines: TNBC (red), luminal A (green), luminal B (blue) and HER2+ (yellow). (<b>B</b>). Western blot results showing AXL expression level in a panel of nine TNBC cell lines. Beta-actin serves as a loading control. (<b>C</b>). MDA-MB-231 cells were treated with the indicated concentration (ng/mL) of GAS6 for 10 min. Phosphorylation levels of AXL and AKT and their total protein expression levels were examined by Western blot analysis. (<b>D</b>). MDA-MB-231 cells were treated with 400 ng/mL GAS6 for different time points as indicated. Phosphorylation levels of AXL and AKT and their total protein expression levels were examined by Western blot analysis as shown. (The original western blot images were included in <a href="#app1-cancers-13-04234" class="html-app">Figure S1</a>).</p>
Full article ">Figure 2
<p>A schematic depicting the strategy used for temporal quantitative phosphotyrosine proteomic profiling of GAS6-treated MDA-MB-231 cells. All experiments were performed in triplicate. Cells were serum starved for 18 h and treated with 400 ng/mL GAS6 for 5 min, 10 min, 1 h or 5 h. After the GAS6 treatment, cells were harvested and lysed in 8 M urea buffer, followed by trypsin digestion and desalting. Tyrosine phosphorylated peptides were enriched with anti-pTyr antibody (pY1000). Enriched peptides were labeled with the TMTpro kit and analyzed by Orbitrap Eclipse.</p>
Full article ">Figure 3
<p>Dynamic regulation of tyrosine phosphorylation induced by the activation of GAS6/AXL signaling cascades. (<b>A</b>). Plot showing the number of upregulated and downregulated tyrosine phosphorylation sites at the indicated time points. (<b>B</b>). Venn diagram showing the overlap of regulated tyrosine phosphorylation sites at different times. (<b>C</b>). Fuzzy C-means clustering showing clusters of phosphotyrosine sites with dynamic regulation patterns. (<b>D</b>). A heatmap visualizing the hierarchical clustering of phosphotyrosine sites regulated by AXL. Each column represents the intensity at the indicated time points after GAS6 treatment.</p>
Full article ">Figure 4
<p>Signaling pathways activated by AXL. (<b>A</b>,<b>B</b>). Enriched signaling pathways for proteins with upregulated (<b>A</b>) or downregulated (<b>B</b>) tyrosine phosphorylation. Pathway analysis was performed using DAVID [<a href="#B29-cancers-13-04234" class="html-bibr">29</a>] based on the KEGG [<a href="#B42-cancers-13-04234" class="html-bibr">42</a>] pathway database and replotted using an R package. The size of the circle represents the number of proteins identified in each pathway. The color scheme represents −log<sub>10</sub>(<span class="html-italic">p</span>-value) as indicated. (<b>C</b>). Bubble plot displaying the alterations in tyrosine phosphorylation of protein tyrosine phosphatases (PTP), serine/threonine kinases, non-receptor tyrosine kinases (NRTK) and receptor tyrosine kinases (RTK). The color scheme represents the fold-changes upon GAS6 treatment of individual tyrosine phosphosites at the indicated time points. The size of the circle represents −log<sub>10</sub>(<span class="html-italic">p</span>-value) compared to untreated MDA-MB-231 cells.</p>
Full article ">Figure 5
<p>Site-specific regulation of tyrosine kinases and protein tyrosine phosphatases by AXL activation. (<b>A</b>). Schematic representation of receptor tyrosine kinases (top left panel), non-receptor tyrosine kinases (top right panel) and protein tyrosine phosphatases (middle panel) along with phosphorylation sites regulated by AXL activation and the time course. Black bars with red filled circles at the end denote phosphorylation modification. The embedded column plots depict the regulation pattern of individual phosphotyrosine sites. The domain structures of the molecules are shown with the key in the panel at the bottom. (<b>B</b>). Multiple sequences alignment of the amino acid sequences of the tyrosine kinase domain of the RTKs and NTRKs regulated by Axl. The box indicates the activation loop of the kinase domain. Regulated phosphotyrosine residues identified in this study were highlighted in pink.</p>
Full article ">Figure 6
<p>Western blot-based validation of selected molecules in the Axl signaling pathway. (<b>A</b>). MDA-MB-231 cells were treated with 400 ng/mL GAS6 for the indicated duration. Cell lysates were analyzed with Western blot analysis using phospho-specific antibodies against EPHA2 (pY772), AXL (pY703/pY702), PZR (pY263), SHP2/PTPN11 (pY580), PTPRA (pY798) and CDK1 (pY15) and antibodies against the corresponding protein as indicated. β-actin served as a loading control. The arrows indicate the three isoforms of PZR protein. (<b>B</b>). Western blot analysis to confirm the dynamic regulation of phosphorylation of the indicated proteins in HCC1395 cells treated with 400 ng/mL GAS6 for different time courses (as indicated). (<b>C</b>). MDA-MB-231 cells were pretreated with 2 µM R428 prior to GAS6 stimulation for 5 or 10 min, respectively. Protein phosphorylation levels were detected by indicated phospho-specific antibodies. (The original western blot images were included in <a href="#app1-cancers-13-04234" class="html-app">Figure S1</a>).</p>
Full article ">
27 pages, 6608 KiB  
Review
Tumor Microenvironment in Breast Cancer—Updates on Therapeutic Implications and Pathologic Assessment
by Joshua J. Li, Julia Y. Tsang and Gary M. Tse
Cancers 2021, 13(16), 4233; https://doi.org/10.3390/cancers13164233 - 23 Aug 2021
Cited by 96 | Viewed by 7464
Abstract
The tumor microenvironment (TME) in breast cancer comprises local factors, cancer cells, immune cells and stromal cells of the local and distant tissues. The interaction between cancer cells and their microenvironment plays important roles in tumor proliferation, propagation and response to therapies. There [...] Read more.
The tumor microenvironment (TME) in breast cancer comprises local factors, cancer cells, immune cells and stromal cells of the local and distant tissues. The interaction between cancer cells and their microenvironment plays important roles in tumor proliferation, propagation and response to therapies. There is increasing research in exploring and manipulating the non-cancerous components of the TME for breast cancer treatment. As the TME is now increasingly recognized as a treatment target, its pathologic assessment has become a critical component of breast cancer management. The latest WHO classification of tumors of the breast listed stromal response pattern/fibrotic focus as a prognostic factor and includes recommendations on the assessment of tumor infiltrating lymphocytes and PD-1/PD-L1 expression, with therapeutic implications. This review dissects the TME of breast cancer, describes pathologic assessment relevant for prognostication and treatment decision, and details therapeutic options that interacts with and/or exploits the TME in breast cancer. Full article
(This article belongs to the Special Issue Radiotherapy and Chemotherapy for Cancers)
Show Figures

Figure 1

Figure 1
<p>Effects of (<b>a</b>) Radiotherapy and (<b>b</b>) Chemotherapy on the Breast Cancer Tumor Microenvironment.</p>
Full article ">Figure 1 Cont.
<p>Effects of (<b>a</b>) Radiotherapy and (<b>b</b>) Chemotherapy on the Breast Cancer Tumor Microenvironment.</p>
Full article ">Figure 2
<p>Tumor infiltrating lymphocytes on H&amp;E sections. (<b>a</b>) Breast cancer with dense stromal infiltrating lymphocytes, H&amp;E, 200× magnification; (<b>b</b>) Breast cancer with sparse lymphocytic infiltrates, H&amp;E, 200×.</p>
Full article ">Figure 3
<p>Fibrotic Foci on H&amp;E Sections. (<b>a</b>) H&amp;E, 20×; (<b>b</b>) H&amp;E, 100×.</p>
Full article ">Figure 4
<p>Immunohistochemistry for immune cells in the tumor microenvironment. (<b>a</b>) CD8 stain highlights cytotoxic T-cells among lymphocytes, CD8, 200×; (<b>b</b>) CD68 highlights tumor associated macrophages which are difficult to identify by H&amp;E, CD68, 200×.</p>
Full article ">Figure 5
<p>PD-L1 Immunohistochemistry. (<b>a</b>) Example of negative PD-L1 expression in tumor and immune cells, Ventana SP142, 200×; (<b>b</b>) Positive PD-L1 staining in tumor and immune cells, Ventana SP142, 200×; (<b>c</b>) Positive staining in immune cells only, Ventana SP142, 400×.</p>
Full article ">Figure 6
<p>CD31 stain for assessment of microvessel density, CD31, 200×.</p>
Full article ">
12 pages, 1095 KiB  
Article
Exploration of Novel Prognostic Markers in Grade 3 Neuroendocrine Neoplasia
by Rebecca Abdelmalak, Mark P. Lythgoe, Joanne Evans, Michael Flynn, Justin Waters, Andy Webb, David J. Pinato and Rohini Sharma
Cancers 2021, 13(16), 4232; https://doi.org/10.3390/cancers13164232 - 23 Aug 2021
Cited by 8 | Viewed by 2029
Abstract
Background: High-grade neuroendocrine tumours and carcinomas (NET/NECs) behave aggressively, typically presenting at an advanced stage. Prognosis is poor, with median survival between 5 and 34 months. The mainstay of treatment is palliative systemic therapy. However, therapy carries a risk of toxicity, which can [...] Read more.
Background: High-grade neuroendocrine tumours and carcinomas (NET/NECs) behave aggressively, typically presenting at an advanced stage. Prognosis is poor, with median survival between 5 and 34 months. The mainstay of treatment is palliative systemic therapy. However, therapy carries a risk of toxicity, which can reduce quality of life. Therefore, accurate prognostic scores for risk stratification of patients with high-grade NET/NECs are needed to help guide patient management to decide whether active treatment is likely to improve overall survival (OS). We aimed to compare the prognostic ability of published prognostic scores to predict OS in a cohort of patients with high-grade NET/NECs of any primary site. Methods: Treatment, biochemical and clinicopathological data were collected retrospectively from 77 patients with high-grade NET/NECs across three hospitals between 2016 and 2020. Variables including performance status (PS), Ki-67, age at diagnosis, previous treatment and presence of liver metastases were recorded. Pre-treatment neutrophil-lymphocyte ratio (NLR), platelet-lymphocyte ratio, modified Glasgow prognostic score (mGPS), and gastrointestinal neuroendocrine carcinoma (GI-NEC) score were derived. Univariable and multivariable survival analyses were used to assess prognostic ability. Results: The median age of the cohort was 63 years (range: 31–85); 53% of subjects were female. Grade 3 NETs (G3-NETs) were identified in 32 patients and NECs in 45 patients. The median OS was 13.45 months (range: 0.87–65.37) with no difference observed between G3-NETs and NECs. Univariable analysis revealed that NLR (n = 72, p = 0.049), mGPS (n = 56, p = 0.003), GI-NEC score (n = 27, p = 0.0007) and Ki-67 (n = 66, p = 0.007) were significantly associated with OS. Multivariable analysis confirmed that elevated mGPS (p = 0.046), GI-NEC score (p = 0.036), and Ki-67 (p = 0.02) were independently prognostic for reduced OS across the entire cohort. mGPS was identified as an independent prognostic factor in G3-NETs. Independent predictors of OS in NECs were PS and Ki-67. Conclusions: mGPS, PS and Ki-67 are independent prognostic markers in high-grade NET/NEC patients. Our study supports the use of these prognostic scores for risk stratification of patients with high grade cancers and as useful tools to guide treatment decisions. Full article
(This article belongs to the Special Issue Advanced Neuroendocrine Tumors)
Show Figures

Figure 1

Figure 1
<p>Kaplan–Meier survival curves of OS using different prognostic scores in the entire patient cohort. G3-NETs and NECs were divided into groups defined by different prognostic scores: modified Glasgow prognostic score (mGPS) (<b>A</b>); neutrophil-lymphocyte ratio (NLR) (<b>B</b>); gastrointestinal neuroendocrine carcinoma (GI-NEC) score (<b>C</b>) and Ki-67 using 55% as the cut-off (<b>D</b>). Median OS, range and log-rank test <span class="html-italic">p</span> value are reported; <span class="html-italic">p</span> &lt; 0.05 was regarded as significant. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
16 pages, 3924 KiB  
Article
MET Mutation Is a Potential Therapeutic Target for Advanced Endometrial Cancer
by Yu-Min Yeh, Pei-Ying Wu, Peng-Chan Lin, Pei-Fang Su, Ya-Ting Hsu, Keng-Fu Hsu and Meng-Ru Shen
Cancers 2021, 13(16), 4231; https://doi.org/10.3390/cancers13164231 - 23 Aug 2021
Cited by 4 | Viewed by 3022
Abstract
An optimal therapeutic regimen for endometrial cancer with extra-uterine metastasis is unavailable. This study aims to improve our understanding of the genomic landscape of advanced endometrial cancer and identify potential therapeutic targets. The clinical and genomic profiles of 81 patients with stage III [...] Read more.
An optimal therapeutic regimen for endometrial cancer with extra-uterine metastasis is unavailable. This study aims to improve our understanding of the genomic landscape of advanced endometrial cancer and identify potential therapeutic targets. The clinical and genomic profiles of 81 patients with stage III or IV endometrial cancer were integrated. To identify genomic aberrations associated with clinical outcomes, Cox proportional hazard regression was used. The impacts of the genomic aberrations were validated in vitro and in vivo. The mutation status of MET, U2AF1, BCL9, PPP2R1A, IDH2, CBL, BTK, and CHEK2 were positively correlated with poor clinical outcomes. MET mutations occurred in 30% of the patients who presented with poor overall survival (hazard ratio, 2.606; 95% confidence interval, 1.167~5.819; adjusted p-value, 0.067). Concurrent MET and KRAS mutations presented with the worst outcomes. MET mutations in hepatocyte growth factor (HGF)-binding (58.1%) or kinase (16.2%) domains resulted in differential HGF-induced c-MET phosphorylation. Different types of MET mutations differentially affected tumor growth and displayed different sensitivities to cisplatin and tyrosine kinase inhibitors. MET N375S mutation is a germline variant that causes chemoresistance to cisplatin, with a high incidence in Eastern Asia. This study highlights the ethnic differences in the biology of the disease, which can influence treatment recommendations and the genome-guided clinical trials of advanced endometrial cancer. Full article
Show Figures

Figure 1

Figure 1
<p>Impact of <span class="html-italic">MET</span> mutation and concurrent genetic mutations on the clinical outcome of advanced endometrial cancer. (<b>A</b>) Kaplan–Meier curves for overall survival (OS) in patients with and without <span class="html-italic">MET</span> mutation compared using the log-rank test. (<b>B</b>–<b>D</b>) Kaplan–Meier curves for OS were analyzed between the <span class="html-italic">MET</span> wild-type and <span class="html-italic">MET</span> mutant patients with or without concurrent <span class="html-italic">PIK3R1</span> (<b>B</b>), <span class="html-italic">PIK3CA</span> mutation (<b>C</b>), and <span class="html-italic">KRAS</span> (<b>D</b>).</p>
Full article ">Figure 2
<p>Differential effects of various <span class="html-italic">MET</span> mutations on c-MET phosphorylation and cellular functions. (<b>A</b>) The structure of c-MET (left) and distribution of <span class="html-italic">MET</span> mutations (right). Sema, semaphoring domain; PSI, plexin–semaphorin–integrin domain; IPT, immunoglobulin–plexin–transcription domain. (<b>B</b>) Representative images and grading of c-MET expression in endometrial cancer. The expression of c-MET was scored as grade 1–3, according to the percentage of the positively-stained tumor cells and the intensity of the staining. (<b>C</b>) Time course of c-MET phosphorylation in response to HGF (50 ng/mL) stimulation. Cell lysates from various clones of the endometrial cancer RL95-2 cells were immunoblotted with anti-c-MET, anti-phospho-c-MET, and anti-α-tubulin antibodies (upper). The expression levels were quantified, and the ratios of phosphor-c-MET to total c-MET are shown (bottom). (<b>D</b>) Cell proliferation, invasion, and migration. (<b>E</b>) Assays using various clones of the RL95-2 cells. Each value represents the mean ± standard error of mean (SEM) from at least three experiments. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Antitumor effects of cisplatin and tyrosine kinase inhibitors in endometrial cancer xenografts harboring different <span class="html-italic">MET</span> mutations. 1 × 10<sup>6</sup> RL95-2 cells carrying wild-type or mutant <span class="html-italic">MET</span>, including N375S, G1085R, and G1087E, were injected into the posterior flank of SCID mice subcutaneously. The tumors were allowed to grow and were treated by observation only (<b>A</b>,<b>B</b>), cisplatin (1 mg/kg) twice a week (<b>C</b>,<b>D</b>), crizotinib (25 mg/kg) by oral gavage twice a week (<b>E</b>,<b>F</b>), or SU11274 (6 mg/kg) twice a week (<b>G</b>,<b>H</b>) (<span class="html-italic">n</span> = 5 in each treatment group). Tumor size and body weight were measured twice a week. Data are presented as the mean tumor size ± SEM. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; NS, non-significance.</p>
Full article ">Figure 4
<p>The clinical impact, distribution, and frequency of the germline <span class="html-italic">MET</span> N375S variant. (<b>A</b>) OS of advanced endometrial cancer patients with and without the <span class="html-italic">MET</span> N375S variant. (<b>B</b>) The distribution and frequency of <span class="html-italic">MET</span> variants in 35 paired normal and tumor tissues. Whole-genome sequencing was used to analyze DNA extracted from blood samples or adjacent non-tumor specimens to determine germline <span class="html-italic">MET</span> variants. (<b>C</b>) The distribution and frequency of germline <span class="html-italic">MET</span> variants in cancer patients and the normal Taiwan population. The database from the Taiwan Biobank, which contained germline whole-genome sequencing data of 499 normal Taiwanese individuals, was used to compare the distribution and frequency of germline <span class="html-italic">MET</span> variants in the NCKUH cohort and the normal Taiwan population. (<b>D</b>) The alternative allele frequency of <span class="html-italic">MET</span> c.1124 (N375S) in the different ethnic groups. The frequency of the variant allele in <span class="html-italic">MET</span> c.1124 was compared among the NCKUH cohort, the normal Taiwanese population, and different ethnic groups enrolled in the 1000 Genomes Project.</p>
Full article ">
10 pages, 420 KiB  
Article
Concordance of Child Self-Reported and Parent Proxy-Reported Posttraumatic Growth in Childhood Cancer Survivors
by Veronika Koutná, Marek Blatný and Martin Jelínek
Cancers 2021, 13(16), 4230; https://doi.org/10.3390/cancers13164230 - 23 Aug 2021
Cited by 5 | Viewed by 2864
Abstract
This article aimed to analyze concordance of parent- and child-reported child posttraumatic growth (PTG) following pediatric cancer, the influence of the parents’ own level of PTG on the level of concordance and the influence of the parents’ and the child’s own level of [...] Read more.
This article aimed to analyze concordance of parent- and child-reported child posttraumatic growth (PTG) following pediatric cancer, the influence of the parents’ own level of PTG on the level of concordance and the influence of the parents’ and the child’s own level of PTG on the parents’ proxy reports of PTG in the child. The sample included 127 parent–child dyads. The children provided self-reports of PTG and the parents provided reports of their own as well as the child’s PTG. Overall, the results showed poor parent–child agreement on the child PTG, with the parents proxy-reporting higher levels of PTG than the children. The parents’ proxy reports of the child PTG were the most accurate at the lowest levels of the parents’ own level of PTG. The parents’ own level of PTG was a stronger predictor of the parents’ proxy reports than the child self-reported PTG. The results suggest that parents are not very accurate reporters of PTG in the child; therefore, their reports should be completed with child self-reports whenever possible. Full article
Show Figures

Figure 1

Figure 1
<p>Level of child PTG according to self-reports and the parents’ proxy reports and the level of PTG in parents. * <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">
27 pages, 2898 KiB  
Review
Recent Advances in the Treatment of Bone Metastases and Primary Bone Tumors: An Up-to-Date Review
by Adrian Emilian Bădilă, Dragoș Mihai Rădulescu, Adelina-Gabriela Niculescu, Alexandru Mihai Grumezescu, Marius Rădulescu and Adrian Radu Rădulescu
Cancers 2021, 13(16), 4229; https://doi.org/10.3390/cancers13164229 - 23 Aug 2021
Cited by 44 | Viewed by 13577
Abstract
In the last decades, the treatment of primary and secondary bone tumors has faced a slow-down in its development, being mainly based on chemotherapy, radiotherapy, and surgical interventions. However, these conventional therapeutic strategies present a series of disadvantages (e.g., multidrug resistance, tumor recurrence, [...] Read more.
In the last decades, the treatment of primary and secondary bone tumors has faced a slow-down in its development, being mainly based on chemotherapy, radiotherapy, and surgical interventions. However, these conventional therapeutic strategies present a series of disadvantages (e.g., multidrug resistance, tumor recurrence, severe side effects, formation of large bone defects), which limit their application and efficacy. In recent years, these procedures were combined with several adjuvant therapies, with different degrees of success. To overcome the drawbacks of current therapies and improve treatment outcomes, other strategies started being investigated, like carrier-mediated drug delivery, bone substitutes for repairing bone defects, and multifunctional scaffolds with bone tissue regeneration and antitumor properties. Thus, this paper aims to present the types of bone tumors and their current treatment approaches, further focusing on the recent advances in new therapeutic alternatives. Full article
(This article belongs to the Special Issue Diagnosis and Treatment for Bone Tumor and Sarcoma)
Show Figures

Figure 1

Figure 1
<p>Anatomical distribution of common bone sarcoma. Reprinted from an open-access source [<a href="#B8-cancers-13-04229" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Bone sarcoma stem cells overview. (<b>a</b>) osteosarcoma; (<b>b</b>) Ewing sarcoma; (<b>c</b>) chondrosarcoma. (<b>d</b>) A broad spectrum of CSC markers (black) and the molecular mechanisms underlying CSC phenotypes (blue) have been documented for each sarcoma. Several anti-CSC compounds (red) have been preclinically tried to inhibit CSC phenotypes. Reprinted from an open-access source [<a href="#B20-cancers-13-04229" class="html-bibr">20</a>].</p>
Full article ">Figure 3
<p>Graphical representation of the “vicious cycle” caused by cancer bone metastasis. Reprinted from an open-access source [<a href="#B36-cancers-13-04229" class="html-bibr">36</a>].</p>
Full article ">Figure 4
<p>Comparison of bone tumor microenvironment in metastasis versus sarcoma. Created based on information from literature references [<a href="#B36-cancers-13-04229" class="html-bibr">36</a>,<a href="#B42-cancers-13-04229" class="html-bibr">42</a>,<a href="#B43-cancers-13-04229" class="html-bibr">43</a>,<a href="#B44-cancers-13-04229" class="html-bibr">44</a>,<a href="#B45-cancers-13-04229" class="html-bibr">45</a>].</p>
Full article ">Figure 5
<p>Various nanocarriers for anticancer drugs. Reprinted from an open-access source [<a href="#B94-cancers-13-04229" class="html-bibr">94</a>].</p>
Full article ">Figure 6
<p>Graphical representation of functionalized scaffolds. (<b>a</b>) Physically immobilized bioactive molecules; (<b>b</b>) Covalently bound bioactive molecules; (<b>c</b>) Scaffold coated with ECM molecules. Reprinted from an open-access source [<a href="#B16-cancers-13-04229" class="html-bibr">16</a>].</p>
Full article ">
10 pages, 1163 KiB  
Article
Prognostic Role of Pre-Treatment [18F]FDG PET/CT in Patients with Anaplastic Thyroid Cancer
by Hyun Jeong Kim, Hang-Seok Chang and Young Hoon Ryu
Cancers 2021, 13(16), 4228; https://doi.org/10.3390/cancers13164228 - 23 Aug 2021
Cited by 8 | Viewed by 3123
Abstract
Anaplastic thyroid carcinoma (ATC) is a rare but highly lethal disease. Therefore, its diagnosis at an early stage and a rapid and accurate establishment of a proper treatment strategy is warranted. Tumor glycolysis assessed by 18fluorodeoxyglucose ([18F]FDG) positron emission tomography [...] Read more.
Anaplastic thyroid carcinoma (ATC) is a rare but highly lethal disease. Therefore, its diagnosis at an early stage and a rapid and accurate establishment of a proper treatment strategy is warranted. Tumor glycolysis assessed by 18fluorodeoxyglucose ([18F]FDG) positron emission tomography (PET)/computed tomography (CT) is predictive of many cancers despite its limited proven applicability to ATC. We investigated the prognostic capability of [18F]FDG PET/CT in patients with ATC. Forty patients with ATC were subjected to [18F]FDG PET/CT for pre-treatment evaluation. The tumor size and stage, overall survival (OS), and PET parameters, including the maximum standardized uptake value (SUVmax), metabolic tumor volume (MTV), and total lesion glycolysis (TLG) were analyzed. The 1-year OS rate was 17.5% with a mean life expectancy of 7.1 months. Distant metastasis was detected solely using PET/CT in 37.5% of cases. High SUVmax, MTV, and TLG were significantly associated with poor prognosis (p < 0.001, p = 0.002, and p < 0.001, respectively). A significant difference (p < 0.001) was observed in OS between patients with a high and low tumor SUVmax. Glucose metabolism assessed by [18F]FDG PET/CT was significantly associated with the OS of patients with ATC. PET-derived parameters such as SUVmax, MTV, and TLG are useful prognostic biomarkers for ATC. Full article
(This article belongs to the Special Issue Biomarkers of Thyroid Cancer)
Show Figures

Figure 1

Figure 1
<p>Representative <sup>18</sup>fluorodeoxyglucose ([<sup>18</sup>F]FDG) positron emission tomography (PET)/computed tomography (CT) images of patients with anaplastic thyroid cancer and lung metastases. (<b>a</b>) The maximum intensity projection image of a 74-year-old female patient. (<b>b</b>) PET axial and (<b>c</b>) fusion axial images show intense FDG uptake in the thyroid tumor. The maximum standardized uptake value (SUVmax) was 35.98, and the patient expired 2.1 months after diagnosis. (<b>d</b>–<b>f</b>) The maximum intensity projection, PET axial, and fusion axial images of a 63-year-old female patient with a tumor SUVmax of 8.81. Her overall survival time was 8.1 months.</p>
Full article ">Figure 2
<p>Kaplan–Meier analyses of patients with anaplastic thyroid cancer. (<b>a</b>) Overall survival curves according to the maximum standardized uptake value (SUVmax) of tumor. (<b>b</b>) Overall survival curves according to total lesion glycolysis (TLG). (<b>c</b>) Overall survival curves according to metabolic tumor volume (MTV).</p>
Full article ">
17 pages, 1831 KiB  
Article
Casein Kinase 1D Encodes a Novel Drug Target in Hedgehog—GLI-Driven Cancers and Tumor-Initiating Cells Resistant to SMO Inhibition
by Elisabeth Peer, Sophie Karoline Aichberger, Filip Vilotic, Wolfgang Gruber, Thomas Parigger, Sandra Grund-Gröschke, Dominik Patrick Elmer, Florian Rathje, Andrea Ramspacher, Mirko Zaja, Susanne Michel, Svetlana Hamm and Fritz Aberger
Cancers 2021, 13(16), 4227; https://doi.org/10.3390/cancers13164227 - 23 Aug 2021
Cited by 8 | Viewed by 3775
Abstract
(1) Background: Aberrant activation of the hedgehog (HH)—GLI pathway in stem-like tumor-initiating cells (TIC) is a frequent oncogenic driver signal in various human malignancies. Remarkable efficacy of anti-HH therapeutics led to the approval of HH inhibitors targeting the key pathway effector smoothened (SMO) [...] Read more.
(1) Background: Aberrant activation of the hedgehog (HH)—GLI pathway in stem-like tumor-initiating cells (TIC) is a frequent oncogenic driver signal in various human malignancies. Remarkable efficacy of anti-HH therapeutics led to the approval of HH inhibitors targeting the key pathway effector smoothened (SMO) in basal cell carcinoma and acute myeloid leukemia. However, frequent development of drug resistance and severe adverse effects of SMO inhibitors pose major challenges that require alternative treatment strategies targeting HH—GLI in TIC downstream of SMO. We therefore investigated members of the casein kinase 1 (CSNK1) family as novel drug targets in HH—GLI-driven malignancies. (2) Methods: We genetically and pharmacologically inhibited CSNK1D in HH-dependent cancer cells displaying either sensitivity or resistance to SMO inhibitors. To address the role of CSNK1D in oncogenic HH signaling and tumor growth and initiation, we quantitatively analyzed HH target gene expression, performed genetic and chemical perturbations of CSNK1D activity, and monitored the oncogenic transformation of TIC in vitro and in vivo using 3D clonogenic tumor spheroid assays and xenograft models. (3) Results: We show that CSNK1D plays a critical role in controlling oncogenic GLI activity downstream of SMO. We provide evidence that inhibition of CSNK1D interferes with oncogenic HH signaling in both SMO inhibitor-sensitive and -resistant tumor settings. Furthermore, genetic and pharmacologic perturbation of CSNK1D decreases the clonogenic growth of GLI-dependent TIC in vitro and in vivo. (4) Conclusions: Pharmacologic targeting of CSNK1D represents a novel therapeutic approach for the treatment of both SMO inhibitor-sensitive and -resistant tumors. Full article
(This article belongs to the Special Issue Stemness and Differentiation in Cancer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>RNAi-mediated inhibition of CSNK1D interferes with canonical oncogenic HH—GLI signaling. (<b>A</b>) Simplified schematic illustration of the canonical HH—GLI signaling pathway with pathway activators in green and inhibitors in red. PTCH and SUFU inactivate the pathway, while it is turned on by the central pathway activator SMO causing activation of the GLI transcription factor. Pharmacologically, the SMO agonist SAG activates the pathway, while it is inhibited by the SMO inhibitor vismodegib. (<b>B</b>) Representative Western blot analysis of GLI1 and CSNK1D in Daoy medulloblastoma cells. Cells were lentivirally transduced with shCSNK1D or control shRNA (shCtrl) and treated with or without SAG [100 nM]. Relative quantification of Western blot bands was conducted via densitometric image analysis using Image Lab 5.0 software (Bio-Rad, Vienna, Austria). Relative protein levels normalized to the loading control tERK and to the shCtrl + SAG sample are shown above each protein band. (<b>C</b>) qPCR analysis of <span class="html-italic">GLI1</span> and <span class="html-italic">HHIP</span> mRNA levels of Daoy cells treated as described in (<b>B</b>) (<span class="html-italic">n</span> = 3). Student’s <span class="html-italic">t</span> test was used for statistical analysis (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>RNAi-mediated inhibition of CSNK1D interferes with oncogenic HH—GLI signaling in SMOi resistant Ewing sarcoma cells. (<b>A</b>) Simplified schematic illustration of non-canonical, SMO-independent HH—GLI signaling in Ewing sarcoma with EWS–FLI1-driven transcription of <span class="html-italic">GLI1</span>. (<b>B</b>) Representative Western blot analysis of GLI1 in A673 cells treated with vismodegib [1 μM] or HPI-1 [20 μM]. (<b>C</b>) Representative Western blot analysis of GLI1 in A673 cells lentivirally transduced with shCSNK1D (#1, #2) or control shRNA (shCtrl). Relative quantification of Western blot bands was conducted via densitometric image analysis using Image Lab 5.0 software (Bio-Rad, Vienna, Austria). Relative protein levels normalized to the loading control tERK and to the Ctrl sample are shown above each protein band.</p>
Full article ">Figure 3
<p>Pharmacological targeting of CSNK1D inhibits SMO-dependent GLI activation. (<b>A</b>) mRNA expression levels of <span class="html-italic">GLI1</span> were analyzed by qPCR and expressed as percentage relative to the control sample (<span class="html-italic">n</span> = 3). (<b>B</b>) Representative Western blot analysis of GLI1 protein levels in Daoy cells treated with SAG [100 nM] and increasing concentrations of CK1D008 [0.01–10 μM] or SR-3029 [0.01–1 μM]. Relative quantification of Western blot bands was conducted via densitometric image analysis using Image Lab 5.0 software (Bio-Rad, Vienna, Austria). Relative protein levels normalized to the loading control tERK and to the SAG-treated sample are shown above each protein band.</p>
Full article ">Figure 4
<p>Pharmacological targeting of CSNK1D inhibits oncogenic HH—GLI signaling in SMOi-resistant cells. (<b>A</b>) Simplified schematic illustration of SMO-independent activation of HH—GLI signaling in Daoy medulloblastoma cells caused by the loss of the negative HH—GLI pathway regulator SUFU. (<b>B</b>) Representative Western blot analysis of GLI1 protein levels in Daoy-shSUFU cells treated with vismo [1 μM], HPI-1 [20 μM], CK1D008 [0.1–3 μM] or SR-3029 [0.01–1 μM]. (<b>C</b>) Representative Western blot analysis of GLI1 in A673 cells treated with vismo [1 μM], HPI-1 [20 μM], CK1D008 [0.1–1 μM] or SR-3029 [0.03–0.3 μM]. Relative quantification of Western blot bands was conducted via densitometric image analysis using Image Lab 5.0 software (Bio-Rad, Vienna, Austria). Relative protein levels normalized to the loading control tERK and to the Ctrl sample are shown above each protein band.</p>
Full article ">Figure 5
<p>Targeting CSNK1D reduces GLI-dependent 3D tumor spheroid formation in vitro as well as tumor engraftment in vivo. (<b>A</b>) A673 cells were lentivirally transduced with shCSNK1D or control shRNA (shCtrl) and cultured under anchorage-independent conditions. Representative images of formed 3D spheroid colonies (left panel). The number of 3D spheroid colonies was counted (right panel) (<span class="html-italic">n</span> = 8). (<b>B</b>) A673 cells were cultured under anchorage-independent conditions and treated with the GLI inhibitor HPI-1 [4 μM], CK1D008 [0.1–1 μM] or SR-3029 [0.03–0.3 μM]. The data represent the mean of six 3D culture experiments. (<b>C</b>) A673 cells were retrovirally transduced with <span class="html-italic">GLI1</span>, active <span class="html-italic">GLI2</span> (GLI2act) or control overexpression constructs, cultured under anchorage-independent conditions and treated with CK1D008 at non-toxic concentration [0.3 μM]. The number of 3D spheroid colonies was counted (<span class="html-italic">n</span> = 4 experiments). (<b>D</b>) A673 cells were lentivirally transduced with shCSNK1D or control shRNA (shCtrl) and engrafted in the flanks of NSG mice (<span class="html-italic">n</span> = 6 mice each). Tumor volume was measured every 3–4 days. Student’s <span class="html-italic">t</span> test was used for statistical analysis (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
31 pages, 7838 KiB  
Review
Role of Inflammatory Mediators, Macrophages, and Neutrophils in Glioma Maintenance and Progression: Mechanistic Understanding and Potential Therapeutic Applications
by Abdul Samad Basheer, Faridah Abas, Iekhsan Othman and Rakesh Naidu
Cancers 2021, 13(16), 4226; https://doi.org/10.3390/cancers13164226 - 23 Aug 2021
Cited by 61 | Viewed by 9414
Abstract
Gliomas are the most common, highly malignant, and deadliest forms of brain tumors. These intra-cranial solid tumors are comprised of both cancerous and non-cancerous cells, which contribute to tumor development, progression, and resistance to the therapeutic regimen. A variety of soluble inflammatory mediators [...] Read more.
Gliomas are the most common, highly malignant, and deadliest forms of brain tumors. These intra-cranial solid tumors are comprised of both cancerous and non-cancerous cells, which contribute to tumor development, progression, and resistance to the therapeutic regimen. A variety of soluble inflammatory mediators (e.g., cytokines, chemokines, and chemotactic factors) are secreted by these cells, which help in creating an inflammatory microenvironment and contribute to the various stages of cancer development, maintenance, and progression. The major tumor infiltrating immune cells of the tumor microenvironment include TAMs and TANs, which are either recruited peripherally or present as brain-resident macrophages (microglia) and support stroma for cancer cell expansion and invasion. These cells are highly plastic in nature and can be polarized into different phenotypes depending upon different types of stimuli. During neuroinflammation, glioma cells interact with TAMs and TANs, facilitating tumor cell proliferation, survival, and migration. Targeting inflammatory mediators along with the reprogramming of TAMs and TANs could be of great importance in glioma treatment and may delay disease progression. In addition, an inhibition of the key signaling pathways such as NF-κB, JAK/STAT, MAPK, PI3K/Akt/mTOR, and TLRs, which are activated during neuroinflammation and have an oncogenic role in glioblastoma (GBM), can exert more pronounced anti-glioma effects. Full article
Show Figures

Figure 1

Figure 1
<p>Represents tumorigenic role of inflammatory mediators, TAMs, TANs, and signaling pathways in glioma. Green color indicates activation/release. TAMs: tumor-associated macrophages; TANs: tumor-associated neutrophils; MCP-1,3: monocyte chemoattractant proteins-1,3; CSF-1,2: colony stimulating factors-1,2; GDNF: glial cell-derived neurotrophic factor; IL-1 α, β, -4, -6, -8, -10: interleukins-1 α, β, -4, -6, -8, -10; CXCL-1, -12; chemokine (C-X-C motif) ligand-1, -12; TNF- α: tumor necrosis factor- α; COX-2: cyclooxygenase-2; PGE2: prostaglandin E2; PDGF: platelet-derived growth factor; TGF β: transforming growth factor β; VEGF: vascular endothelial growth factor; MMP-2, 9: matrix metalloproteinases-2, 9; IGFBP1: insulin-like growth factor-binding protein 1; Bmi-1: B-cell specific Moloney murine virus integration site 1; Arg-1: arginase-1; NF-κB: Nuclear factor kappa-light-chain-enhancer of activated B cells; JAK/STAT: Janus kinase-signal transducer and activator of transcription; MAPK: Mitogen-activated protein kinase; PI3K/Akt/mTOR: Phosphatidylinositol-3-kinase-mammalian target of rapamycin; TLRs: Toll-like receptors.</p>
Full article ">Figure 2
<p>Representation of the key signaling pathways involved in the inflammation-induced glioma genesis. (<b>A</b>) NF-κB pathway; (<b>B</b>) JAK/STAT pathway; (<b>C</b>) MAP Kinase pathway; (<b>D</b>) PI3K/Akt/mTOR pathway; (<b>E</b>) TLRs pathway. NF-κB: Nuclear factor kappa-light-chain-enhancer of activated B cells; JAK/STAT: Janus kinase-signal transducer and activator of transcription; MAPK: Mitogen-activated protein kinase; PI3K/Akt/mTOR: Phosphatidylinositol-3-kinase-mammalian target of rapamycin; TLRs: Toll-like receptors; IL-1β, -4, -6, -13, -22: interleukins-1β, -4, -6, -13, -22; CXCL-12: chemokine (C-X-C motif) ligand-12; TNF- α: tumor necrosis factor- α; COX-2: cyclooxygenase-2; PDGF: platelet-derived growth factor; TGF β: transforming growth factor β; HGF: hepatocyte growth factor; OS: oxidative stress; TROY: tumor necrosis factor receptor superfamily member 19; pSTAT3: phosphorylated signal transducer and activator of transcription 3; GDNF: glial cell-derived neurotrophic factor; Iκβα: inhibitor of kappa B alpha; EGFR: epidermal growth factor receptor; ERK: extracellular signal-regulated kinase; JNK: c-Jun N-terminal kinase; VEGF: vascular endothelial growth factor; MMPs: matrix metalloproteinases; Bax/Bcl-2: Bcl-2-associated X <span class="html-italic">protein</span>, B-cell lymphoma 2; AP-1: activator protein-1; IGFs: insulin-like growth factors; LPS: lipopolysaccharide; PI3K: phosphoinositide 3 kinase; Akt: protein kinase B; PDK: 3-phosphoinositide-dependent kinase 1; PIP2/3: phosphatidylinositol-4, 5-bisphosphate/ phosphatidylinositol-3, 4, 5-triphosphate; mTORC1/mTORC2: mechanistic target of rapamycin complex 1 and 2; TSC1/2: tuberous sclerosis complex 1/2; PTEN: phosphatase and tensin homolog: RTK: receptor tyrosine kinase; Rheb: Ras homolog enriched in brain; CD14: cluster of differentiation 14; MD2: myeloid differentiation factor 2; MyD88: myeloid differentiation factor 88; TIRAP: TIR-containing adaptor protein; TRIF: Toll/IL-1R domain-containing adapter inducing interferon-<span class="html-italic">β</span>; TRAF6, -3: TNF receptor-associated factor 6, -3; TRAM: TRIF-related adaptor molecule; RIP1: receptor-interacting protein 1; TAB-1, 2/3: TAK1-binding protein-1, 2/3; TAK1: TGF-β-activated kinase; TRADD: TNF receptor type 1-associated DEATH domain protein; IGF-1: insulin growth factor-1; CREB: cAMP response element binding protein; IRF3: interferon regulatory factor 3.</p>
Full article ">Figure 3
<p>Represents the therapeutic benefits of targeting inflammatory mediators, TAMs, TANs, and signaling pathways in glioma. Red color indicates inhibition/blocking and green color shows activation/upregulation. NF-κB: Nuclear factor kappa-light-chain-enhancer of activated B cells; JAK/STAT: Janus kinase-signal transducer and activator of transcription; MAPK: Mitogen-activated protein kinase; PI3K/Akt/mTOR: Phosphatidylinositol-3-kinase-mammalian target of rapamycin; TLRs: Toll-like receptors; TAMs: tumor-associated macrophages; TANs: tumor-associated neutrophils; IL-1β, -6, -8, -10: interleukins-1β, -6, -8, -10; IL-6R: interleukin-6 receptor; TNF- α: tumor necrosis factor- α; MMP-2/9/14: matrix metalloproteinase-2/9/14; PDGFR-α: platelet-derived growth factor receptor-α; COX-2: cyclooxygenase-2; PGE2: prostaglandin E2; PAI-1: plasminogen activator inhibitor-1; VEGF: vascular endothelial growth factor; NSAIDs: non-steroidal anti-inflammatory drugs; CXCR-4: chemokine receptor type 4; STAT 3: signal transducer and activator of transcription 3; CCL2: chemokine ligand 2; MCP-1: monocyte chemoattractant protein 1; CC5R: chemokine receptor type 5; CBDDC-TANs: cell-based drug delivery carrier—tumor-associated neutrophils; PTX: paclitaxel; pSTAT3: phosphorylated signal transducer and activator of transcription 3; Arg-1: arginase-1; CSF: colony stimulating factor; Akt/PKB: protein kinase B; TMZ: temozolomide; iNOS: inducible nitric oxide synthase; NO: nitric oxide; T cell: T lymphocyte; SOCS: suppressors of cytokine signaling; CTLA-4: cytotoxic T lymphocyte associated protein-4; PD-1: programmed cell death protein-1; CD80/86: cluster of differentiation 80/86; ERK 1/2: extracellular signal-regulated kinase 1/2; JNK: c-Jun N-terminal kinase; RAI14: retinoic acid-induced 14 protein; IFN β: interferon β; B cells: B lymphocytes; DCs: dendritic cells; sil-TRADD: silencing TNF receptor type 1-associated DEATH domain protein; sil-TLR4: silencing Toll-like receptor 4.</p>
Full article ">
20 pages, 3873 KiB  
Article
Tobacco Smoke and Electronic Cigarette Vapor Alter Enhancer RNA Expression That Can Regulate the Pathogenesis of Lung Squamous Cell Carcinoma
by Joseph C. Tsai, Omar A. Saad, Shruti Magesh, Jingyue Xu, Abby C. Lee, Wei Tse Li, Jaideep Chakladar, Mark M. Fuster, Eric Y. Chang, Jessica Wang-Rodriguez and Weg M. Ongkeko
Cancers 2021, 13(16), 4225; https://doi.org/10.3390/cancers13164225 - 23 Aug 2021
Cited by 5 | Viewed by 3553
Abstract
Tobacco is the primary etiologic agent in worsened lung squamous cell carcinoma (LUSC) outcomes. Meanwhile, it has been shown that etiologic agents alter enhancer RNAs (eRNAs) expression. Therefore, we aimed to identify the effects of tobacco and electronic cigarette (e-cigarette) use on eRNA [...] Read more.
Tobacco is the primary etiologic agent in worsened lung squamous cell carcinoma (LUSC) outcomes. Meanwhile, it has been shown that etiologic agents alter enhancer RNAs (eRNAs) expression. Therefore, we aimed to identify the effects of tobacco and electronic cigarette (e-cigarette) use on eRNA expression in relation to LUSC outcomes. We extracted eRNA counts from RNA-sequencing data of tumor/adjacent normal tissue and before/after e-cigarette tissue from The Cancer Genome Atlas (TCGA) and the Gene Expression Omnibus (GEO), respectively. Tobacco-mediated LUSC eRNAs were correlated to patient survival, clinical variables, and immune-associated elements. eRNA expression was also correlated to mutation rates through the Repeated Evaluation of Variables Conditional Entropy and Redundance (REVEALER) algorithm and methylated sites through methylationArrayAnalysis. Differential expression analysis was then completed for the e-cigarette data to compare with key tobacco-mediated eRNAs. We identified 684 downregulated eRNAs and 819 upregulated eRNAs associated with tobacco-mediated LUSC, specifically, with the cancer pathological stage. We also observed a decrease in immune cell abundance in tobacco-mediated LUSC. Yet, we found an increased association of eRNA expression with immune cell abundance in tobacco-mediated LUSC. We identified 16 key eRNAs with significant correlations to 8 clinical variables, implicating these eRNAs in LUSC malignancy. Furthermore, we observed that these 16 eRNAs were highly associated with chromosomal alterations and reduced CpG site methylation. Finally, we observed large eRNA expression upregulation with e-cigarette use, which corresponded to the upregulation of the 16 key eRNAs. Our findings provide a novel mechanism by which tobacco and e-cigarette smoke influences eRNA interactions to promote LUSC pathogenesis and provide insight regarding disease progression at a molecular level. Full article
(This article belongs to the Special Issue New Insight of Non-small Cell Lung Cancer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Project overview. Schematic of project workflow and analyses. Datasets were acquired from the Cancer Genome Atlas (TCGA) and the National Center for Biotechnology Information (NCBI) Gene Expression Omnibus (GEO). Tobacco-smoke-onset LUSC, nonsmoking-onset LUSC, and electronic cigarette smoke lung samples were analyzed in separate cohorts to limit any confounding variables. Differentially expressed eRNAs were filtered by log fold change and false discovery rate (log fold change (FC) &gt; 1 and false discovery rate (FDR) &lt; 0.05), and significantly dysregulated eRNAs were identified and used in survival analysis, immune infiltration analysis, and clinical variable correlations. A total of 16 key eRNAs were determined to be critical to the immune landscape and tumor phenotype of LUSC, based on previous analyses. These critical eRNAs were subsequently examined further for genomic alteration and methylation patterns.</p>
Full article ">Figure 2
<p>Correlation to Clinical Variables and Patient Survival. (<b>A</b>) The total amount of eRNAs identified to be correlated to patient survival and vital status in smoking-onset LUSC and nonsmoking-onset LUSC. Red correlates to smoking-onset LUSC data, and blue correlates to nonsmoking-onset LUSC data. Overall, there was a significantly greater amount of eRNAs associated with patient survival and vital status in the smoking-onset LUSC cohort. (<b>B</b>) Select cox proportional hazards regression plots (Kaplan–Meier) of patient survival rates (in days) in correlation to eRNA expression. (<b>C</b>) Bar graph representing the total significant correlations between eRNA expression and several clinical variables in both the smoking-onset LUSC cohorts and the nonsmoking-onset LUSC cohorts. Clinical variables investigated include tumor necrosis, cancer status, cancer pathological stage, cancer pathological m stage, cancer pathological n stage, cancer pathological t stage, stromal cell expression, and cancer vital status. (<b>D</b>) Select boxplots and scatterplots of eRNAs significantly correlated to clinical variables, including pathological stage T, pathological stage n, pathological stage m, tumor stromal cells, tumor necrosis, and tumor pathological stages 1, 2, 3, and 4. Boxplots plots were created using the Kruskal–Wallis test, while the scatterplots were created using the Spearman test, and all plots were filtered by <span class="html-italic">p</span>-value (<span class="html-italic">p</span>-value &lt; 0.05). The various correlations in the plots suggest that the upregulation of eRNAs identified from previous analyses are associated with advanced cancer pathological stage and are negatively correlated with tumor stromal cells and tumor necrosis percent expression.</p>
Full article ">Figure 3
<p>Correlation to immune-associated elements. (<b>A</b>) Bar graph comparing immune cell composition in smoking-onset LUSC and nonsmoking-onset LUSC. Red indicates smoking-onset LUSC samples, and blue refers to nonsmoking-onset LUSC samples. Overall, immune cell abundance was higher in nonsmoking LUSC samples. Error bars indicate the standard deviation. Homoscedastic <span class="html-italic">t</span>-tests were performed to compare immune cell compositions of the cohorts (** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Bar graph comparing significant eRNA correlations to immune cell abundance in smoking-onset LUSC and nonsmoking-onset LUSC. Red indicates smoking-onset LUSC samples, and blue refers to nonsmoking-onset LUSC samples. The data suggest that there is a greater amount of significant eRNAs associated with immune cells in smoking-onset LUSC samples. (<b>C</b>) Bar graph summarizing immune cell correlations to the 16 critical eRNAs previously identified to be implicated in the LUSC tumor phenotype and immune environment. Blue indicates relative upregulation, while red refers to downregulation in both eRNA expression and abundance of immune cells. (<b>D</b>) Selected scatter plots illustrating the correlations between expression of key eRNAs and immune cell abundance. eRNAs chr19.38575229.38575296 and chr10.21430870.21431276 were downregulated with increasing CD4 and B cell expression, while chr9.27155069.27155179 was upregulated with increasing macrophage expression. The positive and negative correlations between these eRNAs and immune cell abundances suggest potential roles and interactions they might have in promoting LUSC pathogenesis and proliferation.</p>
Full article ">Figure 4
<p>Identification and Analysis of 16 eRNAs identified to be critical to LUSC immune landscape and tumor phenotype. (<b>A</b>) Venn diagram summarizing the significant correlations of the selected 16 key eRNAs to patient survival, immune cell infiltration, and tumor status. Each of the 16 selected eRNAs has a high degree of association with various clinical and molecular measures of tumor severity, implicating their potential role in furthering LUSC pathogenesis and metastasis. (<b>B</b>) Volcano plot across all significantly dysregulated eRNAs illustrating the statistical significance versus the magnitude of change (log fold change) in survival-correlated eRNAs, tumor-presence-correlated eRNAs, and the identified 16 critical eRNAs. The 16 key eRNAs demonstrate both a high degree of statistical significance and magnitude of change in expression. (<b>C</b>) Associations between genomic alterations and eRNA expression in smoking-onset LUSC samples, nonsmoking-onset LUSC, and adjacent normal LUSC samples. Correlations were determined using the REVEALER algorithm (<span class="html-italic">p</span>-value &lt; 0.05 and CIC &gt; |0.3|).</p>
Full article ">Figure 5
<p>Genomic alteration correlation to 16 key eRNAs via REVEALER (<b>A</b>) Mutation heatmaps with red on the left side indicating that an eRNA is overabundant, while red on the left side indicates that the eRNA is under abundant. The following rows below the heatmap specify copy number variant (CNV) type/genetic variants (alteration, deletion, mutation) across all samples. Each bar represents the presence or absence of a CNV in one sample. Correlations were measured using the Repeated Evaluation of Variables Conditional Entropy and Redundancy (REVEALER) algorithm (CIC &gt; |0.3|, <span class="html-italic">p</span> &lt; 0.05); 15/16 plots are shown here as the algorithm failed to produce a plot for chr12.51786511.51786834. (<b>B</b>) Boxplots correlating eRNA expression with the extent of methylation at CpG sites. In this analysis, the M-values were converted to a binary and categorical variable of “HIGH” and “LOW” in order to conduct the Kruskal–Wallis statistical test. The median abundance level for each eRNA across all samples was calculated. Values lower than the median value were assigned an abundance level of “LOW,” and values above the median value were assigned an abundance level of “HIGH.” The <span class="html-italic">x</span>-axis refers to the extent of methylation at the CpG site, and the <span class="html-italic">y</span>-axis refers to the eRNA expression level. (<b>C</b>) Heatmaps comparing significantly differentially expressed eRNAs in both endothelial microparticles and alveolar macrophages before/after electronic cigarette inhalation (<span class="html-italic">n</span> = 10).</p>
Full article ">
12 pages, 601 KiB  
Review
Challenges for Better Diagnosis and Management of Pancreatic and Biliary Tract Cancers Focusing on Blood Biomarkers: A Systematic Review
by Hiroto Tominaga, Juntaro Matsuzaki, Chihiro Oikawa, Kensho Toyoshima, Haruki Manabe, Eriko Ozawa, Atsushi Shimamura, Riko Yokoyama, Yusuke Serizawa, Takahiro Ochiya and Yoshimasa Saito
Cancers 2021, 13(16), 4220; https://doi.org/10.3390/cancers13164220 - 23 Aug 2021
Cited by 1 | Viewed by 2959
Abstract
Background: pancreatic cancer (PCa) and biliary tract cancer (BTC) are cancers with a poor prognosis and few effective treatments. One of the reasons for this is late detection. Many researchers are tackling to develop non-invasive biomarkers for cancer, but few are specific for [...] Read more.
Background: pancreatic cancer (PCa) and biliary tract cancer (BTC) are cancers with a poor prognosis and few effective treatments. One of the reasons for this is late detection. Many researchers are tackling to develop non-invasive biomarkers for cancer, but few are specific for PCa or BTC. In addition, genetic abnormalities occur in cancer tissues, which ultimately affect the expression of various molecules. Therefore, it is important to identify molecules that are altered in PCa and BTC. For this systematic review, a systematic review of Medline and Embase to select biomarker studies of PCa and BTC patients was conducted. Results: after reviewing 72 studies, 79 biomarker candidates were identified, including 22 nucleic acids, 43 proteins, and 14 immune cell types. Of the 72 studies, 61 examined PCa, and 11 examined BTC. Conclusion: PCa and BTC are characterized by nucleic acid, protein, and immune cell profiles that are markedly different from those of healthy subjects. These altered molecules and cell subsets may serve as cancer-specific biomarkers, particularly in blood. Further studies are needed to better understand the diagnosis and prognosis of PCa and BTC. Full article
(This article belongs to the Special Issue Novel Biomarkers of Gastrointestinal Cancer)
Show Figures

Figure 1

Figure 1
<p>Flow diagram of literature search.</p>
Full article ">
9 pages, 2315 KiB  
Article
Duration of Reduced CA19-9 Levels Is a Better Prognostic Factor Than Its Rate of Reduction for Unresectable Locally Advanced Pancreatic Cancer
by Ko Tomishima, Shigeto Ishii, Toshio Fujisawa, Muneo Ikemura, Hiroto Ota, Daishi Kabemura, Mako Ushio, Taito Fukuma, Sho Takahashi, Wataru Yamagata, Yusuke Takasaki, Akinori Suzuki, Koichi Ito, Hiroaki Saito, Akihito Nagahara and Hiroyuki Isayama
Cancers 2021, 13(16), 4224; https://doi.org/10.3390/cancers13164224 - 22 Aug 2021
Cited by 10 | Viewed by 6359
Abstract
A decrease in carbohydrate antigen (CA) 19-9 levels has been proposed as a prognostic marker for survival and recurrence in patients with pancreatic cancer. We evaluated the association between duration of reduced CA 19-9 levels during 6 months after treatment and long-term survival [...] Read more.
A decrease in carbohydrate antigen (CA) 19-9 levels has been proposed as a prognostic marker for survival and recurrence in patients with pancreatic cancer. We evaluated the association between duration of reduced CA 19-9 levels during 6 months after treatment and long-term survival for 79 patients with unresectable locally advanced pancreatic cancer (LAPC). We calculated the differences between pretreatment and monthly CA19-9 levels. We categorized 71 patients with decreases in CA19-9 levels into three groups based on the duration of these reduced levels (>2, >3, and >4 months). The cut-off level for long-term (more than 2 years) survival was identified as a 44% reduction from the baseline, using a ROC curve. A reduction duration >2 months was not associated with overall survival (p = 0.1), while >3 months was significantly associated with survival (p =.04). In multivariate analysis, a reduction duration >3 months predicted a good long-term prognosis (odds ratio = 5.75; 95% confidence interval = 1.47–22.36; p < 0.01). In patients with unresectable LAPC, the duration of reduced CA19-9 levels for more than 3 months, rather than the rate of reduction in CA19-9 levels, during 6 months after treatment was significantly associated with good prognosis. Full article
(This article belongs to the Special Issue Advances in Pancreatic Cancer Imaging)
Show Figures

Figure 1

Figure 1
<p>Patients flow chart. LAPC: locally advanced pancreatic cancer; BSC: best supportive care; RR-C: reduction ratio of CA 19-9.</p>
Full article ">Figure 2
<p>Among CA 19-9 reduction, relationship between within 44% reduction and over 44% reduction.</p>
Full article ">Figure 3
<p>Among CA 19-9 reduction, relationship between &gt;44% reduction of these durations (&gt;2 months, &gt;3 months, and &gt;4 months) and OS.</p>
Full article ">Figure 4
<p>Relationship between &gt;44% reduction duration (&gt;3 months) and ODM. (ODM: occurrence of distant metastasis).</p>
Full article ">Figure 5
<p>Analysis of the reduction ratio of CA19-9 (RR-C) in each survival. (Patients who survive within 12 months (<b>A</b>), from 12 months to 24 months (<b>B</b>), and more than 24 months (<b>C</b>).)</p>
Full article ">
12 pages, 801 KiB  
Article
Validating Methylated HOXA9 in Bronchial Lavage as a Diagnostic Tool in Patients Suspected of Lung Cancer
by Sara W. C. Wen, Rikke F. Andersen, Kristian Rasmussen, Caroline Brenner Thomsen, Torben Frøstrup Hansen, Line Nederby, Henrik Hager, Anders Jakobsen and Ole Hilberg
Cancers 2021, 13(16), 4223; https://doi.org/10.3390/cancers13164223 - 22 Aug 2021
Cited by 4 | Viewed by 1955
Abstract
Diagnosing lung cancer requires invasive procedures with high risk of complications. Methylated tumor DNA in bronchial lavage has previously shown potential as a diagnostic biomarker. We aimed to develop and validate methylated HOXA9 in bronchial lavage as a diagnostic biomarker of lung cancer. [...] Read more.
Diagnosing lung cancer requires invasive procedures with high risk of complications. Methylated tumor DNA in bronchial lavage has previously shown potential as a diagnostic biomarker. We aimed to develop and validate methylated HOXA9 in bronchial lavage as a diagnostic biomarker of lung cancer. Participants were referred on suspicion of lung cancer. Ten mL lavage fluid was collected at bronchoscopy for analysis of methylated HOXA9 based on droplet digital PCR according to our previously published method. HOXA9 status was compared with the final diagnosis. The Discovery and Validation cohorts consisted of 101 and 95 consecutively enrolled participants, respectively. In the discovery cohort, the sensitivity and specificity were 73.1% (95% CI 60.9–83.2%) and 85.3% (95% CI 68.9–95.0%), respectively. In the validation cohort, the values were 80.0% (95% CI 66.3–90.0%) and 75.6% (95% CI 60.5–87.1%), respectively. A multiple logistic regression model including age, smoking status, and methylated HOXA9 status resulted in an AUC of 84.9% (95% CI 77.3–92.4%) and 85.9% (95% CI 78.4–93.4%) for the Discovery and Validation cohorts, respectively. Methylated HOXA9 in bronchial lavage holds potential as a supplementary tool in the diagnosis of lung cancer with a clinically relevant sensitivity and specificity. It remained significant when adjusting for age and smoking status. Full article
(This article belongs to the Section Cancer Biomarkers)
Show Figures

Figure 1

Figure 1
<p>Flow charts illustrating how participants were selected and excluded for (<b>a</b>) the Discovery cohort and (<b>b</b>) the Validation cohort.</p>
Full article ">Figure 2
<p>ROC analysis of the multiple logistic regression models based on clinical variables with and without methylated HOXA9. Receiver operating characteristics (ROC) curves illustrating the diagnostic performance of the two logistic regression models in (<b>a</b>) the Discovery cohort and (<b>b</b>) the Validation cohort. The blue line represents the multiple logistic regression model based on clinical variables, while the red line represents the model including methylated <span class="html-italic">HOXA9</span> status.</p>
Full article ">Figure 2 Cont.
<p>ROC analysis of the multiple logistic regression models based on clinical variables with and without methylated HOXA9. Receiver operating characteristics (ROC) curves illustrating the diagnostic performance of the two logistic regression models in (<b>a</b>) the Discovery cohort and (<b>b</b>) the Validation cohort. The blue line represents the multiple logistic regression model based on clinical variables, while the red line represents the model including methylated <span class="html-italic">HOXA9</span> status.</p>
Full article ">
11 pages, 2289 KiB  
Article
The “K-Sign”—A Novel CT Finding Suggestive before the Appearance of Pancreatic Cancer
by Yuko Kobashi, Masateru Uchiyama and Junichi Matsui
Cancers 2021, 13(16), 4222; https://doi.org/10.3390/cancers13164222 - 22 Aug 2021
Cited by 6 | Viewed by 2614
Abstract
Pancreatic invasive ductal adenocarcinoma (PDAC) has a poor prognosis, and the detection of PDAC during the early stage is thought to improve prognosis. In this study, we retrospectively investigated pancreatic morphological abnormalities that lead to the early diagnosis of PDAC with computed tomography [...] Read more.
Pancreatic invasive ductal adenocarcinoma (PDAC) has a poor prognosis, and the detection of PDAC during the early stage is thought to improve prognosis. In this study, we retrospectively investigated pancreatic morphological abnormalities that lead to the early diagnosis of PDAC with computed tomography (CT) imaging. In total, 41 out of 308 patients diagnosed with pancreatic cancer between 2011 and 2017 in our institution were enrolled. As a control group for the group with pancreatic cancer, 4277 patients without pancreato-biliary diseases were enrolled. We retrospectively reviewed and analyzed the clinical data including patient characteristics, the clinical course and preoperative CT imaging with pancreatic morphological features. Out of 41 patients, 24 patients (58.5%) showed local K-shaped constriction of the pancreatic parenchyma “K-sign” on preoperative CT images. Eight patients (19.5%) showed localized fatty change. Out of 4277 control patients, seven patients (0.16%) showed K-sign. “K-sign” may be used for the early diagnosis of PDAC by CT imaging. Full article
(This article belongs to the Special Issue Recent Highly Advanced Surgery for Pancreatic Cancer)
Show Figures

Figure 1

Figure 1
<p>Enrollment.</p>
Full article ">Figure 2
<p>Illustration and definition of K-sign. K-sign means a localized constriction of the pancreatic parenchyma. It looks “K” on axial CT image.</p>
Full article ">Figure 3
<p>A representative case with K-sign (pStage IIB). (<b>a</b>) October 2008: pancreatic body shows no evidence of pancreatic abnormality. (<b>b</b>) June 2014: K-sign (yellow arrows) and the main pancreatic duct dilatation (an arrowhead) are shown in the pancreatic body. (<b>c</b>) August 2016: pancreatic body cancer with main pancreatic duct dilatation is shown in the K-sign area (a yellow circle and arrow head). (<b>d</b>) Pancreatosplenectomy was performed in September 2016. Macroscopic specimen did not show visible tumor invasion to the pancreatic surface and proof of K-sign. (<b>e</b>) The cross-section surface of the macroscopic specimen showed pancreatic cancer around the main pancreatic duct (yellow arrows).</p>
Full article ">Figure 4
<p>(<b>A</b>) Group A (a representative patient with K-sign in the pancreatic body): (<b>a</b>) April 2005: pancreatic body shows no evidence of pancreatic abnormality (a yellow circle); (<b>b</b>) January 2013: K-sign is shown in the pancreatic body (a yellow circle); (<b>c</b>) December 2016: pancreatic body cancer with main pancreatic duct dilatation is shown in the K-sign area (a yellow circle). (<b>B</b>) Group A (a representative patient with K-sign in the pancreatic head): (<b>a</b>) May 2004: pancreatic head looks normal on this CT (a yellow circle); (<b>b</b>) September 2012: K-sign is shown in the pancreatic head (a yellow arrow), (<b>c</b>) July 2013: pancreatic head cancer is developed at the site of K-sign (a yellow circle). (<b>C</b>) Group A (a representative patient with K-sign in the pancreatic tail): (<b>a</b>) October 2007: K-sign is shown in the pancreatic tail (a yellow circle); (<b>b</b>) October 2011: K-sign expanded compared with the previous CT in 2007 (a yellow circle); (<b>c</b>) August 2012: pancreatic tail cancer with main pancreatic duct dilatation is shown (a yellow circle). (<b>D</b>) Group A (a representative patient with K-sign and patented main pancreatic duct): (<b>a</b>) March 2003: pancreatic body looks normal on this CT (a yellow circle); (<b>b</b>) August 2011: K-sign (a yellow arrow) and main pancreatic duct (a yellow triangle) are shown; (<b>c</b>) August 2013: pancreatic body cancer is shown at the site of K-sign (a yellow circle).</p>
Full article ">Figure 5
<p>(<b>A</b>) Group B (a representative patient with localized fatty change in the pancreatic head): (<b>a</b>) July 2006: pancreatic head looks normal on this CT (a yellow circle); (<b>b</b>) April 2013: localized fatty change is shown in the pancreatic head (a yellow circle); (<b>c</b>) October 2014: pancreatic head cancer is shown at the site of localized fatty change (a yellow circle). (<b>B</b>) Group B (a representative patient with localized fatty change in the pancreatic uncinate): (<b>a</b>) August 2014: localized fatty change is shown in the pancreatic uncinate (a yellow circle); (<b>b</b>) July 2016: localized fatty change expanded compared with the previous CT in 2014 (a yellow circle); (<b>c</b>) January 2017: pancreatic uncinate cancer is shown at the site of localized fatty change (a yellow circle).</p>
Full article ">Figure 6
<p>Group C (a representative patient with no abnormality): (<b>a</b>) April 2009: pancreatic head looks normal with well enhancement (a yellow circle); (<b>b</b>) December 2017: pancreatic head cancer is shown (yellow arrows).</p>
Full article ">
13 pages, 793 KiB  
Article
Survival Impact of Chronic Obstructive Pulmonary Disease or Acute Exacerbation on Patients with Rectal Adenocarcinoma Undergoing Curative Resection: A Propensity Score-Matched, Population-Based Cohort Study
by Jiaqiang Zhang, Kuo-Chin Chiu, Wei-Chun Lin and Szu-Yuan Wu
Cancers 2021, 13(16), 4221; https://doi.org/10.3390/cancers13164221 - 22 Aug 2021
Cited by 1 | Viewed by 2334
Abstract
Purpose: The survival effect of current smoking-related chronic obstructive pulmonary disease (COPD) and COPD with acute exacerbation (COPDAE) is unclear for patients with rectal adenocarcinoma undergoing curative resection. Methods: We recruited patients with clinical stage I–IIIC rectal adenocarcinoma from the Taiwan Cancer Registry [...] Read more.
Purpose: The survival effect of current smoking-related chronic obstructive pulmonary disease (COPD) and COPD with acute exacerbation (COPDAE) is unclear for patients with rectal adenocarcinoma undergoing curative resection. Methods: We recruited patients with clinical stage I–IIIC rectal adenocarcinoma from the Taiwan Cancer Registry Database who had received surgery. The Cox proportional hazards model was used to analyze all-cause mortality. We categorized the patients into two groups by using propensity score matching based on COPD status to compare overall survival outcomes: Group 1 (current smokers with COPD) and Group 2 (nonsmokers without COPD). Results: In the multivariate Cox regression analyses, the adjusted hazard ratio (aHR; 95% confidence interval (CI)) of all-cause mortality for Group 1 compared with Group 2 was 1.25 (1.04–1.51). The aHRs (95% cis) of all-cause mortality for frequency of ≥1 hospitalizations for COPDAE or ≥2 hospitalizations within 1 year before diagnosis were 1.17 (1.05–1.51) and 1.48 (1.03–2.41) compared with no COPDAE in patients with rectal adenocarcinoma undergoing curative resection. Conclusion: In patients with rectal adenocarcinoma undergoing curative resection, being a current smoker with COPD (Group 1) was associated with worse survival outcomes than being a nonsmoker without COPD (Group 2). Being hospitalized at least once for COPDAE within 1 year before the diagnosis of rectal adenocarcinoma is an independent risk factor for poor overall survival in these patients, and a higher number of hospitalizations for COPDAE within 1 year before diagnosis was associated with poorer survival. Full article
(This article belongs to the Special Issue New Insights into Colorectal Cancer)
Show Figures

Figure 1

Figure 1
<p>Kaplan–Meier survival curves of patients with rectal adenocarcinoma with or without smoking-related chronic obstructive pulmonary disease before surgery after propensity score matching, COPD, chronic obstruction pulmonary disease.</p>
Full article ">Figure 2
<p>Kaplan–Meier survival curves of patients with rectal adenocarcinoma with hospitalization(s) for acute exacerbations of chronic obstructive pulmonary disease within 1 year before surgery. COPDAE, chronic obstruction pulmonary disease with acute exacerbation.</p>
Full article ">
22 pages, 1551 KiB  
Article
Analytical Performance of NGS-Based Molecular Genetic Tests Used in the Diagnostic Workflow of Pheochromocytoma/Paraganglioma
by Balazs Sarkadi, Istvan Liko, Gabor Nyiro, Peter Igaz, Henriett Butz and Attila Patocs
Cancers 2021, 13(16), 4219; https://doi.org/10.3390/cancers13164219 - 22 Aug 2021
Cited by 4 | Viewed by 3052
Abstract
Next Generation Sequencing (NGS)-based methods are high-throughput and cost-effective molecular genetic diagnostic tools. Targeted gene panel and whole exome sequencing (WES) are applied in clinical practice for assessing mutations of pheochromocytoma/paraganglioma (PPGL) associated genes, but the best strategy is debated. Germline mutations of [...] Read more.
Next Generation Sequencing (NGS)-based methods are high-throughput and cost-effective molecular genetic diagnostic tools. Targeted gene panel and whole exome sequencing (WES) are applied in clinical practice for assessing mutations of pheochromocytoma/paraganglioma (PPGL) associated genes, but the best strategy is debated. Germline mutations of at the least 18 PPGL genes are present in approximately 20–40% of patients, thus molecular genetic testing is recommended in all cases. We aimed to evaluate the analytical and clinical performances of NGS methods for mutation detection of PPGL-associated genes. WES (three different library preparation and bioinformatics workflows) and an in-house, hybridization based gene panel (endocrine-onco-gene-panel- ENDOGENE) was evaluated on 37 (20 WES and 17 ENDOGENE) samples with known variants. After optimization of the bioinformatic workflow, 61 additional samples were tested prospectively. All clinically relevant variants were validated with Sanger sequencing. Target capture of PPGL genes differed markedly between WES platforms and genes tested. All known variants were correctly identified by all methods, but methods of library preparations, sequencing platforms and bioinformatical settings significantly affected the diagnostic accuracy. The ENDOGENE panel identified several pathogenic mutations and unusual genotype–phenotype associations suggesting that the whole panel should be used for identification of genetic susceptibility of PPGL. Full article
Show Figures

Figure 1

Figure 1
<p>Coverage of PPGL associated genes by whole exome sequencing. Data is represented as mean ± SD.</p>
Full article ">Figure 2
<p>Schematic presentation of sequencing reads containing the variants SDHB(NM_003000.3):c.263C&gt;T (p.Thr88Ile), SDHB(NM_003000.3):c.268C&gt;G (p.Arg90Gly), SDHB(NM_003000.3):c.271_273del(p.Arg91del) (Case 38, (<b>A</b>)) and NF1(NM_001042492.2):c.5047_5053delinsGGAG(p.Asn1683_Ser1684_Trp1685delinsGlyGly) (Case 24, (<b>B</b>)). Each line represents one read. Half of the reads shows normal sequence (<b>upper part</b>) and half of the reads (<b>lower part</b>) show the mutated sequences.</p>
Full article ">
12 pages, 3049 KiB  
Article
Surgical Management of Jugular Foramen Schwannomas
by Amir Kaywan Aftahy, Maximilian Groll, Melanie Barz, Denise Bernhardt, Stephanie E. Combs, Bernhard Meyer, Chiara Negwer and Jens Gempt
Cancers 2021, 13(16), 4218; https://doi.org/10.3390/cancers13164218 - 22 Aug 2021
Cited by 14 | Viewed by 4039
Abstract
Background: Resection of jugular foramen schwannomas (JFSs) with minimal cranial nerve (CN) injury remains difficult. Reoperations in this vital region are associated with severe CN deficits. Methods: We performed a retrospective analysis at a tertiary neurosurgical center of patients who underwent surgery for [...] Read more.
Background: Resection of jugular foramen schwannomas (JFSs) with minimal cranial nerve (CN) injury remains difficult. Reoperations in this vital region are associated with severe CN deficits. Methods: We performed a retrospective analysis at a tertiary neurosurgical center of patients who underwent surgery for JFSs between June 2007 and May 2020. We included nine patients (median age 60 years, 77.8% female, 22.2% male). Preoperative symptoms included hearing loss (66.6%), headache (44.4%), hoarseness (33.3%), dysphagia (44.4%), hypoglossal nerve palsy (22.2%), facial nerve palsy (33.3%), extinguished gag reflex (22.2%), and cerebellar dysfunction (44.4%). We observed Type A, B, C, and D tumors in 3, 1, 1, and 4 patients, respectively. A total of 77.8% (7/9) underwent a retrosigmoid approach, and 33.3% (3/9) underwent an extreme lateral infrajugular transcondylar (ELITE) approach. Gross total resection (GTR) was achieved in all cases. The rate of shunt-dependent hydrocephalus was 22.2% (2/9). No further complications requiring surgical intervention occurred during follow-up. The median follow-up time was 16.5 months (range 3–84 months). Conclusions: Considering the satisfying outcome, the GTR of JFSs is feasible in performing well-known skull base approaches. Additional invasive and complicated approaches were not needed. Radiosurgery may be an effective alternative for selected patients. Full article
(This article belongs to the Special Issue Skull Base Tumours)
Show Figures

Figure 1

Figure 1
<p>A 66-year-old female patient presenting with headache, vertigo, and cerebellar dysfunction. (<b>A</b>) Preoperative axial and (<b>B</b>) coronal T1-weighted gadolinium-enhanced MRI, displaying cystic JFS (arrows), involving the CPA (Samii Type A). (<b>C</b>) Postoperative axial and (<b>D</b>) coronal MRI control, indicating complete resection (arrows) through a classic retrosigmoid approach. Postoperatively, mild facial nerve palsy (House and Brackmann II) occurred and remained during follow-up.</p>
Full article ">Figure 2
<p>(<b>A</b>) For the dorsolateral approach, a lazy S or bigger C incision is used, about 1 cm posterior to the mastoid bone, extending inferiorly along the hairline. For the anterolateral ELITE procedure, a retroauricular curvilinear C-shaped, or question-mark-shaped, skin incision is begun approximately 2 to 3 cm posterior to the upper border of the ear. Inferiorly, this incision is carried down into the neck, traversing the border of the sternocleidomastoid muscle and running parallel to the body of the mandible (fingerbreadths below). For the dorsolateral approach, the sternocleidomastoid muscle is retracted anteriorly; for the anterolateral approach, it is retracted posteriorly. (<b>B</b>) Exposure by different techniques: (1) suprajugular, infralabyrinthine; (2) infrajugular, transcondylar; and (3) high cervical. SS, sigmoid sinus; ICA, internal carotid artery; IJV, internal jugular vein. (<b>C</b>) Anterolateral techniques with high cervical dissection. AICA, anterior inferior cerebellar artery; PICA, posterior inferior cerebellar artery; VA, vertebral artery; JB, jugular bulb.</p>
Full article ">Figure 3
<p>A 56-year-old female patient presented with headache, cerebellar dysfunctions, hearing loss, hoarseness, and extinguished gag reflex. (<b>A</b>) Preoperative axial and (<b>B</b>) coronal T1-weighted gadolinium-enhanced MRI, showing an impressive JFS (arrows) with a quasi-dumbbell shape and both intra- and extracranial components through the JF (Samii Type D). (<b>C</b>) Postoperative axial and (<b>D</b>) coronal MRI control, indicating complete resection (arrows) through a dorsolateral ELITE approach. Postoperatively, no new deficits occurred, and the patient recovered from the cerebellar dysfunctions, hoarseness, and extinguished gag reflex.</p>
Full article ">Figure 4
<p>A 26-year-old male patient presented with progressive hoarseness, dysphagia, and extinguished gag reflex. (<b>A</b>) Preoperative axial (arrow), (<b>B</b>) sagittal (circle), and (<b>C</b>) coronal (circle) T1-weighted gadolinium-enhanced MRI, showing the JFS with intracranial extension (Samii Type B). (<b>D</b>) Postoperative axial, (<b>E</b>) sagittal, and (<b>F</b>) coronal MRI control, indicating complete resection through a dorsolateral ELITE approach. Postoperatively, no new deficits occurred. He suffered from postoperative temporary vertigo; hoarseness showed slight improvement during follow-up.</p>
Full article ">Figure 5
<p>A 46-year-old male patient presented with headache, hoarseness, dysphagia, and extinguished gag reflex. (<b>A</b>) Preoperative axial (arrow), (<b>B</b>) sagittal (circle), and (<b>C</b>) coronal (circle) T1-weighted gadolinium-enhanced MRI, showing a massive space-occupying extracranial jugular foramen schwannoma with extension into the JF (Samii Type C). (<b>D</b>) Postoperative axial, (<b>E</b>) sagittal, and (<b>F</b>) coronal MRI control, indicating complete resection through an anterolateral ELITE approach. Postoperatively, no new deficits occurred, and the patient recovered from the hoarseness and extinguished gag reflex. Dysphagia remained but improved during follow-up as well.</p>
Full article ">
Previous Issue
Back to TopTop