Spandidos Publications Logo
  • About
    • About Spandidos
    • Aims and Scopes
    • Abstracting and Indexing
    • Editorial Policies
    • Reprints and Permissions
    • Job Opportunities
    • Terms and Conditions
    • Contact
  • Journals
    • All Journals
    • Oncology Letters
      • Oncology Letters
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Oncology
      • International Journal of Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular and Clinical Oncology
      • Molecular and Clinical Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Experimental and Therapeutic Medicine
      • Experimental and Therapeutic Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Molecular Medicine
      • International Journal of Molecular Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Biomedical Reports
      • Biomedical Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Reports
      • Oncology Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular Medicine Reports
      • Molecular Medicine Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • World Academy of Sciences Journal
      • World Academy of Sciences Journal
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Functional Nutrition
      • International Journal of Functional Nutrition
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Epigenetics
      • International Journal of Epigenetics
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Medicine International
      • Medicine International
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
  • Articles
  • Information
    • Information for Authors
    • Information for Reviewers
    • Information for Librarians
    • Information for Advertisers
    • Conferences
  • Language Editing
Spandidos Publications Logo
  • About
    • About Spandidos
    • Aims and Scopes
    • Abstracting and Indexing
    • Editorial Policies
    • Reprints and Permissions
    • Job Opportunities
    • Terms and Conditions
    • Contact
  • Journals
    • All Journals
    • Biomedical Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Experimental and Therapeutic Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Epigenetics
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Functional Nutrition
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Molecular Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Medicine International
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular and Clinical Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular Medicine Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Letters
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • World Academy of Sciences Journal
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
  • Articles
  • Information
    • For Authors
    • For Reviewers
    • For Librarians
    • For Advertisers
    • Conferences
  • Language Editing
Login Register Submit
  • This site uses cookies
  • You can change your cookie settings at any time by following the instructions in our Cookie Policy. To find out more, you may read our Privacy Policy.

    I agree
Search articles by DOI, keyword, author or affiliation
Search
Advanced Search
presentation
Oncology Letters
Join Editorial Board Propose a Special Issue
Print ISSN: 1792-1074 Online ISSN: 1792-1082
Journal Cover
December-2025 Volume 30 Issue 6

Full Size Image

Sign up for eToc alerts
Recommend to Library

Journals

International Journal of Molecular Medicine

International Journal of Molecular Medicine

International Journal of Molecular Medicine is an international journal devoted to molecular mechanisms of human disease.

International Journal of Oncology

International Journal of Oncology

International Journal of Oncology is an international journal devoted to oncology research and cancer treatment.

Molecular Medicine Reports

Molecular Medicine Reports

Covers molecular medicine topics such as pharmacology, pathology, genetics, neuroscience, infectious diseases, molecular cardiology, and molecular surgery.

Oncology Reports

Oncology Reports

Oncology Reports is an international journal devoted to fundamental and applied research in Oncology.

Experimental and Therapeutic Medicine

Experimental and Therapeutic Medicine

Experimental and Therapeutic Medicine is an international journal devoted to laboratory and clinical medicine.

Oncology Letters

Oncology Letters

Oncology Letters is an international journal devoted to Experimental and Clinical Oncology.

Biomedical Reports

Biomedical Reports

Explores a wide range of biological and medical fields, including pharmacology, genetics, microbiology, neuroscience, and molecular cardiology.

Molecular and Clinical Oncology

Molecular and Clinical Oncology

International journal addressing all aspects of oncology research, from tumorigenesis and oncogenes to chemotherapy and metastasis.

World Academy of Sciences Journal

World Academy of Sciences Journal

Multidisciplinary open-access journal spanning biochemistry, genetics, neuroscience, environmental health, and synthetic biology.

International Journal of Functional Nutrition

International Journal of Functional Nutrition

Open-access journal combining biochemistry, pharmacology, immunology, and genetics to advance health through functional nutrition.

International Journal of Epigenetics

International Journal of Epigenetics

Publishes open-access research on using epigenetics to advance understanding and treatment of human disease.

Medicine International

Medicine International

An International Open Access Journal Devoted to General Medicine.

Journal Cover
December-2025 Volume 30 Issue 6

Full Size Image

Sign up for eToc alerts
Recommend to Library

  • Article
  • Citations
    • Cite This Article
    • Download Citation
    • Create Citation Alert
    • Remove Citation Alert
    • Cited By
  • Similar Articles
    • Related Articles (in Spandidos Publications)
    • Similar Articles (Google Scholar)
    • Similar Articles (PubMed)
  • Download PDF
  • Download XML
  • View XML
Article Open Access

Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle

  • Authors:
    • Xiao Han
    • Ling Liu
    • Jing Li
    • Yunxiao Zhi
    • Lanlan Zhao
    • Limin Yuan
    • Xuezhe Ouyang
    • Jie Liu
  • View Affiliations / Copyright

    Affiliations: Department of Medical Genetic and Prenatal Diagnosis, The Third Affiliated Hospital of Zhengzhou University, Zhengzhou, Henan 450052, P.R. China, Department of Oncology, The Affiliated Cancer Hospital of Zhengzhou University and Henan Cancer Hospital, Zhengzhou University, Zhengzhou, Henan 450008, P.R. China
    Copyright: © Han et al. This is an open access article distributed under the terms of Creative Commons Attribution License.
  • Article Number: 539
    |
    Published online on: September 22, 2025
       https://doi.org/10.3892/ol.2025.15285
  • Expand metrics +
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Metrics: Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )
Cited By (CrossRef): 0 citations Loading Articles...

This article is mentioned in:



Abstract

Ovarian cancer (OC) is the most common cause of gynecological cancer‑related death. Histone acetyltransferase 1 (HAT1) has generated interest as a potential target for therapy due to it being involved in a variety of diseases, including cancer. However, to the best of our knowledge, the role of HAT1 in OC has not yet been investigated. In the present study, HAT1 was upregulated in OC and the high expression of HAT1 was associated with unfavorable prognosis. The transcription factor forkhead box protein A1 (FOXA1) transcriptionally regulated HAT1 expression. Furthermore, HAT1 knockdown in OC cells significantly suppressed cell proliferation and colony formation. In addition, the inhibition of HAT1 promoted cell cycle arrest, and reduced cyclin‑dependent kinase (CDK)2, CDK4 and cyclin E levels in OC cells. Taken together, the present data suggested that HAT1 served an oncogenic role in OC; therefore, HAT1 may represents a new potential therapeutic target in OC treatment.

Introduction

Ovarian cancer (OC) has the greatest fatality rate among all global gynecological malignancies, with epithelial OC being the most frequently occurring type (1,2). Due to a lack of accurate early screening techniques for OC, diagnosis is delayed and ~50% of patients die within 5 years of diagnosis (3,4). Despite recent therapeutic breakthroughs, OC still has poor results due to the unknown gene regulation network that underpins its etiology (5). Notably, there is an abundance of clinical prognostic indicators for OC, such as CA125; however, the identification of meaningful biomarkers for early detection and improving treatment outcomes in OC is required.

Histone acetyltransferases (HATs) and deacetylases are capable of controlling DNA transcription by regulating histone acetylation and deacetylation. Overexpression or inappropriate recruitment of HATs have been associated with the development and progression of malignant tumors (6). HAT1 can only acetylate newly generated histone H4, and was the first HAT discovered and remains one of the insufficiently studied members of the HAT family (7). HAT1 is upregulated in numerous types of solid tumors, including lung, breast, esophageal and liver cancer, and functions as an oncogene to promote carcinogenesis (8–11). It has also been revealed that HAT1 operates as a transcription factor (TF), modulating cancer cell metabolism and treatment resistance (12,13).

To date, to the best of our knowledge, the role and function of HAT1 in OC have not been investigated. The present study aimed to elucidate the HAT1 expression levels in specimens from patients with OC and healthy human samples. Then the molecular mechanism of upstream regulator of HAT1 will be studied, as well as the function of HAT1 in regulating OC cell proliferation and colony formation activities. Furthermore, the role of HAT1 in regulation of cell cycle pathway in OC cells will be analyzed, and whether knockdown of HAT1 could decrease cell cycle-related protein expression levels in OC cells. Thus, the present results may contribute to a better understanding of the potential role of HAT1 in OC, which could provide new biomarkers for novel therapeutic strategies in OC.

Materials and methods

Cell culture and reagents

Human OC cell lines OVCAR3, HEY, A2780 and SKOV3, and the immortalized ovarian epithelial cell line IOSE386 were grown in RPMI-1640 medium with 10% FBS (both Gibco; Thermo Fisher Scientific, Inc.), 100 IU/ml penicillin and 100 µg/ml streptomycin (Thermo Fisher Scientific, Inc.). The cells were cultured in a 37°C incubator with 5% CO2, and cells were detached using 0.25% trypsin and passaged every 2 days. Small interfering RNA (siRNA) transfection was used to silence HAT1 expression in 3×106 HEY and SKOV3 cells at 150 nM concentration, and a siRNA transfection reagent (cat. no. sc-29528, Santa Cruz Biotechnology, Inc.) was used to transfect cells at room temperature for 48 h according to the manufacturer's guidelines. siRNAs including scrambled siRNA (siSCR) and siHAT1 were purchased from the Dharmacon (Revvity, Inc.). The ON-TARGETplus Human HAT1 siRNA-SMART pool contained four target sequences of HAT1 together in the tube: 5′-GCUACAGACUGGAUAUUAAUU-3′, 5′-CAGAUGAACCAAAUAGAAAUU-3′, 5′-CAACACAGCAAUUGAACUAUU-3′ and 5′-AGGAACUAGUGGAAGAUUAUU-3′; the siRNA also contained four target sequences of siSCR: 5′-UGGUUUACAUGUCGACUAA-3′, 5′-UGGUUUACAUGUUGUGUGA-3′, 5′-UGGUUUACAUGUUUUCUGA-3′ and 5′-UGGUUUACAUGUUUUCCUA-3′. The FOXA1 overexpression plasmid was synthesized by YouBio Biotechnology. A total of 1 µg pcDNA3-FOXA1 plasmid and 1 µg pcDNA3-Flag control plasmid was transfected into 3×106 HEY cells for 48 h using the FuGENE® HD transfection reagent at room temperature according to the manufacturer's instructions (Roche Diagnostics).

Western blot analysis

Lysates from whole cells (IOSE386, OVCAR3, HEY, A2780 and SKOV3) were extracted using RIPA buffer (Thermo Fisher Scientific, Inc.) with a protease inhibitor cocktail. Cell suspensions were homogenized and centrifuged at 16,000 × g for 20 min at 4°C to extract the supernatant. Equal amounts of protein (15 µg) were separated by SDS-PAGE (Upper layer 4% stacking gel and lower layer 8% separating gel) and transferred to a PVDF membrane, which was blocked with 5% non-fat milk for 2 h at room temperature. Subsequently, the membrane was treated with the indicated primary antibodies at 4°C overnight and then horseradish peroxidase-conjugated secondary antibodies (1:2,000; Cat.31430, Cat.31460; Thermo Fisher Scientific, USA) for another 2 h at room temperature. Signals were detected using an enhanced chemiluminescence reagent from Pierce (Thermo Fisher Scientific, USA) and analyzed with a ImageQuant LAS 4000 and ImageQuant TL 8.1 software (GE Healthcare). The following primary antibodies were used in the present study: Anti-HAT1 (1:1,000; cat no. 11432-1-AP; Proteintech Group, Inc.), anti-CDK2 (1:1,000; cat no. 10122-1-AP; Proteintech Group, Inc.), anti-FOXA1 (1:1,000; cat no. 20411-1-AP; Proteintech Group, Inc.), anti-CDK4 (1:1,000; cat no. 12790; Cell Signaling Technology, Inc.), anti-cyclin E (1:1,000; cat no. 20808; Cell Signaling Technology, Inc.) and anti-GAPDH (1:5,000; cat no. 60004-1-Ig; Proteintech Group, Inc.).

RNA isolation and reverse transcription-quantitative PCR (qPCR)

TRIzol® reagent (Invitrogen; Thermo Fisher Scientific, Inc.) was adopted for total mRNA isolation from cell lines (IOSE386, OVCAR3, HEY, A2780 and SKOV3). mRNA was reverse transcribed with PrimeScript™ RT Master Mix (Takara Bio, Inc.) as follows: Primer annealing at 25°C, 5 min; Reverse transcription: 42°C, 30–60 min; Enzyme inactivation: 85°C 5 min; and Hold: 4°C. qPCR was performed on QuantStudio 5 qPCR equipment (Applied Biosystems; Thermo Fisher Scientific, Inc.) with Universal SYBR Green Master reagent, and the cycling conditions were: Initial denaturation: 95°C, 3 min; PCR cycles (40 cycles): Denaturation: 95°C, 15 sec and Annealing/extension: 60°C, 30 sec (with fluorescence acquisition); and Melting curve analysis: 60–95°C. The relative gene expression levels were normalized to those of the internal control, GAPDH, using the 2−ΔΔCq method (14). The primer sequences were as follows: HAT1, forward, 5′-AAGCCATTCGGAACCTTACTTC-3′, reverse 5′-AGTGCCATCTTTCATCATCCAC-3′; and GAPDH, forward 5′-GGAGCGAGATCCCTCCAAAAT-3′, reverse 5′-GGCTGTTGTCATACTTCTCATGG-3′.

Colony formation assay

500 Cells (HEY and SKOV3 cells) were trypsinized, assessed and cultivated for 2 weeks in medium containing 10% FBS. Subsequently, the cells were washed in PBS, fixed with 4% paraformaldehyde for 15 min and stained with 1% crystal violet for 30 min at 37°C. The colonies visible at the plates were counted with ImageJ software v1.41 (National Institutes of Health) and images were captured.

Cell viability assay

2000 HEY and SKOV3 cells were trypsinized, measured and seeded into 96-well plates with medium containing 10% FBS. Each well received 10 µl Cell Counting Kit-8 (CCK-8) solution (Dojindo Laboratories, Inc.) and the cells were cultured in an incubator at 37°C for 1 h. At various time points (24, 48, 72, 96, 120 h), the absorbance was measured at 450 nm with a microplate reader and analyzed (Bio-Rad Laboratories, Inc.).

Cell cycle assay

A total of 5×106 HEY and SKOV3 cells were collected, washed with PBS and were subsequently fixed with 70% cold ethanol at 4°C overnight. The cells were then incubated with 500 µl PI/RNase A staining solution (Nanjing KeyGen Biotech Co., Ltd.) for 1 h at room temperature in the dark, according to the manufacturer's instructions, and were examined using a BD FACSCanto flow cytometer (BD Biosciences) and analyzed using FlowJo software version 4.5 (Tree Star, Inc.).

EdU staining

The EdU reagent was used to measure cell proliferation (Guangzhou RiboBio Co., Ltd.). A total of 2×104 HEY cells were cultured in 96-well plates. Cells were then treated with 50 µM EdU solution for an additional 2 h in 37°C. After washing with PBS, the cells were fixed at room temperature with 4% paraformaldehyde for 30 min and then permeabilized with 0.5% Triton X-100 for 10 min. Cells were exposed to 200 µl Apollo staining solution in the EdU kit for 30 min at room temperature in the dark. Finally, the DNA was stained with 200 µl 1 X Hoechst 33342 for 30 min at room temperature in the dark. Cells were then examined under a fluorescence microscope with excitation wavelength at 567 nm.

Cell survival assay

HEY and SKOV3 cells were seeded into 96-well plates (5×103/well) and treated with 0.0, 0.5, 1.0, 5, 10, 20, 30 and 50 µM) of HAT1 inhibitor JG-2016 or DMSO as indicated (Cat.HY-154944; MedChemExpress) for 72 h at room temperature. OD values were determined using a luminometer at an absorbance of 450 nm with the CCK-8 Kit as aforementioned (Dojindo Laboratories, Inc.), and cell viability levels were calculated based on OD values.

Luciferase reporter assay

The potential transcriptional factor FOXA1 binding sites of the HAT1 promoter region were analyzed through the UCSC Genome Browser (genome.ucsc.edu/) and TF database JASPAR (https://jaspar.elixir.no/). The pmiRGLO vectors (Promega Corporation) containing HAT1 wild-type (WT) and HAT1 mutant (MUT) sequences were constructed and Sanger sequenced. Subsequently, the HEY cells were co-transfected with pmiRGLO WT/MUT vectors and an empty or FOXA1 vector using Lipofectamine® 3000 (Invitrogen; Thermo Fisher Scientific, Inc.). Following 48 h of transfection, the Luciferase Assay Kit (Promega Corporation) was used to assess the activities of firefly and Renilla luciferase, and the firefly luciferase activity levels were adjusted to Renilla luciferase.

Bioinformatics analysis

The HAT1 protein expression data from healthy human tissue and OC tissues was derived from antibody-based protein profiling using immunohistochemistry from the Human Protein Atlas database (https://www.proteinatlas.org/). The mRNA and protein expression levels of HAT1 in healthy human tissue (n=133) and OC tissues (n=374) were analyzed using The Cancer Genome Atlas (TCGA data (https://www.cancer.gov/ccg/research/genome-sequencing/tcga) through the TNMplot analysis platform (tnmplot.com/analysis/), the University of Alabama at Birmingham Cancer data analysis Portal (UALCAN) and Clinical Proteomic Tumor Analysis Consortium (ualcan.path.uab.edu/index.html; gdc.cancer.gov/about-gdc/contributed-genomic-data-cancer-research/clinical-proteomic-tumor-analysis-consortium-cptac). Survival analysis was performed using the Kaplan-Meier method and log-rank testing to analyze the prognostic value of the HAT1 gene (kmplot.com/analysis/; http://www.gsea-msigdb.org/gsea/index.jsp); the patient samples were split into two groups according to various quantile expressions of HAT1 (Cutoff value used in analysis: 1383), and the hazard ratios with 95% confidence intervals and log-rank P-values were calculated; the Kaplan-Meier survival plots with the number at risk, hazard ratios and log-rank P-values were obtained using the Kaplan-Meier plotter website. The LinkedOmics database was used to perform HAT1 gene and pathway analysis using both over-representation analysis and gene set enrichment analysis (linkedomics.org; cutoff value, 2000). The binding regions of FOXA1 in HAT1 promoter region were analyzed by Cistrome DB (cistrome.org/db/#/) with the public chromatin immunoprecipitation-sequencing data downloaded from the Gene Expression Omnibus (GEO) datasets GSM 1538430/803461/1858655. The binding peaks of FOXA1 in HAT1 promoter indicated that FOXA1 transcriptionally regulated HAT1 expression. The cBioPortal database (cbioportal.org/) was used to analyze the correlation between FOXA1 and HAT1 through Pearson's correlation coefficient. Transcriptional factor FOXA1 was enriched at the HAT1 promoter region as suggested by Cistrome DB. The association between HAT1 and regulatory proteins of DNA replication and pyrimidine metabolism pathways in OC tissues were analyzed through cBioPortal database (https://www.cbioportal.org/).

Statistical analysis

Experiments were repeated at least three times and data are presented as the mean ± SD, with results analyzed using GraphPad 8.0 software (Dotmatics). The differences between two groups were analyzed via unpaired Student's t-test. HAT1 expression levels between the OC cell lines and the normal IOSE386 cell line were analyzed with one-way ANOVA and Dunnett's post hoc test. The significance of differences in HAT1 expression in the bioinformatics analysis was determined using Welch's t-test, or one-way ANOVA with Dunnett's post hoc test. P<0.05 was considered to indicate a statistically significant difference.

Results

Expression levels of HAT1 are upregulated in OC tissues and higher HAT1 levels indicate poor prognosis

HAT1 was the first HAT) discovered; however, its biological functions together with the cell mechanisms involved in cancer progression are poorly characterized. To identify the potential role of HAT1 in OC, the immunohistochemistry staining results of HAT1 in OC tissue were analyzed in the Human Protein Atlas database, and HAT1 staining was found to be higher in OC compared with in healthy human specimens (Fig. 1A). Through TCGA) database analysis, the OC tissues were found to have higher expression levels of HAT1 compared with healthy human tissues (Fig. 1B). In addition, HAT1 protein levels were increased in OC samples (Fig. 1C) and were associated with tumor grades (Fig. 1D). The association of HAT1 with overall survival was analyzed using the Kaplan-Meier plotter website and the results showed that higher HAT1 levels indicated shorter survival rates (Fig. 1E). These findings suggested that HAT1 may act as a tumor promoter in OC.

HAT1 levels are upregulated in human
ovarian cancer tissue and higher HAT1 levels indicate a poor
prognosis. (A) Representative images depicting the representative
immunostaining of HAT1 in ovarian cancer tissues and healthy human
tissues in The Human Protein Altas database. Scale bar=200 µm. (B)
Analysis of the TNMplot database showed that HAT1 was highly
expressed in ovarian cancer tissues compared with healthy controls.
(C) Protein levels of HAT1 in ovarian cancer and healthy human
samples from healthy controls were analyzed in the CPTAC database.
(D) HAT1 protein levels were upregulated with tumor grade in
ovarian cancer tissues. (E) Higher HAT1 levels were associated with
an unfavorable prognosis in patients with ovarian cancer. Data are
presented as the mean ± SD. **P<0.01. HAT1, histone
acetyltransferase 1; CPTAC, Clinical Proteomic Tumor Analysis
Consortium; HR, hazard ratio.

Figure 1.

HAT1 levels are upregulated in human ovarian cancer tissue and higher HAT1 levels indicate a poor prognosis. (A) Representative images depicting the representative immunostaining of HAT1 in ovarian cancer tissues and healthy human tissues in The Human Protein Altas database. Scale bar=200 µm. (B) Analysis of the TNMplot database showed that HAT1 was highly expressed in ovarian cancer tissues compared with healthy controls. (C) Protein levels of HAT1 in ovarian cancer and healthy human samples from healthy controls were analyzed in the CPTAC database. (D) HAT1 protein levels were upregulated with tumor grade in ovarian cancer tissues. (E) Higher HAT1 levels were associated with an unfavorable prognosis in patients with ovarian cancer. Data are presented as the mean ± SD. **P<0.01. HAT1, histone acetyltransferase 1; CPTAC, Clinical Proteomic Tumor Analysis Consortium; HR, hazard ratio.

HAT1 levels are increased in OC cell lines

To determine the role of HAT1 in OC development, the mRNA expression levels of HAT1 were assessed in different OC cell lines: A2780, HEY, OVCAR3 and SKOV3. The results showed that HAT1 expression levels were generally upregulated in OC cell lines compared with normal IOSE386 cell line, especially in HEY and SKOV3 cells, which were used in subsequent study (Fig. 2A). Similarly, the protein levels of HAT1 were increased in OC cells (Fig. 2B). Thus, HAT1 may induce ovarian tumorigenesis.

HAT1 expression levels are
upregulated in human ovarian cancer cell lines. (A) Relative
expression levels of HAT1 were detected by reverse
transcription-quantitative PCR in A2780, HEY, OVCAR3, SKOV3 and
normal IOSE386 cell lines. (B) Western blot analysis of HAT1
expression in A2780, HEY, OVCAR3, SKOV3 and IOSE386 cell lines.
****P<0.0001 vs. IOSE386. HAT1, histone acetyltransferase 1.

Figure 2.

HAT1 expression levels are upregulated in human ovarian cancer cell lines. (A) Relative expression levels of HAT1 were detected by reverse transcription-quantitative PCR in A2780, HEY, OVCAR3, SKOV3 and normal IOSE386 cell lines. (B) Western blot analysis of HAT1 expression in A2780, HEY, OVCAR3, SKOV3 and IOSE386 cell lines. ****P<0.0001 vs. IOSE386. HAT1, histone acetyltransferase 1.

FOXA1 transcriptionally regulates HAT1 expression in OC

TFs are important for the modulation of various signaling pathways associated with cell homeostasis and disease conditions (15). To identify the upstream regulator of HAT1 in OC, the TF database was used to predict the potential TFs of HAT1 and it was revealed that the TF FOXA1 may regulate HAT1. Among cancer-related TFs, FOXA1 is an important molecule that regulates multiple aspects of cancer cells, including cell proliferation and cancer metastasis. The potential FOXA1 binding sites of the HAT1 promoter region were then analyzed through the UCSC Genome Browser and JASPAR databases (Fig. 3A). Subsequently, WT and MUT HAT1 luciferase reporter plasmids were constructed (Fig. 3A). Overexpression of FOXA1 could promote the activities of HAT1 WT plasmids, as determined through the luciferase reporter assay; however, this was not observed with HAT1 MUT plasmids (Fig. 3B and C). Compared with the vector control, overexpression of FOXA1 significantly increased HAT1 expression levels (Fig. 3D). In addition, the cBioPortal database was adopted to analyze the correlation between FOXA1 and HAT1, and a positive association was observed between FOXA1 and HAT1 levels in OC tissues from TCGA data in the cBioPortal database (Fig. 3E). FOXA1 expression levels were also upregulated in OC tissues as suggested in TCGA database (Fig. 3F). Subsequently, the chromatin immunoprecipitation-sequencing data conveyed that FOXA1 was enriched in the promoter region of HAT1 (Fig. 3G). In conclusion, these data showed that FOXA1 functioned as a TF of HAT1.

FOXA1 transcriptionally regulates
HAT1 expression levels. (A) FOXA1 binding sites in the HAT1
promoter region were predicted using the JASPAR database. MUT
constructs were generated at the binding sequence regions as
indicated. (B) The expression levels of FOXA1 in HEY cells
transfected with an empty or FOXA1 vector. (C) HEY cells were
transfected with pmirGLO reporter vectors containing either WT or
MUT plasmids alongside an empty or FOXA1 vector. Luciferase
activities were determined 24 h after transfection. (D)
Overexpression of FOXA1 induced HAT1 expression levels. (E)
cBioPortal database was adopted to analyze the correlation between
FOXA1 and HAT1. Pearson's rank correlation between HAT1 and FOXA1
was analyzed in ovarian cancer tissues. (F) FOXA1 levels in ovarian
cancer tissues from TCGA were determined using TNMplot database.
(G) FOXA1 was enriched at the HAT1 promoter region as suggested by
Cistrome DB. Data are presented as the mean ± SD of three
replicates. **P<0.01 vs. vector or as indicated; ns, not
significant. ChIP-seq, chromatin immunoprecipitation sequencing;
FOXA1, forkhead box protein A1; HAT1, histone acetyltransferase 1;
WT, wild-type; MUT, mutant.

Figure 3.

FOXA1 transcriptionally regulates HAT1 expression levels. (A) FOXA1 binding sites in the HAT1 promoter region were predicted using the JASPAR database. MUT constructs were generated at the binding sequence regions as indicated. (B) The expression levels of FOXA1 in HEY cells transfected with an empty or FOXA1 vector. (C) HEY cells were transfected with pmirGLO reporter vectors containing either WT or MUT plasmids alongside an empty or FOXA1 vector. Luciferase activities were determined 24 h after transfection. (D) Overexpression of FOXA1 induced HAT1 expression levels. (E) cBioPortal database was adopted to analyze the correlation between FOXA1 and HAT1. Pearson's rank correlation between HAT1 and FOXA1 was analyzed in ovarian cancer tissues. (F) FOXA1 levels in ovarian cancer tissues from TCGA were determined using TNMplot database. (G) FOXA1 was enriched at the HAT1 promoter region as suggested by Cistrome DB. Data are presented as the mean ± SD of three replicates. **P<0.01 vs. vector or as indicated; ns, not significant. ChIP-seq, chromatin immunoprecipitation sequencing; FOXA1, forkhead box protein A1; HAT1, histone acetyltransferase 1; WT, wild-type; MUT, mutant.

HAT1 is involved in DNA replication and pyrimidine metabolism pathways

To explore the regulatory pathways associated with HAT1 expression levels, OC data from TCGA database were downloaded and analyzed, before being separated into HAT1 high- and low-expression groups section. The LinkedOmics database suggested that differentially expressed genes of HAT1 were enriched in several biological processes, such as ‘DNA replication’, ‘pyrimidine metabolism’ and ‘RNA transport’ (Fig. 4A). The gene set enrichment analysis results showed that the HAT1 high-expression group was positively associated with the ‘DNA replication’ and ‘pyrimidine metabolism’ pathways (Fig. 4B). Subsequently, the association between HAT1 and regulatory proteins of DNA replication and pyrimidine metabolism pathways in OC tissues were analyzed, and the expression levels of HAT1 were revealed to be positively correlated with proliferating cell nuclear antigen (PCNA), replication protein A1 and DNA polymerase α catalytic subunit from the cell cycle pathway (Fig. 4C). Furthermore, HAT1 levels were positively correlated with thymidine kinase 1, ribonucleoside-diphosphate reductase subunit M (RRM)1 and RRM2, which are involved in the pyrimidine metabolism pathway (Fig. 4D).

Relationship between HAT1 and DNA
replication and pyrimidine metabolism pathways. (A) LinkedOmics
database suggested that differentially expressed genes of HAT1 were
enriched in several biological processes, such as ‘DNA
replication’, ‘pyrimidine metabolism’ and ‘RNA transport’. (B) Gene
set enrichment analysis revealed that genes altered by HAT1 were
positively associated with ‘DNA replication’ as well as ‘pyrimidine
metabolism’ pathways. (C) cBioPortal database was used to analyze
the correlations between HAT1 and regulatory proteins of the DNA
replication pathway. Pearson's rank correlation between HAT1 and
DNA replication-related proteins (PCNA, RPA1 and POLA1) was
analyzed in ovarian cancer tissues. (D) Pearson's rank correlation
between HAT1 and pyrimidine metabolism-related proteins (TK1, RRM1
and RRM2) was analyzed in ovarian cancer tissues as suggested in
the cBioPortal database. HAT1, histone acetyltransferase 1; PCNA,
proliferating cell nuclear antigen; RPA1, replication protein A1;
POLA1, DNA polymerase α catalytic subunit; TK1, thymidine kinase 1;
RRM, ribonucleoside-diphosphate reductase subunit M; NES,
normalized enrichment score; FDR, false discovery rate.

Figure 4.

Relationship between HAT1 and DNA replication and pyrimidine metabolism pathways. (A) LinkedOmics database suggested that differentially expressed genes of HAT1 were enriched in several biological processes, such as ‘DNA replication’, ‘pyrimidine metabolism’ and ‘RNA transport’. (B) Gene set enrichment analysis revealed that genes altered by HAT1 were positively associated with ‘DNA replication’ as well as ‘pyrimidine metabolism’ pathways. (C) cBioPortal database was used to analyze the correlations between HAT1 and regulatory proteins of the DNA replication pathway. Pearson's rank correlation between HAT1 and DNA replication-related proteins (PCNA, RPA1 and POLA1) was analyzed in ovarian cancer tissues. (D) Pearson's rank correlation between HAT1 and pyrimidine metabolism-related proteins (TK1, RRM1 and RRM2) was analyzed in ovarian cancer tissues as suggested in the cBioPortal database. HAT1, histone acetyltransferase 1; PCNA, proliferating cell nuclear antigen; RPA1, replication protein A1; POLA1, DNA polymerase α catalytic subunit; TK1, thymidine kinase 1; RRM, ribonucleoside-diphosphate reductase subunit M; NES, normalized enrichment score; FDR, false discovery rate.

HAT1 induces OC cell viability and colony formation

To confirm the tumor-promoting role of HAT1, HEY and SKOV3 OC cell lines were used to knockdown HAT1 expression (Fig. 5A). As determined by CCK-8 assay, HAT1 knockdown markedly suppressed cell proliferation (Fig. 5B). In the present study, cells were also treated with the HAT1 inhibitor JG-2016, and the results showed that JG-2016 significantly inhibited cell proliferation (Fig. 5C). In addition, inhibition of HAT1 reduced colony formation activities in HEY and SKOV3 cell lines (Fig. 5D). Thus, the present results indicated that HAT1 contributed to tumorigenic growth.

Inhibition of HAT1 suppresses cell
viability and colony formation in vitro. (A) Western blot
analysis of HAT1 expression in HEY and SKOV3 cells transfected with
HAT1 siRNA or NC siRNA. (B) Cell Counting Kit-8 assays were
performed to determine cell viability after HAT1 was knocked down
in HEY and SKOV3 cells. (C) HEY and SKOV3 cells were treated with
HAT1 inhibitor JG-2016 for 72 h, which significantly inhibited cell
viability. (D) Knockdown of HAT1 decreased the colony formation
capacity of HEY and SKOV3 cells. Data are presented as the mean ±
SD of three replicates. *P<0.05 and **P<0.01 vs. siNC in
figure 5B and DMSO group in figure 5C. HAT1, histone
acetyltransferase 1; si, small interfering; NC, negative
control.

Figure 5.

Inhibition of HAT1 suppresses cell viability and colony formation in vitro. (A) Western blot analysis of HAT1 expression in HEY and SKOV3 cells transfected with HAT1 siRNA or NC siRNA. (B) Cell Counting Kit-8 assays were performed to determine cell viability after HAT1 was knocked down in HEY and SKOV3 cells. (C) HEY and SKOV3 cells were treated with HAT1 inhibitor JG-2016 for 72 h, which significantly inhibited cell viability. (D) Knockdown of HAT1 decreased the colony formation capacity of HEY and SKOV3 cells. Data are presented as the mean ± SD of three replicates. *P<0.05 and **P<0.01 vs. siNC in figure 5B and DMSO group in figure 5C. HAT1, histone acetyltransferase 1; si, small interfering; NC, negative control.

HAT1 expression is associated with CDK2, CDK4 and cyclin E levels

Human HEY OC cells were transfected with negative control and HAT1 siRNAs. These cells were then stained with an EdU probe, which functions as a cell proliferation marker, and the results showed that HAT1 knockdown significantly suppressed EdU activities (Fig. 6A). The cell cycle regulatory pathway integrates into other hallmarks of cancer, including metabolism remodeling and immune escape, and promotes tumorigenesis (16). In the present study, inhibition of HAT1 suppressed cell cycle progression and reduced the percentage of cells at S phase (Fig. 6B). Cell cycle-related proteins levels were then analyzed and it was revealed that suppression of HAT1 decreased CDK2, CDK4 and cyclin E levels (Fig. 6C). Furthermore, positive associations were shown between HAT1 and CDK2, CDK4 and cyclin E levels in OC tissues from TCGA data obtained from the cBioPortal database (Fig. 6D). Thus, HAT1 was suggested to regulate the cell cycle pathway in OC.

HAT1 knockdown inhibits cell
proliferation. (A) EdU assays of HEY cells were performed showing
that suppression of HAT1 attenuated cell proliferation activities.
Scale bar, 20 µm. (B) Cell cycle analysis was carried out on the
HAT1 knockdown and siNC cell lines, and the distribution of
G0/G1, S and G2/M percentages were
analyzed. (C) Western blot analysis of cell cycle-related protein
expression after knockdown of HAT1. (D) Pearson's rank correlation
between HAT1 and CDK2, CDK4 and cyclin E was analyzed in ovarian
cancer tissues. Data are presented as the mean ± SD of three
replicates. **P<0.01 vs. siNC. HAT1, histone acetyltransferase
1; si, small interfering; NC, negative control; CDK,
cyclin-dependent kinase.

Figure 6.

HAT1 knockdown inhibits cell proliferation. (A) EdU assays of HEY cells were performed showing that suppression of HAT1 attenuated cell proliferation activities. Scale bar, 20 µm. (B) Cell cycle analysis was carried out on the HAT1 knockdown and siNC cell lines, and the distribution of G0/G1, S and G2/M percentages were analyzed. (C) Western blot analysis of cell cycle-related protein expression after knockdown of HAT1. (D) Pearson's rank correlation between HAT1 and CDK2, CDK4 and cyclin E was analyzed in ovarian cancer tissues. Data are presented as the mean ± SD of three replicates. **P<0.01 vs. siNC. HAT1, histone acetyltransferase 1; si, small interfering; NC, negative control; CDK, cyclin-dependent kinase.

Discussion

OC has become one of the most frequently diagnosed cancers in female patients worldwide, with an expected 313,000 cases and 207,000 mortalities in 2020, ranking as the third highest cause of death among gynecological cancers (17). The high mortality rate is partly due to the heterogeneity of the tumor combined with the absence of symptoms at the early stages of OC, which limits early diagnosis, resulting in most patients presenting with advanced stages of OC when identified (18). Histone modifications, as major chromatin regulators, play an important role in the etiology of numerous diseases, especially cancer (19). Acetylation, specifically lysine acetylation, is a dynamic epigenetic process with a notable role in cell cycle, cell survival or to cellular processes, which capable of targeting both histone and non-histone proteins (16,20). Epigenetic modification via acetylation regulates gene transcriptional processes and protein stability, activity and localization (21). HAT1 was the first identified HAT; however, its biological significance remains to be elucidated. Several studies have highlighted the importance of HAT1 in regulating cellular processes that are altered in a variety of diseases, including cancer (13,22,23). HAT1 protein knockdown or knockout in various pre-clinical models has conveyed that HAT1 is likely a new potential therapeutic target in OC treatment (13,22,23). In the present study, the mRNA and protein expression levels of HAT1 were found to be elevated in OC tissues, and elevated HAT1 levels were associated with shorter survival rate.

FOXA1 belongs to the FOXA TF family and has a forkhead (or winged helix) DNA-binding domain of ~100 amino acids. FOXA1 functions as a TF and attaches to the chromosome to induce nucleosome remodeling, which allows other TFs to bind to the chromosome and perform tissue-specific transcriptional programs (24–28). FOXA1 can be pro- or antitumorigenic in various human malignancies (29). Up to 80% of estrogen receptor-positive breast cancer cases are FOXA1-positive, and FOXA1 expression is linked with improved prognosis (29). Furthermore, FOXA1 upregulation has been shown to be inversely linked with interferon signature and activated tumor-specific CD8+T cells in patients with prostate cancer, and to increase cancer immunoresistance and resistance to chemotherapy in nude mice and patients with prostate cancer (30). FOXA1 acts as a tumor suppressor during its development of endometrial cancer (31). In the present study, the TF FOXA1 was shown to regulate HAT1 levels, and there was a positive association between FOXA1 and HAT1 in OC.

The functional importance of HAT1 has been investigated in a number of cell models to improve understanding of its biological function. Previous investigation has suggested that HAT1 is briefly localized on chromatin, adjacent to DNA replication sites, where it plays a regulatory role in replication fork stalling (32). According to additional research, the loss of HAT1 enhances genomic instability, making cells vulnerable to DNA double-strand breaks (33). The significance of HAT1 in replication and chromatin assembly has previously been shown (34).

In the present study, the expression levels of HAT1 were positively correlated with PCNA in the DNA replication pathway. PCNA is the eukaryotic DNA sliding clamp that interacts with cell cycle proteins, including cyclins and CDKs, involved in DNA replication (35). A previous study has shown that PCNA is a key factor in DNA replication during the cell cycle and is directly associated with prostate tissue proliferation (36). PCNA inhibition induces S-phase arrest of human prostatic epithelial cancer cell lines (37). PCNA also interacts with the cyclin-CDK complex in several eukaryotic cells (38). It has been noted that CDK2 could directly interact with PCNA in numerous cells, supporting the linkage between cell cycle checkpoints and PCNA (39). Thus, PCNA is considered an accessory protein in cellular processes, contributing to cell division, DNA replication and various cell cycle controls. In the present study, HAT1 was suggested to regulate the cell cycle pathway in OC. The cell cycle is a series of interconnected events that allow the cell to continue growing and proliferating. Cell cycle arrest serves as a survival mechanism that enables cancer cells to compensate for their own DNA that has been damaged. CDKs are key components of the cell cycle machinery that, when activated, enable the cell to move from one phase to the next (40). Cyclins positively regulate CDKs, whereas naturally generated CDK inhibitors (CDKIs) adversely regulate them. Cancer is characterized by cell cycle dysregulation, with cells upregulating cyclins or lacking CDKIs continuing to grow abnormally (41). The cell cycle also protects the cell against DNA damage. The present research conveyed that suppression of HAT1 decreased CDK2, CDK4 and cyclin E levels; thus, HAT1 may be a potential therapeutic target in OC.

Notably, findings have revealed the importance of HAT1 in a variety of physiological processes, encouraging additional research to determine its biological function (12,42). HAT1 has been linked to a variety of cellular functions, including proliferation, DNA replication and cellular metabolism (12,42). The JG-2016 compound has shown relative specificity toward HAT1 compared with other acetyltransferases, suppressing the proliferation of human cancer cell lines and interfering with tumor growth (43). The present study is, to the best of our knowledge, the first report of a small-molecule inhibitor of the HAT1 enzyme complex and represents a step toward targeting the HAT1 pathway for cancer therapy. In the present study, the inhibition of HAT1 was discovered to promote cell cycle arrest, and reduce CDK2, CDK4 and cyclin E levels in OC cells. Thus, the present work highlighted the importance of FOXA1 as an upstream regulator mediating HAT1 expression levels and HAT1 as a potential oncogene in OC. Therefore, specifically inhibiting HAT1 expression may provide a unique strategy for OC treatment.

Acknowledgements

Not applicable.

Funding

The present work was supported by the Science and Technology Research Project of Henan province (grant no. LHGJ20220532).

Availability of data and materials

The data generated in the present study may be requested from the corresponding author.

Authors' contributions

XH conceptualized and designed the study, performed experiments and wrote the manuscript. JLi, YZ, LZ, LY and XO developed methodology, analyzed data and performed experiments. LL and JLiu performed experiments, analyzed data and confirm the authenticity of all the raw data. All authors have read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Obermair A, Beale P, Scott CL, Beshay V, Kichenadasse G, Simcock B, Nicklin J, Lee YC, Cohen P and Meniawy T: Insights into ovarian cancer care: report from the ANZGOG ovarian cancer webinar series 2020. J Gynecol Oncol. 32:e952021. View Article : Google Scholar : PubMed/NCBI

2 

Xia CF, Dong XS, Li H, Cao M, Sun D, He S, Yang F, Yan X, Zhang S, Li N and Chen W: Cancer statistics in China and United States, 2022: Profiles, trends, and determinants. Chin Med J (Engl). 135:584–590. 2022. View Article : Google Scholar : PubMed/NCBI

3 

Garrido MP, Fredes AN, Lobos-Gonzalez L, Valenzuela-Valderrama M, Vera DB and Romero C: Current treatments and new possible complementary therapies for epithelial ovarian cancer. Biomedicines. 10:772021. View Article : Google Scholar : PubMed/NCBI

4 

Jayde V, White K and Blomfield P: Symptoms and diagnostic delay in ovarian cancer: A summary of the literature. Contemp Nurse. 34:55–65. 2009. View Article : Google Scholar : PubMed/NCBI

5 

Konstantinopoulos PA and Matulonis UA: Clinical and translational advances in ovarian cancer therapy. Nat Cancer. 4:1239–1257. 2023. View Article : Google Scholar : PubMed/NCBI

6 

Orr JA and Hamilton PW: Histone acetylation and chromatin pattern in cancer. A review. Anal Quant Cytol Histol. 29:17–31. 2007.PubMed/NCBI

7 

Makowski AM, Dutnall RN and Annunziato AT: Effects of acetylation of histone H4 at lysines 8 and 16 on activity of the Hat1 histone acetyltransferase. J Biol Chem. 276:43499–43502. 2001. View Article : Google Scholar : PubMed/NCBI

8 

Jin X, Tian SH and Li PP: Histone acetyltransferase 1 promotes cell proliferation and induces cisplatin resistance in hepatocellular carcinoma. Oncol Res. 25:939–946. 2017. View Article : Google Scholar : PubMed/NCBI

9 

Klett-Mingo JI, Pinto-Díez C, Cambronero-Plaza J, Carrión-Marchante R, Barragán-Usero M, Pérez-Morgado MI, Rodríguez-Martín E, Toledo-Lobo MV, González VM and Martín ME: Potential therapeutic use of aptamers against HAT1 in lung cancer. Cancers (Basel). 15:2272022. View Article : Google Scholar : PubMed/NCBI

10 

Xue L, Hou J, Wang Q, Yao L, Xu S and Ge D: RNAi screening identifies HAT1 as a potential drug target in esophageal squamous cell carcinoma. Int J Clin Exp Patho. 7:3898–3907. 2014.PubMed/NCBI

11 

Sarkar T, Dhar S, Chakraborty D, Pati S, Bose S, Panda AK, Basak U, Chakraborty S, Mukherjee S, Guin A, et al: FOXP3/HAT1 axis controls treg infiltration in the tumor microenvironment by inducing CCR4 expression in breast cancer. Front Immunol. 13:7405882022. View Article : Google Scholar : PubMed/NCBI

12 

Sun Y, Ren D, Zhou Y, Shen J, Wu H and Jin X: Histone acetyltransferase 1 promotes gemcitabine resistance by regulating the PVT1/EZH2 complex in pancreatic cancer. Cell Death Dis. 12:8782021. View Article : Google Scholar : PubMed/NCBI

13 

Yang G, Yuan Y, Yuan H, Wang J, Yun H, Geng Y, Zhao M, Li L, Weng Y, Liu Z, et al: Histone acetyltransferase 1 is a succinyltransferase for histones and non-histones and promotes tumorigenesis. Embo Rep. 22:e509672021. View Article : Google Scholar : PubMed/NCBI

14 

Livak KJ and Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(−Delta Delta C(T)) method. Methods. 25:402–408. 2001. View Article : Google Scholar : PubMed/NCBI

15 

Weidemüller P, Kholmatov M, Petsalaki E and Zaugg JB: Transcription factors: Bridge between cell signaling and gene regulation. Proteomics. 21:e20000342021. View Article : Google Scholar : PubMed/NCBI

16 

Hanahan D: Hallmarks of cancer: New dimensions. Cancer Discov. 12:31–46. 2022. View Article : Google Scholar : PubMed/NCBI

17 

Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A and Bray F: Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 71:209–249. 2021.PubMed/NCBI

18 

Webb PM and Jordan SJ: Epidemiology of epithelial ovarian cancer. Best Pract Res Clin Obstet Gynaecol. 41:3–14. 2017. View Article : Google Scholar : PubMed/NCBI

19 

Shirvaliloo M: The landscape of histone modifications in epigenomics since 2020. Epigenomics. 14:1465–1477. 2022. View Article : Google Scholar : PubMed/NCBI

20 

Shvedunova M and Akhtar A: Modulation of cellular processes by histone and non-histone protein acetylation. Nat Rev Mol Cell Bio. 23:329–349. 2022. View Article : Google Scholar : PubMed/NCBI

21 

Sun LC, Zhang HF and Gao P: Metabolic reprogramming and epigenetic modifications on the path to cancer. Protein Cell. 13:877–919. 2022. View Article : Google Scholar : PubMed/NCBI

22 

Fan P, Zhao J, Meng Z, Wu H, Wang B, Wu H and Jin X: Overexpressed histone acetyltransferase 1 regulates cancer immunity by increasing programmed death-ligand 1 expression in pancreatic cancer. J Exp Clin Cancer Res. 38:472019. View Article : Google Scholar : PubMed/NCBI

23 

Gruber JJ, Geller B, Lipchik AM, Chen J, Salahudeen AA, Ram AN, Ford JM, Kuo CJ and Snyder MP: HAT1 coordinates histone production and acetylation via H4 promoter binding. Mol Cell. 75:711–724.e5. 2019. View Article : Google Scholar : PubMed/NCBI

24 

Carroll JS, Liu XS, Brodsky AS, Li W, Meyer CA, Szary AJ, Eeckhoute J, Shao W, Hestermann EV, Geistlinger TR, et al: Chromosome-wide mapping of estrogen receptor binding reveals long-range regulation requiring the forkhead protein FoxA1. Cell. 122:33–43. 2005. View Article : Google Scholar : PubMed/NCBI

25 

Cirillo LA, Lin FR, Cuesta I, Friedman D, Jarnik M and Zaret KS: Opening of compacted chromatin by early developmental transcription factors HNF3 (FoxA) and GATA-4. Mol Cell. 9:279–289. 2002. View Article : Google Scholar : PubMed/NCBI

26 

Cirillo LA and Zaret KS: An early developmental transcription factor complex that is more stable on nucleosome core particles than on free DNA. Mol Cell. 4:961–969. 1999. View Article : Google Scholar : PubMed/NCBI

27 

Laganière J, Deblois G, Lefebvre C, Bataille AR, Robert F and Giguère V: From the Cover: Location analysis of estrogen receptor alpha target promoters reveals that FOXA1 defines a domain of the estrogen response. Proc Natl Acad Sci USA. 102:11651–11656. 2005. View Article : Google Scholar : PubMed/NCBI

28 

Zaret K: Developmental competence of the gut endoderm: Genetic potentiation by GATA and HNF3/fork head proteins. Dev Biol. 209:1–10. 1999. View Article : Google Scholar : PubMed/NCBI

29 

Albergaria A, Paredes J, Sousa B, Milanezi F, Carneiro V, Bastos J, Costa S, Vieira D, Lopes N, Lam EW, et al: Expression of FOXA1 and GATA-3 in breast cancer: The prognostic significance in hormone receptor-negative tumours. Breast Cancer Res. 11:R402009. View Article : Google Scholar : PubMed/NCBI

30 

He Y, Wang L, Wei T, Xiao YT, Sheng H, Su H, Hollern DP, Zhang X, Ma J, Wen S, et al: FOXA1 overexpression suppresses interferon signaling and immune response in cancer. J Clin Invest. 131:e1470252021. View Article : Google Scholar : PubMed/NCBI

31 

Abe Y, Ijichi N, Ikeda K, Kayano H, Horie-Inoue K, Takeda S and Inoue S: Forkhead box transcription factor, forkhead box A1, shows negative association with lymph node status in endometrial cancer, and represses cell proliferation and migration of endometrial cancer cells. Cancer Sci. 103:806–812. 2012. View Article : Google Scholar : PubMed/NCBI

32 

Agudelo Garcia PA, Lovejoy CM, Nagarajan P, Park D, Popova LV, Freitas MA and Parthun MR: Histone acetyltransferase 1 is required for DNA replication fork function and stability. J Biol Chem. 295:8363–8373. 2020. View Article : Google Scholar : PubMed/NCBI

33 

Yang X, Li L, Liang J, Shi L, Yang J, Yi X, Zhang D, Han X, Yu N and Shang Y: Histone acetyltransferase 1 promotes homologous recombination in DNA repair by facilitating histone turnover. J Biol Chem. 288:18271–18282. 2013. View Article : Google Scholar : PubMed/NCBI

34 

Agudelo Garcia PA, Hoover ME, Zhang P, Nagarajan P, Freitas MA and Parthun MR: Identification of multiple roles for histone acetyltransferase 1 in replication-coupled chromatin assembly. Nucleic Acids Res. 45:9319–9335. 2017. View Article : Google Scholar : PubMed/NCBI

35 

Kelman Z and O'Donnell M: Structural and functional similarities of prokaryotic and eukaryotic DNA polymerase sliding clamps. Nucleic Acids Res. 23:3613–3620. 1995. View Article : Google Scholar : PubMed/NCBI

36 

Taftachi R, Ayhan A, Ekici S, Ergen A and Ozen H: Proliferating-cell nuclear antigen (PCNA) as an independent prognostic marker in patients after prostatectomy: A comparison of PCNA and Ki-67. BJU Int. 95:650–654. 2005. View Article : Google Scholar : PubMed/NCBI

37 

Tan Z, Wortman M, Dillehay KL, Seibel WL, Evelyn CR, Smith SJ, Malkas LH, Zheng Y, Lu S and Dong Z: Small-molecule targeting of proliferating cell nuclear antigen chromatin association inhibits tumor cell growth. Mol Pharmacol. 81:811–819. 2012. View Article : Google Scholar : PubMed/NCBI

38 

Xiong Y, Zhang H and Beach D: D type cyclins associate with multiple protein kinases and the DNA replication and repair factor PCNA. Cell. 71:505–514. 1992. View Article : Google Scholar : PubMed/NCBI

39 

Koundrioukoff S, Jónsson ZO, Hasan S, de Jong RN, van der Vliet PC, Hottiger MO and Hübscher U: A direct interaction between proliferating cell nuclear antigen (PCNA) and Cdk2 targets PCNA-interacting proteins for phosphorylation. J Biol Chem. 275:22882–22887. 2000. View Article : Google Scholar : PubMed/NCBI

40 

Suski JM, Braun M, Strmiska V and Sicinski P: Targeting cell-cycle machinery in cancer. Cancer Cell. 39:759–778. 2021. View Article : Google Scholar : PubMed/NCBI

41 

Cavalu S, Abdelhamid AM, Saber S, Elmorsy EA, Hamad RS, Abdel-Reheim MA, Yahya G and Salama MM: Cell cycle machinery in oncology: A comprehensive review of therapeutic targets. FASEB J. 38:e237342024. View Article : Google Scholar : PubMed/NCBI

42 

Zhou P, Peng X, Zhang K, Cheng J, Tang M, Shen L, Zhou Q, Li D and Yang L: HAT1/HDAC2 mediated ACSL4 acetylation confers radiosensitivity by inducing ferroptosis in nasopharyngeal carcinoma. Cell Death Dis. 16:1602025. View Article : Google Scholar : PubMed/NCBI

43 

Gaddameedi JD, Chou T, Geller BS, Rangarajan A, Swaminathan TA, Dixon D, Long K, Golder CJ, Vuong VA, Banuelos S, et al: Acetyl-click screening platform identifies small-molecule inhibitors of histone acetyltransferase 1 (HAT1). J Med Chem. 66:5774–5801. 2023. View Article : Google Scholar : PubMed/NCBI

Related Articles

  • Abstract
  • View
  • Download
  • Twitter
Copy and paste a formatted citation
Spandidos Publications style
Han X, Liu L, Li J, Zhi Y, Zhao L, Yuan L, Ouyang X and Liu J: Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle. Oncol Lett 30: 539, 2025.
APA
Han, X., Liu, L., Li, J., Zhi, Y., Zhao, L., Yuan, L. ... Liu, J. (2025). Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle. Oncology Letters, 30, 539. https://doi.org/10.3892/ol.2025.15285
MLA
Han, X., Liu, L., Li, J., Zhi, Y., Zhao, L., Yuan, L., Ouyang, X., Liu, J."Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle". Oncology Letters 30.6 (2025): 539.
Chicago
Han, X., Liu, L., Li, J., Zhi, Y., Zhao, L., Yuan, L., Ouyang, X., Liu, J."Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle". Oncology Letters 30, no. 6 (2025): 539. https://doi.org/10.3892/ol.2025.15285
Copy and paste a formatted citation
x
Spandidos Publications style
Han X, Liu L, Li J, Zhi Y, Zhao L, Yuan L, Ouyang X and Liu J: Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle. Oncol Lett 30: 539, 2025.
APA
Han, X., Liu, L., Li, J., Zhi, Y., Zhao, L., Yuan, L. ... Liu, J. (2025). Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle. Oncology Letters, 30, 539. https://doi.org/10.3892/ol.2025.15285
MLA
Han, X., Liu, L., Li, J., Zhi, Y., Zhao, L., Yuan, L., Ouyang, X., Liu, J."Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle". Oncology Letters 30.6 (2025): 539.
Chicago
Han, X., Liu, L., Li, J., Zhi, Y., Zhao, L., Yuan, L., Ouyang, X., Liu, J."Histone acetyltransferase 1 promotes ovarian cancer progression by regulating cell proliferation and the cell cycle". Oncology Letters 30, no. 6 (2025): 539. https://doi.org/10.3892/ol.2025.15285
Follow us
  • Twitter
  • LinkedIn
  • Facebook
About
  • Spandidos Publications
  • Careers
  • Cookie Policy
  • Privacy Policy
How can we help?
  • Help
  • Live Chat
  • Contact
  • Email to our Support Team