[go: up one dir, main page]

 
 

Topic Editors

Department of Anatomy, Cell Biology, and Physiological Sciences, Faculty of Medicine, American University of Beirut, Beirut 1107-2020, Lebanon
The Arkadi M. Rywlin Department of Pathology and Laboratory Medicine, Mount Sinai Medical Center, Miami Beach, FL 33140, USA

Molecular and Cellular Mechanisms of Cancers: Colorectal Cancer

Abstract submission deadline
closed (28 December 2022)
Manuscript submission deadline
closed (28 February 2023)
Viewed by
32059

Topic Information

Dear Colleagues,

Colorectal cancer (CRC) is the third most common cancer in both men and women and the third leading cause of cancer-related deaths in the United States. It has always been recognized as a heterogeneous disease, where the clinical course is unique to every patient in terms of prognosis and treatment response. With the rapid advancement in the medical field and the wide use of next-generation sequencing techniques to understand CRC, researchers and physicians are moving away from a “one size fits all” strategy in treating the disease to identifying novel biomarkers that can be targeted to specifically treat every patient. Nevertheless, treatment of most CRC patients requires a multimodality approach that includes surgical intervention, radiation therapy, and chemotherapy. It is thus crucial to decipher the molecular and cellular mechanisms underlying the initiation and progression of this intractable cancer and accordingly identify the unique biomarkers associated with it to aid in cancer diagnosis and improve prognosis. This approach in the management of CRC—by applying basic and translational research with bed-side clinical research—is the hub for “personalized medicine” in the 21st century, and it is an area of great interest to all physicians and researchers working in the field, in particular molecular pathologists and hematology-oncologists. This issue focuses on research and experiences related to CRC. This may include deciphering the molecular and cellular mechanisms underlying the initiation, promotion, and progression of CRC in addition to the identification of novel cancer biomarkers and therapeutic targets, with focus on the prospects for improving personalized cancer care using new technologies to enhance cancer patient diagnoses, management, and outcomes. In this Special Issue, we invite researchers in molecular pathology, hematology–oncology, hematopathology, stem cells, and other fields of cancer research to submit high-quality empirical papers or reviews related to the issues in this research area.

Dr. Wassim Abou-Kheir
Dr. Hisham Bahmad
Topic Editors

Keywords

  • colorectal cancer
  • biomarkers
  • therapeutic target
  • personalized medicine
  • targeted therapy
  • genetic aberrations
  • molecular signatures
  • next generation sequencing

Participating Journals

Journal Name Impact Factor CiteScore Launched Year First Decision (median) APC
Biomolecules
biomolecules
4.8 9.4 2011 16.3 Days CHF 2700
Cancers
cancers
4.5 8.0 2009 16.3 Days CHF 2900
Cells
cells
5.1 9.9 2012 17.5 Days CHF 2700
Journal of Molecular Pathology
jmp
- - 2020 25.4 Days CHF 1000
Organoids
organoids
- - 2022 15.0 days * CHF 1000

* Median value for all MDPI journals in the first half of 2024.


Preprints.org is a multidiscipline platform providing preprint service that is dedicated to sharing your research from the start and empowering your research journey.

MDPI Topics is cooperating with Preprints.org and has built a direct connection between MDPI journals and Preprints.org. Authors are encouraged to enjoy the benefits by posting a preprint at Preprints.org prior to publication:

  1. Immediately share your ideas ahead of publication and establish your research priority;
  2. Protect your idea from being stolen with this time-stamped preprint article;
  3. Enhance the exposure and impact of your research;
  4. Receive feedback from your peers in advance;
  5. Have it indexed in Web of Science (Preprint Citation Index), Google Scholar, Crossref, SHARE, PrePubMed, Scilit and Europe PMC.

Published Papers (12 papers)

Order results
Result details
Journals
Select all
Export citation of selected articles as:
22 pages, 13691 KiB  
Article
Pyrroline-5-Carboxylate Reductase-2 Promotes Colorectal Carcinogenesis by Modulating Microtubule-Associated Serine/Threonine Kinase-like/Wnt/β-Catenin Signaling
by Raju Lama Tamang, Balawant Kumar, Sagar M. Patel, Ishwor Thapa, Alshomrani Ahmad, Vikas Kumar, Rizwan Ahmad, Donald F. Becker, Dundy (Kiran) Bastola, Punita Dhawan and Amar B. Singh
Cells 2023, 12(14), 1883; https://doi.org/10.3390/cells12141883 - 18 Jul 2023
Cited by 2 | Viewed by 1941
Abstract
Background: Despite significant progress in clinical management, colorectal cancer (CRC) remains the third most common cause of cancer-related deaths. A positive association between PYCR2 (pyrroline-5-carboxylate reductase-2), a terminal enzyme of proline metabolism, and CRC aggressiveness was recently reported. However, how PYCR2 promotes colon [...] Read more.
Background: Despite significant progress in clinical management, colorectal cancer (CRC) remains the third most common cause of cancer-related deaths. A positive association between PYCR2 (pyrroline-5-carboxylate reductase-2), a terminal enzyme of proline metabolism, and CRC aggressiveness was recently reported. However, how PYCR2 promotes colon carcinogenesis remains ill understood. Methods: A comprehensive analysis was performed using publicly available cancer databases and CRC patient cohorts. Proteomics and biochemical evaluations were performed along with genetic manipulations and in vivo tumor growth assays to gain a mechanistic understanding. Results: PYCR2 expression was significantly upregulated in CRC and associated with poor patient survival, specifically among PYCR isoforms (PYCR1, 2, and 3). The genetic inhibition of PYCR2 inhibited the tumorigenic abilities of CRC cells and in vivo tumor growth. Coinciding with these observations was a significant decrease in cellular proline content. PYCR2 overexpression promoted the tumorigenic abilities of CRC cells. Proteomics (LC-MS/MS) analysis further demonstrated that PYCR2 loss of expression in CRC cells inhibits survival and cell cycle pathways. A subsequent biochemical analysis supported the causal role of PYCR2 in regulating CRC cell survival and the cell cycle, potentially by regulating the expression of MASTL, a cell-cycle-regulating protein upregulated in CRC. Further studies revealed that PYCR2 regulates Wnt/β-catenin-signaling in manners dependent on the expression of MASTL and the cancer stem cell niche. Conclusions: PYCR2 promotes MASTL/Wnt/β-catenin signaling that, in turn, promotes cancer stem cell populations and, thus, colon carcinogenesis. Taken together, our data highlight the significance of PYCR2 as a novel therapeutic target for effectively treating aggressive colon cancer. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>PYCR2 expression increases significantly in colon cancer. (<b>A</b>–<b>C</b>) PYCR2 mRNA expression in Asian and European CRC cohorts (<span class="html-italic">p</span> = 0.000201 for Korean cohort, <span class="html-italic">p</span> = 2.003 × 10<sup>−11</sup> for French cohort, and <span class="html-italic">p</span> = 0.01577093 for Amsterdam cohort). (<b>D</b>) Analysis of PYCR2 protein expression in colorectal cancer patients in the CPTAC database (adjacent normal vs. primary tumor, <span class="html-italic">p</span> = 1.49 × 10<sup>−43</sup>). (<b>E</b>) PYCR2 protein expression in different stages of colorectal cancer (normal vs. stage 1, <span class="html-italic">p</span> = 6.5 × 10<sup>−5</sup>; normal vs. stage 2, <span class="html-italic">p</span> = 2.13 × 10<sup>−22</sup>; normal vs. stage 3, <span class="html-italic">p</span> = 2.94 × 10<sup>−16</sup>; and normal vs. stage 4, <span class="html-italic">p</span> = 5.49 × 10<sup>−4</sup>). (<b>F</b>) Representative image showing PYCR2 expression in colon tumors of APC<sup>min</sup> mice. (<b>G</b>) Representative images of immunohistochemical analysis of PYCR2 expression in colon adenoma and adenocarcinoma in comparison to normal adjacent human colon. (<b>H</b>) Quantitative analysis of PYCR2 expression in human colon polyps and CRC samples. **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>PYCR2 protein expression is significantly upregulated in all histological types of CRC adenomas. (<b>Ai</b>–<b>Av</b>) Representative images of immunohistochemical analysis of PYCR2 expression in different types of colon adenomas and adjacent normal colon (TMA, N#109). (<b>B</b>) Scoring analysis of PYCR2 immunostaining intensity in normal adjacent colon vs. colon adenoma (<span class="html-italic">p &lt;</span> 0.0001 for SSA, <span class="html-italic">p</span> = 0.002 for TV, <span class="html-italic">p</span> = 0.0015 for TA, and TA + SSA respectively). The data are presented as mean + SEM. Statistical significance was determined using one-way ANOVA and a post hoc Tukey’s test for pairwise comparison. **** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001, and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Genetic manipulation of PYCR2 expression modulates oncogenic properties of CRC cells. (<b>A</b>) Western blot analysis of PYCR2 expression in different CRC cell lines. The IEC-6 cells served as normal intestinal epithelial cells. (<b>Bi</b>,<b>Bii</b>) Immunoblot analysis of control and genetically manipulated PYCR2 HCT116 and SW480 cells and densitometric analysis (<span class="html-italic">p</span> = 0.0021 and <span class="html-italic">p</span> = 0.015). (<b>C</b>,<b>D</b>) Representative immunoblot analysis and densitometric analysis of EpCAM, E-cadherin, and vimentin in control and PYCR2-KO HCT116 cells (<span class="html-italic">p</span> = 0.00012 for EpCAM and <span class="html-italic">p</span> = 0.021 and 0.019 for E-cadherin and vimentin). (<b>E</b>) Immunofluorescence staining images for EpCAM expression in control and PYCR2-KO HCT116 cells. (<b>F</b>) Cell proliferation assays using the HCT116 control and PYCR2-KD cells (<span class="html-italic">p &lt;</span> 0.0001), (<b>Gi</b>,<b>Gii</b>) Soft agar assay using the HCT116 control and PYCR2-KD cells (<span class="html-italic">p</span> = 0.0257), (<b>Hi</b>,<b>Hii</b>) Cell migration assay using the HCT116 control and PYCR2-KD cells (<span class="html-italic">p</span> = 0.0355 at 48 h and <span class="html-italic">p</span> = 0.0048 at 72 h), (<b>Ii</b>,<b>Iii</b>) Cell invasion in HCT116 control and PYCR2-KD cells (<span class="html-italic">p</span> = 0.0017), and quantitative analysis. (<b>J</b>,<b>K</b>) Representative images of the immunoblot analysis of EpCAM, E-cadherin, and vimentin in control and PYCR2-overexpressing SW480 cells and densitometric evaluation (<span class="html-italic">p</span> = 0.031 for EpCAM and <span class="html-italic">p</span> = 0.0015 and 0.00029 for E-cadherin and vimentin). (<b>L</b>) Representative data for the effect of PYCR2 overexpression on cell proliferation (<span class="html-italic">p</span> = 0.00019). (<b>Mi</b>,<b>Mii</b>) Representative data for the cell invasion (<span class="html-italic">p</span> = 0.0024) in control and PYCR2-overexpressing SW480 cells. Data are presented as mean + SEM. Statistical significance was determined using Student’s <span class="html-italic">t</span>-test and one-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Inhibition of PYCR2 expression inhibits xenograft tumor growth and promotes apoptosis. (<b>A</b>) Schematics of in vivo studies using murine models of subcutaneous xenograft tumor growth and colonoscopy-guided cancer cell transplantation into the colon wall. (<b>Bi</b>) Representative images of the tumors isolated from athymic/nude mice injected with control or PYCR2-inhibited SW620 cells. (<b>Bii</b>,<b>Biii</b>) Statistical analysis showing % change in tumor volume (<span class="html-italic">p</span> = 0.0048) and fold change in tumor weight (<span class="html-italic">p</span> = 0.0019). (<b>Ci</b>) The analysis of the probability of mouse survival after colonoscopy-guided injection. (<b>Cii</b>–<b>Civ</b>) Representative images of the quantification of the % of tumor development; respective images of colon tumors and tumor size quantification (<span class="html-italic">p</span> = 0.0014, control vs. PYCR2 KD). (<b>Di</b>) Representative H&amp;E images of the tumors. (<b>Dii</b>–<b>Dv</b>) Representative images of IHC using anti-cleaved caspase-3 and p-H<sub>2</sub>AX antibodies in xenograft tumors and quantitative analysis (<span class="html-italic">p</span> = 0.0349 and <span class="html-italic">p</span> = 0.0018). (<b>E</b>) Immunoblot analysis for p-H<sub>2</sub>AX and cleaved PARP in HCT116 control and PYCR2-KD cells. (<b>F</b>,<b>G</b>) FACS analysis for early and late apoptosis in HCT116 control and PYCR2-KD cells and quantitative analysis (<span class="html-italic">p</span> = 0.0085). The representative figure has four quadrants where A = live cells, B = early apoptosis, C = late apoptosis, and D = necrosis. Data are presented as mean + SEM, and significance was determined using Student’s <span class="html-italic">t</span>-test and one-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>LC-MS/MS proteomics analysis to determine effects of PYCR2 loss of expression. (<b>A</b>) Schematics of the LC-MS/MS proteomics analysis. (<b>B</b>) Principal component analysis of the proteins differentially expressed in the control and PYCR2-KO HCT116 cells. (<b>C</b>,<b>D</b>) Analyses of the KEGG pathway and the GO biological function for differentially expressed proteins in PYCR2-KO versus control cells.</p>
Full article ">Figure 6
<p>PYCR2 regulates cell survival pathways and cancer stem cell population. (<b>A</b>) Heatmap analysis of proteins involved in cell apoptosis and proliferation. (<b>B</b>–<b>G</b>) Immunoblotting and densitometric analysis examining the expression of p-AKT and cyclin D1 in control and PYCR2-inhibited HCT116 and SW620 cells. (<b>H</b>–<b>J</b>) Immunoblotting and densitometric analysis examining the expression of p-AKT and cyclin D1 in control and PYCR2-overexpressing SW480 cells. (<b>K</b>) mRNA expression analysis for colonic CSC markers in HCT116 control cells and PYCR2-KD cells (<span class="html-italic">p &lt;</span> 0.011 for CD133, and <span class="html-italic">p</span> = 0.00014 for CD44 and 0.9484 for Sox2 (ns). (<b>L</b>) Representative immunoblots for the analysis of colonic CSC markers in control and PYCR2-KO HCT116 cells. (<b>Mi</b>,<b>Mii</b>) Sphere-forming assay using HCT116 control cells and PYCR2-KD cells and quantitative analysis (<span class="html-italic">p</span> = 0.0018). (<b>N</b>) mRNA expression analysis of colonic CSC markers in control and PYCR2-overexpressing SW480 cells (<span class="html-italic">p</span> &lt; 0.769 for CD133 (ns), 0.00156 for CD44, and <span class="html-italic">p &lt;</span> 0.0001 for Sox2). (<b>O</b>) Representative immunoblots for the analysis of colonic CSC markers in control and SW480-PYCR2 cells. (<b>Pi</b>–<b>Piii</b>) Sphere-forming assay using control and SW480-PYCR2 cells and quantitative analysis (<span class="html-italic">p</span> = 0.0158 for number of spheres, and <span class="html-italic">p</span> = 0.00013 for size). Data are presented as mean + SEM. Statistical significance was determined using Student’s <span class="html-italic">t</span>-test and one-way ANOVA. ns = non-significant, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>PYCR2 regulates the cell cycle and modulates MASTL/Wnt/β-catenin signaling. (<b>A</b>,<b>B</b>) Representative images of the cell cycle analysis using the control and PYCR2-KO HCT116 cells showing cell cycle arrest in PYCR2-KO cells at the G2/M phase and the subsequent quantification of the % of cells arrested at the G2/M phase. (<b>Ci</b>,<b>Cii</b>) Representative images of immunoblots and densitometric analysis examining the effects of PYCR2 on MASTL expression in PYCR2-KO HCT116. (<b>Di</b>,<b>Dii</b>) Immunoblots and densitometric analysis for MASTL expression in control and SW480-PYCR2 cells. (<b>Ei</b>,<b>Eii</b>) Representative images of immunoblots and densitometric analysis examining the effects of PYCR2 on Wnt signaling (p-β catenin s552) using the control and PYCR2-KO HCT116. (<b>F</b>) TOP-flash luciferase-based analysis of control and PYCR2-KO HCT116 cells. (<b>Gi</b>,<b>Gii</b>) Effect of PYCR2 overexpression on Wnt signaling (p-β catenin s552) in control and SW480-PYCR2 cells followed by densitometric analysis. (<b>H</b>) TOP-flash activity analysis of control and SW480-PYCR2 cells. Data are presented as mean + SEM. Statistical significance was determined using Student’s <span class="html-italic">t</span>-test and one-way ANOVA. <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>MASTL mediates CRC-promoting effects of PYCR2 expression. (<b>Ai</b>–<b>Aiii</b>) Immunoblot analysis determining the effects of MASTL overexpression in PYCR2-KO HCT116 cells and densitometric analysis of MASTL and p-βcatenin s552 expression in control, PYCR2 KO, and MASTL overexpression in PYCR2 KO HCT116 cells. (<b>Bi</b>–<b>Biii</b>) Immunoblot analysis determining the effect of the GKI-an inhibitor on MASTL expression/activity in PYCR2-overexpressing SW480 cells. A densitometric analysis of MASTL and p-βcatenin s552 expression in control, PYCR2 overexpression, and MASTL-inhibited SW480 cells is also presented. (<b>C</b>,<b>D</b>) Cell proliferation assay of HCT116-KD and SW480-PYCR2 cells after MASTL overexpression and inhibition, respectively. (<b>E</b>) Schematics summarizing our findings on the regulatory role of PYCR2 in CRC progression caused by modulating MASTL/Wnt/β-catenin signaling. Data are presented as mean + SEM. Statistical significance was determined using Student’s <span class="html-italic">t</span>-test and one-way ANOVA. ns = non-significant, * <span class="html-italic">p &lt;</span> 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
14 pages, 3484 KiB  
Article
Loss of Tumor Suppressor C9orf9 Promotes Metastasis in Colorectal Cancer
by Erfei Chen, Fangfang Yang, Qiqi Li, Tong Li, Danni Yao, Lichao Cao and Jin Yang
Biomolecules 2023, 13(2), 312; https://doi.org/10.3390/biom13020312 - 7 Feb 2023
Cited by 2 | Viewed by 1572
Abstract
The whole genome sequencing of tumor samples identifies thousands of somatic mutations. However, the function of these genes or mutations in regulating cancer progression remains unclear. We previously performed exome sequencing in patients with colorectal cancer, and identified one splicing mutation in C9orf9 [...] Read more.
The whole genome sequencing of tumor samples identifies thousands of somatic mutations. However, the function of these genes or mutations in regulating cancer progression remains unclear. We previously performed exome sequencing in patients with colorectal cancer, and identified one splicing mutation in C9orf9. The subsequent target sequencing of C9orf9 gene based on a validation cohort of 50 samples also found two function mutations, indicating that the loss of wild-type C9orf9 may participate in the tumorigenesis of colorectal cancer. In this research, we aimed to further confirm the function of C9orf9 in the CRC phenotype. Our Q-PCR analysis of the tumor and matched normal samples found that C9orf9 was downregulated in the CRC samples. Function assays revealed that C9orf9 exerts its tumor suppressor role mainly on cancer cell migration and invasion, and its loss was essential for certain tumor-microenvironment signals to induce EMT and metastasis in vivo. RNA-sequencing showed that stable-expressing C9orf9 can inhibit the expression of several metastasis-related genes and pathways, including vascular endothelial growth factor A (VEGFA), one of the essential endothelial cell mitogens which plays a critical role in normal physiological and tumor angiogenesis. Overall, our results showed that the loss of C9orf9 contributes to the malignant phenotype of CRC. C9orf9 may serve as a novel metastasis repressor for CRC. Full article
Show Figures

Figure 1

Figure 1
<p><b>Exome sequencing and deep sequencing identified three somatic mutations in <span class="html-italic">C9orf9</span> gene</b>. In-3: patient from initial exome sequencing; Ex-6, Ex-36, patients from extended validation cohort. Somatic mutations were visualized and analyzed by SnapGene v4.2.</p>
Full article ">Figure 2
<p><b>Gene expression analysis of C9orf9 in multiple samples of digestive tumor patients</b>. (<b>A</b>). Analysis of C9orf9 expression in five digestive tumors from TCGA and GTEx samples. * <span class="html-italic">p</span> &lt; 0.05. T: tumor (red), N: normal (grey). (<b>B</b>). C9orf9 expression level in tumor and matched normal tissues from TCGA and local validation cohort. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. (<b>C</b>). Correlation between copy number variation (CNV) and C9orf9 expression level in TCGA cohort. **** <span class="html-italic">p</span> &lt; 0.0001. (<b>D</b>). C9orf9 is downregulated in metastatic tumor samples of CRC, data from GSE41258. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p><b>C9orf9 has limit effect on cell growth.</b> (<b>A</b>). Overexpression of C9orf9 in SW480 and LoVo cells did not affect cell proliferation. (<b>B</b>). Knockdown of C9orf9 slightly promoted cell proliferation in LoVo cells, but not in SW480. (<b>C</b>). Cell apoptosis analysis using Annexin/PI double staining in LoVo and SW480 cells transfected with C9orf9 expression plasmid. (<b>D</b>). Cell apoptosis analysis using Annexin/PI double staining in LoVo and SW480 cells transfected with C9orf9-specific siRNAs. Ns, not significant, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p><b>C9orf9 regulates LoVo and SW480 cell migration and invasion capacity.</b> Transwell assays (Matrigel-free) or Transwell invasion assays (coated with Matrigel) were performed, respectively, in control and C9orf9-overexpression LoVo (<b>A</b>) and SW480 (<b>B</b>) cells. (<b>C</b>). Cell migration and invasion assays in control and C9orf9-knockdown LoVo cells. (<b>D</b>). Cell migration and invasion assays in control and C9orf9-knockdown SW480 cells. * <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 ">Figure 5
<p><b>C9orf9 inhibits tumor metastasis in vivo.</b> (<b>A</b>). Representative images of luciferase expression from lung metastasis of the normal saline (<span class="html-italic">n</span> = 3), LV-NC (<span class="html-italic">n</span> = 5), and LV-C9orf9 (<span class="html-italic">n</span> = 5) groups. (<b>B</b>). Quantification of the total flux, compared with the LV-NC group. (<b>C</b>). HE staining of lungs from the normal saline, LV-NC, and LV-C9orf9 groups. (<b>D</b>). mRNA level of metastasis-related marker N-cadherin and Vimentin. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p><b>C9orf9 responds to and involves in FGF, EGF, and IL6 induced EMT.</b> (<b>A</b>). Q-PCR analysis of C9orf9 and EMT-related markers (E-cadherin, N-cadherin, Vimentin) in FGF (20 ng/mL), EGF (20 ng/mL), and IL6 (50 ng/mL)-stimulated SW480 cells at 0, 12, and 24 h. (<b>B</b>) Transwell (Matrigel-free) assays of LV-NC and LV-C9orf9 SW480 cells with or without cytokine stimulation. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p><b>C9orf9 regulates metastasis-related gene and pathway.</b> (<b>A</b>). Volcano plot of differentially expressed genes (DEGs). (<b>B</b>). Top 20 KEGG pathway enrichment from DEGs. (<b>C</b>). GSEA plot of oxidative phosphorylation, EMT, and hypoxia gene set. (<b>D</b>). Q-PCR validation of angiogenesis, oxidative phosphorylation, and EMT related genes. * <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; 0.0001.</p>
Full article ">
20 pages, 9976 KiB  
Article
Prognostic Biomarker SPOCD1 and Its Correlation with Immune Infiltrates in Colorectal Cancer
by Lin Gan, Changjiang Yang, Long Zhao, Shan Wang, Zhidong Gao and Yingjiang Ye
Biomolecules 2023, 13(2), 209; https://doi.org/10.3390/biom13020209 - 20 Jan 2023
Cited by 2 | Viewed by 1829
Abstract
The biological role of the spen paralogue and orthologue C-terminal domain containing 1 (SPOCD1) has been investigated in human malignancies, but its function in colorectal cancer (CRC) is unclear. This study investigated the association between SPOCD1 expression and clinicopathological features of CRC cases, [...] Read more.
The biological role of the spen paralogue and orthologue C-terminal domain containing 1 (SPOCD1) has been investigated in human malignancies, but its function in colorectal cancer (CRC) is unclear. This study investigated the association between SPOCD1 expression and clinicopathological features of CRC cases, as well as its prognostic value and biological function based on large-scale databases and clinical samples. The results showed that the expression level of SPOCD1 was elevated in CRC, which was generally associated with shortened survival time and poor clinical indexes, including advanced T, N, and pathologic stages. Multivariate Cox regression analysis showed that elevated SPOCD1 expression was an independent factor for poor prognosis in CRC patients. Functional enrichment analysis of SPOCD1 and its co-expressed genes revealed that SPOCD1 could act as an oncogene by regulating gene expression in essential functions and pathways of tumorigenesis, such as extracellular matrix organization, chemokine signaling pathways, and calcium signaling pathways. In addition, immune cell infiltration results showed that SPOCD1 expression was associated with various immune cells, especially macrophages. Furthermore, our findings suggested a possible function for SPOCD1 in the polarization of macrophages from M1 to M2 in CRC. In conclusion, SPOCD1 is a promising diagnostic and prognostic marker for CRC, opening new avenues for research and treatment. Full article
Show Figures

Figure 1

Figure 1
<p>mRNA and protein expression levels of SPOCD1 in pan-cancer and colorectal cancer (CRC) versus normal samples. (<b>A</b>,<b>B</b>) SPOCD1 mRNA expression was upregulated in pan-cancer tissues compared with normal tissues based on TCGA. (<b>C</b>,<b>D</b>) SPOCD1 mRNA expression was upregulated in CRC compared with normal tissues based on TCGA. (<b>E</b>) ROC analysis of SPOCD1 in the diagnosis of CRC. (<b>F</b>–<b>H</b>) SPOCD1 mRNA expression was upregulated in CRC compared with normal tissues based on GEO. (<b>I</b>,<b>J</b>) IHC staining showing SPOCD1 protein expression was upregulated in CRC samples compared with normal tissues. (ns represents no significance, * <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 ">Figure 2
<p>Associations between SPOCD1 expression and clinicopathologic characteristics and its prognostic significance in CRC based on TCGA database. SPOCD1 expression was significantly associated with T stage (<b>A</b>) but not with N stage (<b>B</b>), M stage (<b>C</b>), and pathological stage (<b>D</b>). (<b>E</b>–<b>G</b>) The Kaplan-Meier curves show that CRC patients with a higher expression of SPOCD1 had a shorter overall survival time, disease-specific survival time, and progress-free interval time. (<b>H</b>) Multivariate Cox analyses of factors affecting the overall survival of CRC patients in the TCGA database. (<b>I</b>) immunohistochemistry (IHC) staining showing SPOCD1 protein expression in CRC samples of the different pathological stages. (<b>J</b>) Stacked histogram shows the proportion of CRC at different stages in the high and low SPOCD1 expression groups. (ns represents no significance, * <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 ">Figure 3
<p>Co-expression gene analysis and subsequent GO and KEGG enrichment analysis of SPOCD1. (<b>A</b>) Heatmap illustrating the top 50 genes positively correlated with SPOCD1 in CRC. Red represents high expression, and blue represents low expression. (<b>B</b>) Correlation analysis of SPOCD1 and SCL11A1. (<b>C</b>) The Kaplan-Meier plots for overall survival for CRC patients according to the difference SLC11A1 expressions. (<b>D</b>) GO and KEGG pathway analysis of SPOCD1 co-expressed gene with a correlation coefficient &gt; 0.6 in CRC. (<b>E</b>–<b>G</b>) GSEA enrichment plots showed positive correlations of SPOCD1 co-expression genes with the apical junction, calcium, and chemokine signaling pathway (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>The correlation of SPOCD1 expression with tumor microenvironment and immune infiltration level in CRC. (<b>A</b>–<b>C</b>) The correlation between SPOCD1 expression and the ESTIMATE, stromal, and immune scores based on the ESTIMATE algorithm. (<b>D</b>) The lollipop plot shows the correlation of SPOCD1 expression with immune cell infiltration conducted by ssGSEA. (<b>E</b>) Enrichment score of macrophages according to the difference SPOCD1 expressions. (<b>F</b>–<b>K</b>) The correlation of SPOCD1 expression with immune cell infiltration in CRC acquired from the TIMER online tool. (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Correlation between SPOCD1 expression level and macrophage markers in CRC. (<b>A</b>–<b>C</b>) Correlation between SPOCD1 expression and macrophage markers (CD68, CD163, and MRC1). (<b>D</b>) CD163 is upregulated in SPOCD1 high-expression group. (<b>E</b>) Expression of SPOCD1 and M2 macrophage marker CD163 in CRC tissues (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Correlation between SPOCD1 expression level and immunological microenvironment biomarkers in CRC. (<b>A</b>) Correlation between SPOCD1 and immunostimulatory gene expression in CRC. (<b>B</b>) Correlation between SPOCD1 and MHC molecule expression in CRC. (<b>C</b>) Correlation between SPOCD1 and immunoinhibitory gene expression in CRC. (<b>D</b>) Correlation between SPOCD1 expression and TGFB1 in CRC (ns represents no significance, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 7
<p>Correlation analysis between SPOCD1 expression and chemokines and chemokine receptors in CRC. (<b>A</b>) Heatmap analysis of the correlation between SPOCD1 and chemokine receptors in tumors. (<b>B</b>) Heatmap analysis of the correlation between SPOCD1 and chemokines in tumors. (<b>C</b>) Correlation between SPOCD1 expression and chemokines CCL18, chemokine receptors CCR1 in CRC (ns represents no significance, * <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 ">Figure 8
<p>Gene-drug sensitivity analysis based on the CellMiner database: the top 12 drugs with high correlation with SPOCD1 expression in CRC were screened.</p>
Full article ">
15 pages, 2356 KiB  
Review
Effects of Long Non-Coding RNAs Induced by the Gut Microbiome on Regulating the Development of Colorectal Cancer
by Shiying Fan, Juan Xing, Zhengting Jiang, Zhilin Zhang, Huan Zhang, Daorong Wang and Dong Tang
Cancers 2022, 14(23), 5813; https://doi.org/10.3390/cancers14235813 - 25 Nov 2022
Cited by 8 | Viewed by 2128
Abstract
Although an imbalanced gut microbiome is closely associated with colorectal cancer (CRC), how the gut microbiome affects CRC is not known. Long non-coding RNAs (lncRNAs) can affect important cellular functions such as cell division, proliferation, and apoptosis. The abnormal expression of lncRNAs can [...] Read more.
Although an imbalanced gut microbiome is closely associated with colorectal cancer (CRC), how the gut microbiome affects CRC is not known. Long non-coding RNAs (lncRNAs) can affect important cellular functions such as cell division, proliferation, and apoptosis. The abnormal expression of lncRNAs can promote CRC cell growth, proliferation, and metastasis, mediating the effects of the gut microbiome on CRC. Generally, the gut microbiome regulates the lncRNAs expression, which subsequently impacts the host transcriptome to change the expression of downstream target molecules, ultimately resulting in the development and progression of CRC. We focused on the important role of the microbiome in CRC and their effects on CRC-related lncRNAs. We also reviewed the impact of the two main pathogenic bacteria, Fusobacterium nucleatum and enterotoxigenic Bacteroides fragilis, and metabolites of the gut microbiome, butyrate, and lipopolysaccharide, on lncRNAs. Finally, available therapies that target the gut microbiome and lncRNAs to prevent and treat CRC were proposed. Full article
Show Figures

Figure 1

Figure 1
<p>Biogenesis and the mode of action of lncRNAs. (<b>A</b>) Signal: some lncRNAs are specifically transcribed under different TFs and signaling pathways, and they subsequently participate in specific signaling pathways as signaling molecules. (<b>B</b>) Decoy: a certain type of lncRNAs are transcribed and bind directly to protein targets but are inactive, thus inhibiting the function of the molecule and the signaling pathway. (<b>C</b>) Guidance: certain lncRNAs, when bound to proteins, can direct the localization of ribonucleoprotein complex to specific targets (colored flags), thereby regulating the transcription of downstream molecules. (<b>D</b>) Scaffolding: some lncRNAs act as a “central platform” to bring multiple proteins together to form ribonucleoprotein complexes, enabling the intersection and integration of information between various signaling pathways and facilitating the rapid feedback and regulation of the body or cell to external signals and stimuli. (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>). lncRNAs: long non-coding RNAs; TFs: transcription factors.</p>
Full article ">Figure 2
<p><span class="html-italic">F. nucleatum</span> promotes CRC metastasis as well as enhances drug resistance by upregulating <span class="html-italic">EVADR</span>, KRT7-AS, and ENO1-IT1. Elevated <span class="html-italic">EVADR</span> directs RBP YBX1 to recruit EMT-related factors, including Snail, Slug, and Zeb1, to polyribosomes, and enhances the translation of these EMT-related factors, thereby inducing EMT, ultimately facilitating colorectal cancer metastasis. <span class="html-italic">F. nucleatum</span> infection activates the NF-κB pathway, and activated NF-κB <span class="html-italic">P</span>-p65 (activated NF-κB subunit) may upregulate KRT7-AS by increasing the transcriptional activity of KRT7-AS, then activating the downstream target of KRT7-AS, KRT7, which regulates CRC metastasis. <span class="html-italic">F. nucleatum</span> upregulates the binding efficiency of the transcription factor SP1 to the promoter region of lncRNA ENO1-IT1 to activate the transcription of lncRNA ENO1- IT1 and subsequently recruit KAT7 to the promoter of the ENO1 gene to regulate ENO1 transcription via epigenetic modulation, thereby activating CRC glycolysis, which enhances the drug resistance. (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>). <span class="html-italic">F. nucleatum</span>: <span class="html-italic">Fusobacterium nucleatum</span>; EVADR: endogenous retroviral-associated adenocarcinoma lncRNA; KRT7-AS: Keratin7-antisense; ENO1-IT1: enolase1-intronic transcript 1; RBP: RNA-binding protein; KRT7: Keratin7; EMT: epithelial mesenchymal transition; lncRNA: long non-coding RNA.</p>
Full article ">Figure 3
<p>ETBF induces CRC cells growth, proliferation, and metastasis by regulating BFAL1 and AERRIE. BFAL1 competes with miR-155-5p and miR-200a-3p, thus impeding the inhibitory effect of miR-155-5p and miR-200a-3p on RHEB, which upregulates the target RHEB mRNA expression. RHEB can bind directly to the mTOR complex and regulate the mTOR-signaling pathway by phosphorylating the p70 S6K, leading to the activation of the mTOR pathway, thus promoting ETBF-induced CRC cells growth. ETBF can also increase the levels of JMJD2B, and then induce the expression of lncRNA AERRIE, which in turn induces the expression of SULF1. Elevated SULF1 can activate the canonical Wnt pathway, which has been demonstrated to promote the proliferation and metastasis of CRC cells. (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>). ETBF: Enterotoxigenic <span class="html-italic">Bacteroides fragilis</span>; BFAL1: <span class="html-italic">B. fragilis</span>-associated lncRNA1; RHEB: Ras homolog enriched in brain; S6K: S6 Kinase; JMJD2B: Jumonji domain-containing protein 2B; lncRNA: long non-coding RNA; SULF1: sulfatase 1.</p>
Full article ">Figure 4
<p>Butyrate as the metabolite of the gut microbiome inhibits the intestinal inflammation to lower the risk of CRC by regulating lncRNA LncLy6C. Butyrate-induced LncLy6C binds to the C/EBPβ and H3K4me3, specifically encouraging the enrichment of C/EBPβ and H3K4me3 marks on the promoter region of Nr4A1, thus promoting the expression of Nr4A1. This promotes the differentiation of Ly6C<sup>high</sup> inflammatory monocytes into Ly6C<sup>int/neg</sup> resident macrophages, leading to the inhibition of inflammation. As a result, the risk of developing CAC goes down. (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>). lncRNA: long non-coding RNA; C/EBPβ: CCAAT/enhancer binding protein β; CAC: colitis-associated colorectal cancer.</p>
Full article ">Figure 5
<p>The gut microbiome-derived LPS promotes cancer cells migration and invasion by regulating LINC00152. The bacteria-derived LPS introduces histone lactonization on the promoter of LINC00152 and reduces the binding efficiency of YY1 to LINC00152, thus upregulating the expression of LINC00152. The LPS-induced overexpression of LINC00152 was associated with cancer cells migration and invasion. (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>). LPS: lipopolysaccharide.</p>
Full article ">
23 pages, 1144 KiB  
Review
Short Linear Motifs in Colorectal Cancer Interactome and Tumorigenesis
by Candida Fasano, Valentina Grossi, Giovanna Forte and Cristiano Simone
Cells 2022, 11(23), 3739; https://doi.org/10.3390/cells11233739 - 23 Nov 2022
Cited by 4 | Viewed by 2303
Abstract
Colorectal tumorigenesis is driven by alterations in genes and proteins responsible for cancer initiation, progression, and invasion. This multistage process is based on a dense network of protein–protein interactions (PPIs) that become dysregulated as a result of changes in various cell signaling effectors. [...] Read more.
Colorectal tumorigenesis is driven by alterations in genes and proteins responsible for cancer initiation, progression, and invasion. This multistage process is based on a dense network of protein–protein interactions (PPIs) that become dysregulated as a result of changes in various cell signaling effectors. PPIs in signaling and regulatory networks are known to be mediated by short linear motifs (SLiMs), which are conserved contiguous regions of 3–10 amino acids within interacting protein domains. SLiMs are the minimum sequences required for modulating cellular PPI networks. Thus, several in silico approaches have been developed to predict and analyze SLiM-mediated PPIs. In this review, we focus on emerging evidence supporting a crucial role for SLiMs in driver pathways that are disrupted in colorectal cancer (CRC) tumorigenesis and related PPI network alterations. As a result, SLiMs, along with short peptides, are attracting the interest of researchers to devise small molecules amenable to be used as novel anti-CRC targeted therapies. Overall, the characterization of SLiMs mediating crucial PPIs in CRC may foster the development of more specific combined pharmacological approaches. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of a novel in silico methodology developed by our group to search for new interactors of an oncoprotein of interest by taking advantage of a library of SLiMs (tripeptides P1–P19) able to bind it in vitro, as identified by surface plasmon resonance (SPR) analysis [<a href="#B124-cells-11-03739" class="html-bibr">124</a>]. Tripeptides P1–P19 are then used as in silico probes to identify human proteins containing them, which are therefore candidate interactors of the oncoprotein of interest.</p>
Full article ">
23 pages, 2249 KiB  
Review
The Role of Fusobacterium nucleatum in Colorectal Cancer Cell Proliferation and Migration
by Zihong Wu, Qiong Ma, Ying Guo and Fengming You
Cancers 2022, 14(21), 5350; https://doi.org/10.3390/cancers14215350 - 30 Oct 2022
Cited by 10 | Viewed by 3325
Abstract
Colorectal cancer (CRC) is a common cancer worldwide with poor prognosis. The presence of Fusobacterium nucleatum (Fn) in the intestinal mucosa is associated with the progression of CRC. In this review, we explore the mechanisms by which Fn contributes to proliferation and migration [...] Read more.
Colorectal cancer (CRC) is a common cancer worldwide with poor prognosis. The presence of Fusobacterium nucleatum (Fn) in the intestinal mucosa is associated with the progression of CRC. In this review, we explore the mechanisms by which Fn contributes to proliferation and migration of CRC cells from the following four aspects: induction of the epithelial–mesenchymal transition (EMT), regulation of the tumor microenvironment (TME), expression of oncogenic noncoding RNAs, and DNA damage. This review outlines the scientific basis for the use of Fn as a biomarker and therapeutic target in CRC. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>An overview of the EMT process during which Fn is involved in the development of CRC.</p>
Full article ">Figure 2
<p>Fn reprograms the TME by recruiting tumor-infiltrating immune cells and prometastatic cytokines.</p>
Full article ">Figure 3
<p>A schematic of the effect of FadA in disrupting adherence junctions in CRC cells. In adherence junctions, E-cadherin and β-catenin form complexes on epithelial cell membranes and bind to the actin cytoskeleton to maintain cell–cell adhesion. When FadA invades, it can bind to E-cadherin, resulting in its separation from β-catenin and subsequent activation of β-catenin signaling, leading to a reduction in cell–cell adherence between cancer cells.</p>
Full article ">Figure 4
<p>A schematic of oncogene expression regulated by Fn in CRC metastasis.</p>
Full article ">Figure 5
<p>An overview of the molecular mechanisms by which Fn is associated with CRC proliferation and migration.</p>
Full article ">
17 pages, 1228 KiB  
Review
Zinc Finger Proteins: Functions and Mechanisms in Colon Cancer
by Shujie Liu, Xiaonan Sima, Xingzhu Liu and Hongping Chen
Cancers 2022, 14(21), 5242; https://doi.org/10.3390/cancers14215242 - 26 Oct 2022
Cited by 10 | Viewed by 3600
Abstract
According to the global cancer burden data for 2020 issued by the World Health Organization (WHO), colorectal cancer has risen to be the third-most frequent cancer globally after breast and lung cancer. Despite advances in surgical treatment and chemoradiotherapy for colon cancer, individuals [...] Read more.
According to the global cancer burden data for 2020 issued by the World Health Organization (WHO), colorectal cancer has risen to be the third-most frequent cancer globally after breast and lung cancer. Despite advances in surgical treatment and chemoradiotherapy for colon cancer, individuals with extensive liver metastases still have depressing prognoses. Numerous studies suggest ZFPs are crucial to the development of colon cancer. The ZFP family is encoded by more than 2% of the human genome sequence and is the largest transcriptional family, all with finger-like structural domains that could combine with Zn2+. In this review, we summarize the functions, molecular mechanisms and recent advances of ZFPs in colon cancer. We also discuss how these proteins control the development and progression of colon cancer by regulating cell proliferation, EMT, invasion and metastasis, inflammation, apoptosis, the cell cycle, drug resistance, cancer stem cells and DNA methylation. Additionally, several investigations have demonstrated that Myeloid zinc finger 1 (MZF1) has dual functions in colon cancer, which may both promote cancer proliferation and inhibit cancer progression through apoptosis. Generally, a comprehensive understanding of the action mechanisms of ZFPs in colon cancer will not only shed light on the discovery of new diagnostic and prognosis indicators but will also facilitate the design of novel targeted therapies. Full article
Show Figures

Figure 1

Figure 1
<p>The function of ZFPs in regulating the cellular biological processes of colon cancer. ZFPs play important roles in the regulation of cell proliferation, epithelial–mesenchymal transition (EMT), invasion and metastasis, inflammation, cell cycle, cancer stem cells and DNA methylation in colon cancer cells. (This figure was created with <a href="http://biorender.com" target="_blank">biorender.com</a>).</p>
Full article ">Figure 2
<p>The possible mechanisms of ZFPs in regulating the cell cycle of colon cancer. There are various zinc finger proteins involved in cell cycle processes, such as ZFP91, ZFP278, ZFP692, Slug, KLF6-SV2, and P52-ZER6. The underlying mechanisms involve cyclin D, cyclin E, E2F, p21, p53, Bax, and MDM2. The underlying mechanism affects multiple molecules, such as cyclin D, cyclin E, E2F, p21, p53, Bax, and MDM2, eventually inducing cell progression or inhibiting cell proliferation in colon cancer. These ZFPs have great potential as novel therapeutic targets for colon cancer. (This figure was created with <a href="http://biorender.com" target="_blank">biorender.com</a>).</p>
Full article ">Figure 3
<p>In colon cancer, MZF1 plays dual and opposite roles in different signaling pathways: (<b>a</b>) MZF1 transcriptionally activates the downstream target gene Axl and stimulates various signaling pathways, such as PI3K, FAK, Grb2/Ras, MEK/ERK, advancing cell proliferation, EMT transformation, invasion, and metastasis in colon cancer. (<b>b</b>) MZF1 transcriptionally activates the downstream target gene p55PIK; stimulates diverse signaling pathways such as PI3K/Akt and PI3K/RAC; and activates a series of downstream target genes such as CDC2, ALDH, BCL2, TWIST, SNAIL, and SLUG, promoting cell proliferation, EMT transformation, invasion, and metastasis in colon cancer. (<b>c</b>) MZF1 transcriptionally triggers the downstream target gene c-myc and multiple downstream target genes, such as MINA53, ID2, BCL2, and PTMA, promoting proliferation in colon cancer. (<b>d</b>) Sulfide sulindac sulfide induces the upregulation of MZF1. MZF1 promotes the expression of DR5 that interacts with FADD to activate caspases, promoting cell apoptosis and eventually inhibiting metastasis in colon cancer. (This figure was created with <a href="http://biorender.com" target="_blank">biorender.com</a>).</p>
Full article ">
18 pages, 4132 KiB  
Article
CREPT Disarms the Inhibitory Activity of HDAC1 on Oncogene Expression to Promote Tumorigenesis
by Yajun Cao, Bobin Ning, Ye Tian, Tingwei Lan, Yunxiang Chu, Fangli Ren, Yinyin Wang, Qingyu Meng, Jun Li, Baoqing Jia and Zhijie Chang
Cancers 2022, 14(19), 4797; https://doi.org/10.3390/cancers14194797 - 30 Sep 2022
Cited by 2 | Viewed by 1863
Abstract
Histone deacetylases 1 (HDAC1), an enzyme that functions to remove acetyl molecules from ε-NH3 groups of lysine in histones, eliminates the histone acetylation at the promoter regions of tumor suppressor genes to block their expression during tumorigenesis. However, it remains unclear why HDAC1 [...] Read more.
Histone deacetylases 1 (HDAC1), an enzyme that functions to remove acetyl molecules from ε-NH3 groups of lysine in histones, eliminates the histone acetylation at the promoter regions of tumor suppressor genes to block their expression during tumorigenesis. However, it remains unclear why HDAC1 fails to impair oncogene expression. Here we report that HDAC1 is unable to occupy at the promoters of oncogenes but maintains its occupancy with the tumor suppressors due to its interaction with CREPT (cell cycle-related and expression-elevated protein in tumor, also named RPRD1B), an oncoprotein highly expressed in tumors. We observed that CREPT competed with HDAC1 for binding to oncogene (such as CCND1, CLDN1, VEGFA, PPARD and BMP4) promoters but not the tumor suppressor gene (such as p21 and p27) promoters by a chromatin immunoprecipitation (ChIP) qPCR experiment. Using immunoprecipitation experiments, we deciphered that CREPT specifically occupied at the oncogene promoter via TCF4, a transcription factor activated by Wnt signaling. In addition, we performed a real-time quantitative PCR (qRT-PCR) analysis on cells that stably over-expressed CREPT and/or HDAC1, and we propose that HDAC1 inhibits CREPT to activate oncogene expression under Wnt signaling activation. Our findings revealed that HDAC1 functions differentially on tumor suppressors and oncogenes due to its interaction with the oncoprotein CREPT. Full article
Show Figures

Figure 1

Figure 1
<p>The correlation of HDAC1 and CREPT in human tumors. (<b>A</b>). CREPT expression is positively correlated to HDAC1 at the protein level in colon cancer cells. The endogenous protein levels of HDAC1 and CREPT in colon cancer cell lines (SW620, SW480, HCT116, DLD1) and epithelial cells (NCM460) are shown. β-actin was used as a loading control. (<b>B</b>,<b>C</b>). Both HDAC1 (<b>B</b>) and CREPT (<b>C</b>) are highly expressed in colon cancer cells when compared with the normal cells at the mRNA level. The mRNA levels were normalized to the fold change relative to β-actin. (<b>D</b>). Both HDAC1 and CREPT are elevated in the tumor tissues compared with the adjacent normal tissues from human colon cancers. Western blot analyses were based on β-actin as a loading control. N represents paired adjacent normal tissue and T represents tumor tissue from the same patient. The bands were quantified using ImageJ and presented as a value normalized according to the moderate level of a band in each blot. (<b>E</b>). The relative level of CREPT and HDAC1 in adjacent normal and tumor tissues are shown. The ratio of CREPT or HDAC1 vs actin was calculated by quantifying the bands from the Western blot in (<b>D</b>). The closed circle represents normal tissue for CREPT/actin, the square represents tumor tissue for CREPT/actin, the triangle represents normal tissue for HDAC1/actin and the inverted triangle represents tumor tissue for HDAC1/actin. (<b>F</b>). HDAC1 and CREPT were co-stained in human colon tumor tissue. Immunohistochemical staining was performed with an antibody against CREPT or HDAC1. (<b>G</b>). A graphical presentation of correlation of CREPT and HDAC1 in colon cancer. The levels of CREPT and HDAC1 in tumor tissue and the adjacent normal tissue were quantified using Image J. The ratio of CREPT (<span class="html-italic">X</span>-axis) or HDAC1 (<span class="html-italic">Y</span>-axis) level in tumor tissue to the adjacent normal tissue was calculated. SPSS was employed to figure out the correlation coefficient between the two proteins. (* <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 ">Figure 2
<p>HDAC1 represses the activity of CREPT in promoting tumorigenesis. Stable cell lines were established for the overexpression of HDAC1 (HDAC1 OE) alone, CREPT (CREPT OE) alone, and both HDAC1 and CREPT (HDAC1&amp;CREPT OE) in DLD1 cells, as well as for the deletion of HDAC1 (siHDAC1) or deletion of CREPT (CREPT KO) in SW480 cells. An empty vector was used as a control (EV) for overexpression experiments in DLD1 cells and a non-specific siRNA (siNC) was used as a control for deletion experiments in SW480 cells. Indicated stable cell lines were cultured for the CCK8 experiment at different days. A colony formation experiment was performed. A statistic result for the colony numbers as scanned by Image J is presented on the right panel. (<b>A</b>–<b>D</b>). Ectopic expression of HDAC1 significantly inhibits cell proliferation (<b>A</b>,<b>B</b>) or colony formation (<b>C</b>,<b>D</b>) in the presence of overexpressed CREPT in DLD1 cells. (<b>E</b>–<b>H</b>). Knocking down the expression of HDAC1 elevates cell proliferation (<b>E</b>,<b>F</b>) or colony formation (<b>G</b>,<b>H</b>) in the absence of CREPT in SW480 cells. The bands were quantified using ImageJ and presented as a value normalized according to the moderate level of a band in each blot. (<b>I</b>–<b>K</b>). Overexpression of HDAC1 inhibits the tumor formation in the presence of overexpressed CREPT by Lovo cells in nude mice. (<b>L</b>). HDAC1 blocks the activity of CREPT in regulating the Wnt signaling pathway. Super-TOP-luciferase reporter (0.1 μg), pRL-TK (50 ng), HA–HDACs (0.1 μg) and Myc–CREPT (0.4 μg) were co-transfected into HEK293T cells. Luciferase activity is presented as a relative value based on the internal control (renilla signal). (<b>M</b>). HDAC1 inhibits the expression of CREPT-activated genes. The mRNA levels of CCND1, CLDN1, VEGFA, BMP4, PPARD, p21 and p27 were examined in DLD1 cell lines. (ns, not significantly, * <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 ">Figure 3
<p>HDAC1 specifically interacts with CREPT. (<b>A</b>). Flag–HDAC1 interacts with Myc–CREPT. Myc–CREPT (3 μg) and Flag–HDAC1/2/3 (3 μg) were co-transfected into HEK293T cells for immunoprecipitation experiments. Antibodies used are indicated. (<b>B</b>,<b>C</b>). HDAC1 interacts with CREPT endogenously. Nuclear extracts were immunoprecipitated with an antibody against HDAC1/2/3 (<b>B</b>) or with an antibody against CREPT (<b>C</b>) in HEK293T and DLD1 cells. (<b>D</b>). HDAC1 and CREPT are co-localized in the nucleus. HEK293T cells expressing GFP-CREPT and Flag–HDAC1 were scanned by a confocal microscope after staining. Scale bar, 10 um. HDAC1 interacts with the CID domain of CREPT in HEK293T. Anti-Myc and anti-Flag antibodies were used. (<b>F</b>). The N-terminus domain of HDAC1 interacts with CREPT in HEK293T cells. Immunoprecipitation was performed (<b>E</b>).</p>
Full article ">Figure 4
<p>HDAC1 and CREPT exclusively interact with TCF4 and β-catenin. (<b>A</b>). HDAC1 and CREPT co-localize with TCF4 in the nucleus. HEK293T cells expressing HA–TCF4 together with GFP–CREPT or Flag–HDAC1 were scanned by confocal microscope. Scale bar, 10 um. (<b>B</b>). The interaction of HDAC1 with TCF4 is shown. Nuclear extracts from HEK293T cells were used. (<b>C</b>). CREPT associates with TCF4. Immunoprecipitation was performed with indicated antibodies. (<b>D</b>). TCF4 interacts with CREPT and HDAC1 exclusively. HEK293T cells were co-transfected with Flag–HDAC1 (10 μg), HA–TCF4 (6 μg) and Myc–CREPT (4 μg) for IP experiments using an antibody against Flag. Precipitants were used for re-IP experiments using an antibody against Myc or an antibody against HA. (<b>E</b>). CREPT impairs the TCF4–HDAC1 complex formation. HEK293T cells were co-transfected with Flag–HDAC1 (2 μg) and HA–TCF4 (2 μg) under stable expression (Myc–CREPT) or deletion (CREPT KO) of CREPT for IP experiments using an antibody against HA. (<b>F</b>). CREPT impedes the β-catenin–HDAC1 complex formation. HEK293T cells were co-transfected with HA–HDAC1 (2 μg) and Flag–β-catenin (2 μg) under stable expression (Myc–CREPT) or deletion (CREPT KO) of CREPT for IP experiments using an antibody against Flag. (<b>G</b>). CREPT represses the interaction between HDAC1 and β-catenin–TCF4 complex endogenously. HEK293T cells under stable expression (Myc–CREPT) or deletion (CREPT KO) of CREPT were collected for IP experiments using an antibody against HDAC1. (<b>H</b>,<b>I</b>). HDAC1 suppresses the CREPT–TCF4 complex (<b>H</b>) and CREPT–β-catenin complex (<b>I</b>) formation. HEK293T cells were co-transfected with Myc–CREPT (2 μg) and HA–TCF4 (2 μg) (<b>H</b>) or Flag–β-catenin (2 μg) (<b>I</b>) under overexpression (Flag–HDAC1) or deletion (siHDAC1) of HDAC1 for IP experiments using an antibody against Myc. (<b>J</b>). HDAC1 represses the interaction between CREPT and β-catenin–TCF4 complex endogenously. HEK293T cells under stable expression (Flag–HDAC1) or deletion (siHDAC1-1/2) of HDAC1 were collected for IP experiments using an antibody against CREPT. (<b>K</b>). CREPT facilitates the β-catenin–TCF4 complex formation. Cells were co-transfected with Flag–β-catenin (3 μg) and HA–TCF4 (3 μg) under stable expression (Myc–CREPT) or deletion (CREPT KO) of CREPT for IP experiments using an antibody against HA. The Myc–CREPT in the lysates was detected with anti-Myc antibody in the left panels of (<b>E</b>,<b>F</b>,<b>K</b>). The endogenous CREPT was detected with anti-CREPT antibody in the right panels of (<b>E</b>,<b>F</b>,<b>K</b>,<b>L</b>). HDAC1 impairs the β-catenin–TCF4 complex formation. HEK293T cells were co-transfected with Flag–β-catenin (3 μg) and HA–TCF4 (3 μg) under stable expression (Flag–HDAC1) or deletion (siHDAC1-1/2) of HDAC1 for IP experiments using an antibody against HA. The Flag–HDAC1 in the lysates was detected with anti-Flag antibody in the left panels of (<b>H</b>,<b>I</b>,<b>L</b>). The endogenous HDAC1 was detected with anti- HDAC1 antibody in the right panels of (<b>H</b>,<b>I</b>,<b>L</b>). (<b>M</b>). CREPT and HDAC1 competitively influence the interaction of β-catenin and TCF4. Flag–β-catenin (2 μg), HA–TCF4 (2 μg), Myc–CREPT (2 μg) (6th column) or (4 μg) (5th column), Flag–HDAC1 (2 μg) (5th column) or (4 μg) (6th column) were co-transfected in cells for IP experiment by using an antibody against HA.</p>
Full article ">Figure 5
<p>CREPT blocks the occupancy of HDAC1 at the TBS region of target genes. (<b>A</b>,<b>B</b>). HDAC1 or CREPT occupies the TCF4 binding sequence (TBS) of Wnt target oncogenes. DLD1 cells overexpressing HDAC1 (<b>A</b>) or CREPT (<b>B</b>) were collected for the ChIP-qPCR assay. (<b>C</b>). CREPT represses the binding of HDAC1 on the TBS region of Wnt target oncogenes. DLD1 cells under stable expression of HA–HDAC1 and Flag–CREPT were collected for ChIP-qPCR assay. (<b>D</b>). Deletion of CREPT facilitates the occupancy of HDAC1 on the TBS region. DLD1 cells under deletion of CREPT were cultured in the presence of over-expressed HA–HDAC1 and harvested for ChIP-qPCR assay using an antibody against HA. (<b>E</b>). HDAC1 impairs the binding of CREPT on the TBS region of Wnt target oncogenes. DLD1 cells under stable expression of HA–HDAC1 and Flag–CREPT were collected for ChIP-qPCR assay. (<b>F</b>). Deletion of HDAC1 promotes the occupancy of CREPT at the TBS region. DLD1 cells under deletion of HDAC1 were cultured in the presence of over-expressed Flag–CREPT and harvested for ChIP-qPCR assay using an antibody against Flag. The occupancy abundance was revealed by the precipitated DNA fragments on the TBS regions using qPCR. Wnt target oncogenes CCND1, CLDN1, VEGFA, PPARD, and BMP4 were examined. Tumor suppressor genes p21 and p27 were used as a control. (<b>G</b>,<b>H</b>). CREPT regulates the histone acetylation at H3K27 in the promoters of CCND1 (<b>G</b>) or CLDN1 (<b>H</b>) depending on HDAC1. (<b>I</b>,<b>J</b>). CREPT failed to influence the acetylation levels of tumor suppressor genes p21 (<b>I</b>) or p27 (<b>J</b>). Cells under stable expression (HA–CREPT) or deletion (CREPT KO) of CREPT were transfected with siRNAs against HDAC1 (si-HDAC1-1, siHDAC1-2). ChIP-qPCR was performed using an antibody against AcH3K27. (ns, not significantly, * <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 ">Figure 6
<p>Wnt3a induces CREPT but reduces HDAC1 binding TBS. (<b>A</b>,<b>B</b>). Wnt3a increases the interaction between CREPT and TCF4 (<b>A</b>)-β-catenin (<b>B</b>) complex. Myc–CREPT (3 μg), HA–TCF4 (3 μg) (<b>A</b>) or Flag–β-catenin (3 μg) (<b>B</b>) were co-transfected in HEK293T cells. Then, Wnt3a was added for 20 h. (<b>C</b>,<b>D</b>). Wnt3a represses the interaction between HDAC1 and TCF4 (<b>C</b>)/β-catenin (<b>D</b>) complex. Flag–HDAC1 (3 μg), HA–TCF4 (3 μg) (<b>C</b>) or Flag–β-catenin (3 μg) (<b>D</b>) were co-transfected cells with or without Wnt3a for 20 h. (<b>E</b>). Wnt3a decreases the interaction between CREPT and HDAC1. HEK293T cells were transfected with Myc–CREPT (3 μg) and Flag–HDAC1 (3 μg) in the presence or absence of Wnt3a. (<b>F</b>). Wnt3a increases the interaction of CREPT with TCF4/β-catenin but reduces the CREPT–HDAC1 interaction. (<b>G</b>). Wnt3a increases the binding of CREPT on TBS regions. (<b>H</b>). Wnt3a reduces the occupancy of HDAC1 on the TBS regions. HEK293T cells were harvested with or without Wnt3a for 24 h. Then cells were collected for a ChIP assay using indicated antibodies. The ChIPed DNA was examined by qPCR. (<b>I</b>). A model demonstrating the competition of CREPT with HDAC1 in regulating gene expression during tumorigenesis. In tumor cells, HDAC1 interacts with transcription factors and maintains histone proteins at the promoters of tumor suppressor genes (TSG) in a deacetylation state (right panel), leading to repressed TSG transcription. However, highly expressed CREPT is recruited to the promoter of Wnt target oncogenes. Then, CREPT interacts and dissociates HDAC1 from TCF4. The dissociation of HDAC1 recovers the promoter in an acetylated state. The subsequent binding of CREPT to TCF4/β-catenin initiates oncogene transcription and promotes tumorigenesis. (ns, not significantly, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
14 pages, 74690 KiB  
Article
Nuclear Beclin 1 Destabilizes Retinoblastoma Protein to Promote Cell Cycle Progression and Colorectal Cancer Growth
by Yang Pan, Zhiqiang Zhao, Juan Li, Jinsong Li, Yue Luo, Weiyuxin Li, Wanbang You, Yujun Zhang, Zhonghan Li, Jian Yang, Zhi-Xiong Jim Xiao and Yang Wang
Cancers 2022, 14(19), 4735; https://doi.org/10.3390/cancers14194735 - 28 Sep 2022
Cited by 5 | Viewed by 2442
Abstract
Autophagy is elevated in colorectal cancer (CRC) and is generally associated with poor prognosis. However, the role of autophagy core-protein Beclin 1 remains controversial in CRC development. Here, we show that the expression of nuclear Beclin 1 protein is upregulated in CRC with [...] Read more.
Autophagy is elevated in colorectal cancer (CRC) and is generally associated with poor prognosis. However, the role of autophagy core-protein Beclin 1 remains controversial in CRC development. Here, we show that the expression of nuclear Beclin 1 protein is upregulated in CRC with a negative correlation to retinoblastoma (RB) protein expression. Silencing of BECN1 upregulates RB resulting in cell cycle G1 arrest and growth inhibition of CRC cells independent of p53. Furthermore, ablation of BECN1 inhibits xenograft tumor growth through elevated RB expression and reduced autophagy, while simultaneous silencing of RB1 restores tumor growth but has little effect on autophagy. Mechanistically, knockdown of BECN1 promotes the complex formation of MDM2 and MDMX, resulting in MDM2-dependent MDMX instability and RB stabilization. Our results demonstrate that nuclear Beclin 1 can promote cell cycle progression through modulation of the MDM2/X-RB pathway and suggest that Beclin 1 promotes CRC development by facilitating both cell cycle progression and autophagy. Full article
Show Figures

Figure 1

Figure 1
<p>Clinical relevance of Beclin 1 or RB expression in human colorectal cancer specimen. (<b>a</b>) Overall survival (OS) and relapse-free survival (RFS) of colon adenocarcinoma patients were based on <span class="html-italic">BECN1</span> mRNA expression and protein expression from TCGA-COAD dataset. Patients were stratified into high- or low-expression groups using optimal cutoff from survminer package. The blue and red areas represent the 95% confidence intervals of the high- and low-expressed groups, respectively. The tables listed below the plots show the number of patients at risk at each time point in each group. (<b>b</b>) Pearson’s correlation coefficient between the protein expressions of Beclin 1 and RB in the TCGA-COADREAD dataset was obtained from the LinkedOmics database (<a href="http://www.linkedomics.org/admin.php" target="_blank">http://www.linkedomics.org/admin.php</a> (accessed on 10 September 2022)). Reverse-Phase Protein Array, RPPA. (<b>c</b>,<b>d</b>) Human colorectal cancer tissue microarrays consisting of cancer specimens from three cancer stages (stage I, <span class="html-italic">n</span> = 6; II, <span class="html-italic">n</span> = 14; III, <span class="html-italic">n</span> = 7) were subjected to IHC staining for RB and Beclin 1, with quantitative analyses using average optical density (AOD). (<b>e</b>) Pearson’s correlation between RB and Beclin 1 in colorectal cancer patients was analyzed. Scale bar = 50 <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Ablation of <span class="html-italic">BECN1</span> leads to cell cycle G1/S arrest and growth inhibition through upregulation of RB protein expression. HCT116 or HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells stably expressing shRNA specific for <span class="html-italic">BECN1</span> (#1 or #2) or control shRNA (shC) were subjected to western blot (<b>a</b>), colony formation assays (<b>b</b>), real-time cell analyses (RTCA) (<b>c</b>), and FACS analyses (<b>d</b>). Quantifications of cell cycle derived from FACS analyses (<b>e</b>). HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells stably expressing shBeclin 1 were infected with lentivirus expressing Beclin 1 or Beclin 1<math display="inline"><semantics> <msup> <mrow/> <mrow> <mi mathvariant="normal">L</mi> <mn>184</mn> <mi mathvariant="normal">A</mi> <mo>+</mo> <mi mathvariant="normal">L</mi> <mn>187</mn> <mi mathvariant="normal">A</mi> </mrow> </msup> </semantics></math> followed by western blot (<b>f</b>), colony formation assays (<b>g</b>), and FACS analyses (<b>h</b>). Data derived from three independent experiments were presented as mean ± SD. * <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. The uncropped blots and molecular weight markers are shown in <a href="#app1-cancers-14-04735" class="html-app">Supplementary File S1</a>.</p>
Full article ">Figure 3
<p>Ablation of <span class="html-italic">BECN1</span> reduces MDMX expression resulting in the RB-dependent cell cycle G1/S arrest and growth inhibition. HCT116 or HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells stably expressing shBeclin 1 were infected with lentivirus shRNA specific for <span class="html-italic">RB1</span> (#1 or #2) or control shRNA and subjected to western blot (<b>a</b>), colony formation assays (<b>b</b>), real-time cell analyses (RTCA) (<b>c</b>), and FACS analyses (<b>d</b>). Quantifications of cell cycle derived from FACS analyses (<b>e</b>). HCT116 or HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells stably expressing shBeclin 1 were infected with lentivirus expressing MDMX or MDMX<math display="inline"><semantics> <msup> <mrow/> <mrow> <mi mathvariant="normal">C</mi> <mn>437</mn> <mi mathvariant="normal">A</mi> </mrow> </msup> </semantics></math> and subjected to western blot (<b>f</b>), colony formation assays (<b>g</b>), real-time cell analyses (RTCA) (<b>h</b>), and FACS analyses (<b>i</b>). Data were derived from three independent experiments and were presented as mean ± SD. * <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. The uncropped blots and molecular weight markers are shown in <a href="#app1-cancers-14-04735" class="html-app">Supplementary File S1</a>.</p>
Full article ">Figure 4
<p>Ablation of <span class="html-italic">BECN1</span> facilities MDM2–MDMX interaction to promote MDMX degradation and RB protein stabilization. (<b>a</b>) HCT116 cells stably expressing shBeclin 1 or control shRNA (shC) were treated with 50 μg/mL cycloheximide (CHX) for the indicated time intervals. (<b>b</b>) The plots of RB protein half-life. (<b>c</b>) HCT116 cells stably expressing shBeclin 1 or control shRNA (shC) were treated with 20 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>M MG132 for 12 h before collection. Total cell lysates were subjected to western blot. (<b>d</b>) 293FT cells stably expressing shC or shBeclin 1 (#1 or #2) were transfected with expressing plasmids of Myc-MDMX, His-Ub. Then, cells were treated with 20 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>M MG132 for 6 h before collection. Ubiquitylation of MDMX was examined by IP-western analyses. (<b>e</b>) HCT116 or HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells were subjected to co-immunoprecipitation (co-IP) using either a specific antibody for Beclin 1 and MDMX or IgG control followed by western blot analyses. (<b>f</b>) HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells stably expressing shC or shBeclin 1 were subjected to co-immunoprecipitation (co-IP) assay using a specific antibody for MDM2 or IgG control followed by western blot analyses as indicated. (<b>g</b>–<b>j</b>) 293FT cells were co-transfected with Flag-Beclin 1 (FL) or an indicated Beclin 1 deletion construct and MDM2 (<b>g</b>), MDMX (<b>h</b>), or RB (<b>i</b>). Total cell lysates were subjected to IP-western analyses. (<b>j</b>) A schematic illustration shows MDM2–Beclin 1, MDMX–Beclin 1, and RB–Beclin 1 interaction domains. (<b>k</b>) HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math> cells transfected with siMDM2, siMDMX, siMDM2/siMDMX were subjected to co-immunoprecipitation (co-IP) using a specific antibody followed by western blot analyses as indicated. (<b>l</b>) A sketch depicts a model of the influence of Beclin 1 on quadruple protein complexes formation among Beclin 1, RB, MDMX, and MDM2. The uncropped blots and molecular weight markers are shown in <a href="#app1-cancers-14-04735" class="html-app">Supplementary File S1</a>.</p>
Full article ">Figure 5
<p>Knockdown of <span class="html-italic">BECN1</span> suppresses xenograft tumor growth, which is reversed by simultaneous ablation of <span class="html-italic">RB1</span>. (<b>a</b>) For tumor growth assay, HCT116 p53<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mo>/</mo> <mo>−</mo> </mrow> </msup> </semantics></math>-Antares2 cells (5 × 10<math display="inline"><semantics> <msup> <mrow/> <mn>6</mn> </msup> </semantics></math>) stably expressing shBeclin 1 or shBeclin 1+shRB were subcutaneously transplanted into nude mice (<span class="html-italic">n</span> = 6/group). Diphenylterazine (DTZ) was injected into tumor regions in anesthetized mice and bioluminescence was subsequently imaged using Caliper IVIS Lumina III. (<b>b</b>) The harvested tumors were photographed. (<b>c</b>) The tumor volumes of cell-derived xenografts were determined every other day. (<b>d</b>) The tumor sections were subjected to immunohistochemistry (IHC) staining for Beclin 1, RB, Ki67, p62, and LC3 (Scale bar = 50 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m; Inset scale bar = 10 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m). (<b>e</b>) The respective quantification data were presented as mean ± SEM. * <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. (<b>f</b>) A model depicts that nuclear Beclin 1 interferes with MDM2–MDMX interaction to stabilize MDMX, leading to the RB protein instability, cell cycle progression, and ultimately tumor growth.</p>
Full article ">
16 pages, 1749 KiB  
Article
Teenage-Onset Colorectal Cancers in a Digenic Cancer Predisposition Syndrome Provide Clues for the Interaction between Mismatch Repair and Polymerase δ Proofreading Deficiency in Tumorigenesis
by Esther Schamschula, Miriam Kinzel, Annekatrin Wernstedt, Klaus Oberhuber, Hendrik Gottschling, Simon Schnaiter, Nicolaus Friedrichs, Sabine Merkelbach-Bruse, Johannes Zschocke, Richard Gallon and Katharina Wimmer
Biomolecules 2022, 12(10), 1350; https://doi.org/10.3390/biom12101350 - 22 Sep 2022
Cited by 12 | Viewed by 2828
Abstract
Colorectal cancer (CRC) in adolescents and young adults (AYA) is very rare. Known predisposition syndromes include Lynch syndrome (LS) due to highly penetrant MLH1 and MSH2 alleles, familial adenomatous polyposis (FAP), constitutional mismatch-repair deficiency (CMMRD), and polymerase proofreading-associated polyposis (PPAP). Yet, 60% of [...] Read more.
Colorectal cancer (CRC) in adolescents and young adults (AYA) is very rare. Known predisposition syndromes include Lynch syndrome (LS) due to highly penetrant MLH1 and MSH2 alleles, familial adenomatous polyposis (FAP), constitutional mismatch-repair deficiency (CMMRD), and polymerase proofreading-associated polyposis (PPAP). Yet, 60% of AYA-CRC cases remain unexplained. In two teenage siblings with multiple adenomas and CRC, we identified a maternally inherited heterozygous PMS2 exon 12 deletion, NM_000535.7:c.2007-786_2174+493del1447, and a paternally inherited POLD1 variant, NP_002682.2:p.Asp316Asn. Comprehensive molecular tumor analysis revealed ultra-mutation (>100 Mut/Mb) and a large contribution of COSMIC signature SBS20 in both siblings’ CRCs, confirming their predisposition to AYA-CRC results from a high propensity for somatic MMR deficiency (MMRd) compounded by a constitutional Pol δ proofreading defect. COSMIC signature SBS20 as well as SBS26 in the index patient’s CRC were associated with an early mutation burst, suggesting MMRd was an early event in tumorigenesis. The somatic second hits in PMS2 were through loss of heterozygosity (LOH) in both tumors, suggesting PPd-independent acquisition of MMRd. Taken together, these patients represent the first cases of cancer predisposition due to heterozygous variants in PMS2 and POLD1. Analysis of their CRCs supports that POLD1-mutated tumors acquire hypermutation only with concurrent MMRd. Full article
Show Figures

Figure 1

Figure 1
<p>Immunohistochemical staining of mismatch-repair (MMR) proteins (PMS2, MLH1, MSH2 and MSH6) in the tumor of the index patient (<b>A</b>) and his sister (<b>B</b>). Loss of PMS2 expression is seen in carcinoma epithelia (red arrows) but not in tumor-infiltrating leukocytes (blue arrows; <b>A</b>,<b>B</b>). PMS2 expression is also absent or strongly reduced in the patient’s non-dysplastic crypts adjacent to carcinoma tissue (black arrows) (<b>A</b>). An enlarged view of the relevant areas (black box) is shown below the upper panels (<b>A</b>,<b>B</b>). In the pedigree of the family (<b>C</b>), the identified germline pathogenic variants (PVs), <span class="html-italic">PMS2</span>:c.2007-786_2174+493del1447 (green points) and <span class="html-italic">POLD1</span>:c.946G&gt;A (blue triangles), are indicated. Cancers are labelled as filled quarters (see figure key). The index patient is depicted by an arrow.</p>
Full article ">Figure 2
<p>Identification and characterization of the familial <span class="html-italic">PMS2</span> exon 12 deletion. Copy number (CN) CN analysis of next-generation sequencing (NGS) data with the SeqNext software shows a loss of one copy of exon 12 of either the <span class="html-italic">PMS2</span> gene or the <span class="html-italic">PMS2CL</span> pseudogene. Bars indicate the relative CN of each <span class="html-italic">PMS2</span> exon (shown on the <span class="html-italic">x</span>-axis) in the patient compared to 12 controls. Red lines indicate thresholds for CN variant calling (<b>A</b>). Direct <span class="html-italic">PMS2</span> gene-specific cDNA sequencing using a reverse primer located in exon 13 (black arrow) shows exon 12 skipping in 50% of the transcripts (<b>B</b>). Sequencing of the deletion-spanning gene-specific amplicon reveals an Alu-mediated 1447 bp-deletion (Δex12). These (green and purple) and other (gray) Alu elements in the introns are indicated as arrows in the schematic illustration of <span class="html-italic">PMS2</span> exons 11 and 12 and flanking intronic sequences. The amplified region using an unspecific, i.e., not discriminating between <span class="html-italic">PMS2</span> and <span class="html-italic">PMS2CL</span>, forward primer (black arrow, universal) and a <span class="html-italic">PMS2</span>-specific reverse primer (blue arrow) is indicated above the scheme. The shortened 1256 bp PCR amplicon generated from the patient’s DNA and the wild-type (wt) 2703 bp amplicon generated from a control (Ct) DNA (Ø: negative control) are visible in the agarose gel shown on the left below the scheme. Sanger sequencing of the deletion-spanning amplicon of the patient shows a transition from AluI (green) to AluII (purple). The last AluI-specific and the first AluII-specific nucleotides are marked in bold in the sequence shown right below the scheme. The intervening 25 bp sequence in which AluI does not differ from AluII is framed in gray (<b>C</b>).</p>
Full article ">Figure 3
<p>Histogram of somatic single nucleotide variants (SNVs) detected in the patient’s (IV-1; (<b>A</b>)) and his sister’s (IV-2; (<b>B</b>)) tumor. Short tandem repeat variants are highlighted in dark turquoise. Late and early events with low and high variant allele frequencies (VAFs), respectively, are shown as framed solid. Possible artefacts are framed dashed (<b>A</b>,<b>B</b>). Pie chart of the signature contribution of all SNVs for the patient’s (IV-1 all events) and his sister’s (IV-2 all events) tumor and signature contribution of separately analyzed late and early SNV events for the patient’s tumor (IV-1 late events; IV-1 early events). For the separate analysis of late and early events, regions suspected to be affected by CN variants and/or loss of heterozygosity (LOH) events were omitted (see Methods and <a href="#app1-biomolecules-12-01350" class="html-app">Figure S3A,B</a>) (<b>C</b>).</p>
Full article ">
23 pages, 3866 KiB  
Article
Anti-Human CD9 Fab Fragment Antibody Blocks the Extracellular Vesicle-Mediated Increase in Malignancy of Colon Cancer Cells
by Mark F. Santos, Germana Rappa, Simona Fontana, Jana Karbanová, Feryal Aalam, Derek Tai, Zhiyin Li, Marzia Pucci, Riccardo Alessandro, Chikao Morimoto, Denis Corbeil and Aurelio Lorico
Cells 2022, 11(16), 2474; https://doi.org/10.3390/cells11162474 - 10 Aug 2022
Cited by 10 | Viewed by 3686
Abstract
Intercellular communication between cancer cells themselves or with healthy cells in the tumor microenvironment and/or pre-metastatic sites plays an important role in cancer progression and metastasis. In addition to ligand–receptor signaling complexes, extracellular vesicles (EVs) are emerging as novel mediators of intercellular communication [...] Read more.
Intercellular communication between cancer cells themselves or with healthy cells in the tumor microenvironment and/or pre-metastatic sites plays an important role in cancer progression and metastasis. In addition to ligand–receptor signaling complexes, extracellular vesicles (EVs) are emerging as novel mediators of intercellular communication both in tissue homeostasis and in diseases such as cancer. EV-mediated transfer of molecular activities impacting morphological features and cell motility from highly metastatic SW620 cells to non-metastatic SW480 cells is a good in vitro example to illustrate the increased malignancy of colorectal cancer leading to its transformation and aggressive behavior. In an attempt to intercept the intercellular communication promoted by EVs, we recently developed a monovalent Fab fragment antibody directed against human CD9 tetraspanin and showed its effectiveness in blocking the internalization of melanoma cell-derived EVs and the nuclear transfer of their cargo proteins into recipient cells. Here, we employed the SW480/SW620 model to investigate the anti-cancer potential of the anti-CD9 Fab antibody. We first demonstrated that most EVs derived from SW620 cells contain CD9, making them potential targets. We then found that the anti-CD9 Fab antibody, but not the corresponding divalent antibody, prevented internalization of EVs from SW620 cells into SW480 cells, thereby inhibiting their phenotypic transformation, i.e., the change from a mesenchymal-like morphology to a rounded amoeboid-like shape with membrane blebbing, and thus preventing increased cell migration. Intercepting EV-mediated intercellular communication in the tumor niche with an anti-CD9 Fab antibody, combined with direct targeting of cancer cells, could lead to the development of new anti-cancer therapeutic strategies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>CD9 expression in SW480 and SW620 cells. (<b>A</b>) CD9 expression was investigated by indirect immunofluorescence labeling using anti-CD9 5H9 Abs on either intact or permeabilized SW480 and SW620 cells cultured on poly-D-lysine-coated dishes. Nuclei were stained with DAPI and samples were observed by CLSM. Composite images (top panels) or single x-y sections (middle and bottom panels) are shown. Arrowheads indicate membrane blebs, while asterisks mark cytoplasmic CD9 immunoreactivity. (<b>B</b>) The amount of surface CD9 antigens per cell detected with FITC-conjugated eBioSN4 Abs was estimated using a flow cytometer calibrated with fluorescent microparticles. (<b>C</b>) Total CD9 antigens were analyzed by immunoblotting (top panel) using 5H9 Abs and quantified (bottom panel). The samples were normalized to β-actin. Molecular mass markers (kDa) are indicated. Arrowhead indicates the protein of interest. (<b>D</b>,<b>E</b>) SW620 cells stably transfected with CD9-GFP were analyzed either by fluorescence microscopy without permeabilization (<b>D</b>) or immunoblotting (<b>E</b>). For the microscopy, nuclei were stained with DAPI and samples were observed by CLSM. A composite image (top panel) or a single x-y section (bottom panel) is displayed. Arrowheads indicate membrane blebs. For immunoblotting, the membrane was probed with 5H9 Abs. The arrow and arrowhead indicate the CD9-GFP fusion protein and the endogenous CD9, respectively. Means ± S.D. and individual values for each experiment are shown (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> values are indicated. Scale bars, 10 µm.</p>
Full article ">Figure 2
<p>Characterization of EVs released by SW620, CD9-GFP<sup>+</sup> SW620, and SW480 cells. (<b>A</b>–<b>E</b>) EVs were recovered from the conditioned media of SW620, CD9-GFP<sup>+</sup> SW620, and SW480 cells by differential centrifugation, and the resulting 200,000× <span class="html-italic">g</span> pellets were analyzed by the ZetaView particle analyzer (<b>A</b>), immunoblotting (<b>B</b>,<b>C</b>), and dSTORM (<b>D</b>,<b>E</b>). The concentration and size of EVs derived from the indicated cells are shown (<b>A</b>). Note the presence of a common population of small particles (&lt;200 nm) with a peak at 100–150 nm (pink area), and larger ones (350–500 nm, gray areas) enriched in SW480 samples. EVs, and for comparison the cells from which they were derived, were probed by immunoblotting for CD9, CD63, CD81, Alix, and Calnexin (<b>B</b>,<b>C</b>). Arrowheads and brackets indicate the endogenous proteins of interest, while the arrow points to the CD9-GFP fusion protein. Molecular mass markers (kDa) are indicated. EVs were imaged after immunolabeling of three tetraspanins using dSTORM (<b>D</b>). The proteins of interest (CD9, CD63, and CD81) were pseudo-colored as indicated. Small single-, double-, and triple-positive EVs were shown (<b>D</b>) and quantified (<b>E</b>). Means ± S.D. and individual values for the three experiments are shown (<span class="html-italic">n</span> &gt; 5000 EVs per experiment). Note that small and large EVs derived from SW480 cells were also quantified (<a href="#app1-cells-11-02474" class="html-app">Supplementary Figure S2</a>). Scale bars, 50 nm.</p>
Full article ">Figure 3
<p>Effects of CD9 Fab and divalent Abs on the internalization of SW620 cell-derived CD9-GFP<sup>+</sup> EVs into SW480 cells. (<b>A</b>–<b>D</b>) SW480 cells and fluorescent EVs derived from CD9-GFP<sup>+</sup> SW620 cells (1 × 10<sup>9</sup> particles) were individually pre-incubated for 30 min without (control) or with different concentrations of anti-CD9 Fab or divalent Ab as indicated before their co-incubation for 5 h in the absence or presence of Abs (protocol #1). Fixed cells were either stained with DAPI (<b>A</b>,<b>B</b>) or immunolabeled for SUN2 (<b>C</b>,<b>D</b>) before observation by CLSM. Note the presence of discrete punctate GFP signals in the cytoplasmic (<b>A</b>,<b>C</b>) or nucleoplasmic (<b>C</b>) compartments (asterisk and circle, respectively). Their intensity (<b>B</b>) or amount (<b>D</b>) was quantified using serial optical sections through a cell (see <a href="#app1-cells-11-02474" class="html-app">Supplementary Figure S3</a>). Means ± S.D. of individual signals from three independent experiments, as indicated by color coding, are shown (<span class="html-italic">n</span> &gt; 15 cells per experiment). <span class="html-italic">p</span> values are indicated. Arrowheads indicate CD9-GFP signals in the nuclear envelope invagination (<b>C</b>). Scale bars, 5 µm.</p>
Full article ">Figure 4
<p>Effects of CD9 Fab and divalent Ab on the pro-metastatic morphological alterations of SW480 cells exposed to SW620 cell-derived EVs. (<b>A</b>–<b>C</b>) SW480 cells and SW620 cell-derived EVs (1 × 10<sup>9</sup> particles) were individually pre-incubated for 30 min with different concentrations of anti-CD9 Fab (red) or divalent Ab (green) as indicated before their co-incubation for 5 h in the presence of Abs, as described for protocol #1. As negative and positive controls, cells were not exposed (SW480, grey) or were exposed to EVs in the absence of Ab (+ SW620 EV, blue), respectively. Afterward, fixed cells were stained with DAPI and fluorochrome-conjugated phalloidin to label nuclei and actin filaments, respectively, before observation by CLSM (<b>A</b>). Single sections are presented. Rounded cell morphology and membrane blebs induced by EVs are indicated by the letter R and the arrowheads, respectively. The percentage of cells with rounded morphology (<b>B</b>) or membrane blebs (<b>C</b>) was quantified. Means ± S.D. and individual values for each experiment are shown (<span class="html-italic">n</span> = 4). At least 100 cells were evaluated for each experiment. <span class="html-italic">p</span> values are indicated. Similar experiments were performed by pre-incubating only EVs or cells with Abs (<a href="#app1-cells-11-02474" class="html-app">Supplementary Figure S5</a>). Scale bars, 10 µm.</p>
Full article ">Figure 5
<p>Effects of CD9 Fab and divalent Ab on the migration of SW480 cells exposed to SW620 cell-derived EVs. (<b>A</b>) The migration–wound healing assay was performed by introducing a scratch on confluent SW480 cell monolayers cultured on 12-well standard cell culture plates and incubating them for 5 h in the absence (negative control, grey) or presence (positive control, blue) of SW620 cell-derived EVs (1 × 10<sup>9</sup> particles/mL). Alternatively, after introducing a scratch on the cell monolayer, cells and EVs were individually pre-incubated for 30 min with different concentrations of anti-CD9 Fab (red) or divalent Ab (green) as indicated before their 5 h co-incubation in the presence of Abs (<b>A</b>). The percentage of remaining wound areas after 5 h was quantified. Baseline (100%, dashed line) refers to wound area at 0 h. (<b>B</b>,<b>C</b>) Solely EVs (<b>B</b>) or cells (<b>C</b>) were pre-incubated for 30 min with Abs prior to co-incubation with cells or EVs for 5 h. Wound area was quantified. Controls (white) are shown for comparison. Means ± S.D. from multiple scratches are shown (<span class="html-italic">n</span> = 4–13). (<b>D</b>–<b>F</b>) The migration–Transwell filter assay was performed using a Transwell chamber as illustrated (<b>D</b>), where SW480 (<b>E</b>) or SW620 (<b>F</b>) cells were added to the upper chamber. Cells and EVs were pre-treated, as described in panel A, using 25 µg/mL CD9 Fab or divalent Ab ((<b>D</b>), steps 1 and 2) before 24 h of co-incubation ((<b>D</b>), step 3). In the case of SW620 cells, they were not incubated with EVs (<b>F</b>). The amount of migrating cells recovered in the lower chamber was then quantified. Each individual value is shown. <span class="html-italic">p</span> values are indicated. n.s., not significant.</p>
Full article ">Figure 6
<p>The lack of CD9 in SW480 cells impedes pro-metastatic morphological alterations produced by SW620 cell-derived EVs. (<b>A</b>) Parental (control) or CD9-knockdown (shCD9) SW480 cells were analyzed by immunoblotting for CD9 and β-actin. Molecular mass markers (kDa) are indicated. Arrowhead indicates the protein of interest. (<b>B</b>) SW480 cells as indicated were incubated with (+) SW620 cell-derived EVs (1 × 10<sup>9</sup> particles) or without (–) for 5 h. Afterward, fixed cells were stained with DAPI and fluorochrome-conjugated phalloidin to label nuclei and actin filaments, respectively, before observation by CLSM. The percentage of cells with rounded morphology (top panel) or membrane blebs (bottom panel) was quantified. Means ± S.D. and individual values for each experiment are shown (<span class="html-italic">n</span> = 3). At least 100 cells were evaluated for each experiment. <span class="html-italic">p</span> values are indicated. n.s., not significant.</p>
Full article ">
17 pages, 3568 KiB  
Article
Single-Cell FISH Analysis Reveals Distinct Shifts in PKM Isoform Populations during Drug Resistance Acquisition
by Seong Ho Kim, Ji Hun Wi, HyeRan Gwak, Eun Gyeong Yang and So Yeon Kim
Biomolecules 2022, 12(8), 1082; https://doi.org/10.3390/biom12081082 - 6 Aug 2022
Cited by 2 | Viewed by 2083
Abstract
The Warburg effect, i.e., the utilization of glycolysis under aerobic conditions, is recognized as a survival advantage of cancer cells. However, how the glycolytic activity is affected during drug resistance acquisition has not been explored at single-cell resolution. Because the relative ratio of [...] Read more.
The Warburg effect, i.e., the utilization of glycolysis under aerobic conditions, is recognized as a survival advantage of cancer cells. However, how the glycolytic activity is affected during drug resistance acquisition has not been explored at single-cell resolution. Because the relative ratio of the splicing isoform of pyruvate kinase M (PKM), PKM2/PKM1, can be used to estimate glycolytic activity, we utilized a single-molecule fluorescence in situ hybridization (SM-FISH) method to simultaneously quantify the mRNA levels of PKM1 and PKM2. Treatment of HCT116 cells with gefitinib (GE) resulted in two distinct populations of cells. However, as cells developed GE resistance, the GE-sensitive population with reduced PKM2 expression disappeared, and GE-resistant cells (Res) demonstrated enhanced PKM1 expression and a tightly regulated PKM2/PKM1 ratio. Our data suggest that maintaining an appropriate PKM2 level is important for cell survival upon GE treatment, whereas increased PKM1 expression becomes crucial in GE Res. This approach demonstrates the importance of single-cell-based analysis for our understanding of cancer cell metabolic responses to drugs, which could aid in the design of treatment strategies for drug-resistant cancers. Full article
Show Figures

Figure 1

Figure 1
<p>Detection of endogenous mRNAs of PKM isoforms in HCT116 cells using SM-FISH-STIC. (<b>A</b>) Schematic diagram of alternative splicing of PKM isoforms. PKM1 and PKM2 are produced via alternative splicing of the mutually exclusive exons 9 and 10. (<b>B</b>) Fluorescence images of single PKM mRNAs in HCT116 cells. PKM1 was detected with AX488 (green), and PKM2 was detected with Cy3. Images were processed as described in the Materials and Methods section and visualized using a maximum intensity projection of the entire <span class="html-italic">z</span>-axis of the cells (20 μm). Scale bar, 10 μm. (<b>C</b>) (<b>Left</b>) 2D plot of PKM isoforms in HCT116 cells. Each point corresponds to the number of mRNA molecules of each PKM isoform in individual cells. A linear regression of all the points is shown as a red line. The slope of the blue line corresponds to 1.2 (the average ratio of PKM2 to PKM1), and it passes through the average value of each PKM isoform (132, 159). (<b>Right</b>) Distribution of the total number of PKM isoform molecules (<b>top</b>) and the relative ratio of PKM2 to PKM1 (<b>bottom</b>) in individual cells.</p>
Full article ">Figure 2
<p>Effects of isoform-specific knockdown on PKM mRNA expression. Cells were treated with a scrambled siRNA (Con), PKM1-specific siRNA (SiM1), or PKM2-specific siRNA (siM2). (<b>A</b>) RT-PCR analysis of PKM isoforms. M1 and M2 corresponds to PKM1 and PKM2, respectively. Error bars represent average ± SD (<span class="html-italic">n</span> = 3) (<b>B</b>) Quantitative SM-FISH-STIC analysis of PKM isoforms. The number of mRNA molecules of each PKM isoform in individual cell was counted, and the average value is presented. M1 and M2 corresponds to PKM1 and PKM2, respectively. Error bars represent average ± standard error of the mean (SEM) (<span class="html-italic">n</span> ≥ 122) (<b>C</b>) 2D plot of each PKM isoform in HCT116 cells. Each point corresponds to the number of PKM isoforms in individual cell. The region enclosed within the dotted lines (the average ± SD in control cells) is highlighted in blue. The green box represents cells with a high expression level of PKM2 when PKM1 was depleted (greater than the average value + SD in control cells), and the red box represents cells with a high expression level of PKM1 when PKM2 was depleted (greater than the average value + SD in control cells). (<b>D</b>) Distribution of PKM1 (<b>left</b>) and PKM2 mRNA (<b>right</b>) from individual cells. The black, dark green, and red lines represent the Gaussian fits of each distribution. The black arrows indicate the corresponding average value + SD of the control. (<b>E</b>) Determination of relative mRNA levels of regulatory proteins involved in PKM splicing determined by RT-PCR. Data are shown as average ± SD of three independent experiments.* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Effects of GE on the mRNA levels of PKM isoforms and proteins involved in PKM regulation. Cells were untreated (−) or GE-treated for 24 h (+). GE-resistant cells (GE Res) were also analyzed. (<b>A</b>) Evaluation of the cell viability. HCT116 cells (Con) or GE-resistant HCT116 cells (Res) were cultured in the absence or presence of 30 μM GE for 24 h, and viable cells were counted as described in the Materials and Methods section. Data are shown as average ± SD of three independent experiments. Scale bar, 200 μm. (<b>B</b>) The effect of GE on cell viability. HCT116 cells (Con) or GE-resistant HCT116 cells (Res) were treated with 30 μM GE for 24 h; then, the degree of EGFR expression and activation was measured by Western blotting. Untreated cells were used as a negative control. (<b>C</b>) RT-PCR analysis for PKM isoforms, c-Myc, hnRNPs, PTBP1, and Sp1. M1 and M2 in the graph correspond to PKM1 and PKM2, respectively. (<b>D</b>) Quantitative SM-FISH-STIC analysis of PKM isoforms. M1 and M2 correspond to PKM1 and PKM2, respectively. Error bars represent average ± SEM (<span class="html-italic">n</span> ≥ 162). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Effects of GE on the distribution of PKM isoforms. Cells were untreated (GE (−)) or GE-treated for 24 h (GE (+)). GE-resistant cells (GE Res) were also analyzed. (<b>A</b>) 2D plot of each PKM isoform in GE (−) or GE (+). Each point corresponds to the number of mRNA molecules of each PKM isoform in individual cells. The dotted lines represent the average value ± SD of each PKM isoform in GE (−). The dark green line represents the value at which the two PKM2 Gaussian fits (shown in orange and blue in <a href="#biomolecules-12-01082-f004" class="html-fig">Figure 4</a>C) overlap. (<b>B</b>) 2D plot of each PKM isoform in GE (−) or GE Res. Each point corresponds to the number of mRNA molecules of each PKM isoform in individual cells. The dotted lines represent the average value ± SD of PKM isoforms from GE (−). (<b>C</b>) Distribution of PKM1 (left) and PKM2 mRNAs (right). The black, dark green, and red lines represent the Gaussian fit of each distribution. In the PKM2 mRNA distribution (right), the dark green line represents the sum of the two Gaussian fits (orange and blue lines) for GE (+). The black arrow highlights the overlap of the two Gaussian fits. (<b>D</b>) Distribution of the ratio of PKM2 to PKM1. The black, green, and red lines represent the Gaussian fits of each distribution.</p>
Full article ">Figure 5
<p>The importance of PKM isoforms in the acquisition of drug resistance. (<b>A</b>) The effect of PKM isoform depletion on the response to GE. Control or GE-resistant HCT116 cells (Con or GE Res) were pretreated with the appropriate siRNAs (siM1 and siM2) for 24 h and exposed to GE (30 μM) for 24 h. The viable cells were counted in three independent experiments as described in the Materials and Methods section. Data are shown as average ± SD of three independent experiments. Scale bar, 300 μm. (<b>B</b>) Expression of PKM proteins upon GE treatment in nuclear and cytoplasmic fractions. Individual fractions were normalized by vinculin (cytoplasmic extract) and Lamin B1 (nuclear extract). Data are shown as average ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span>&lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
Back to TopTop