Next Article in Journal
Lack of Relationship between Fibrosis-Related Biomarkers and Cardiac Magnetic Resonance-Assessed Replacement and Interstitial Fibrosis in Dilated Cardiomyopathy
Previous Article in Journal
Effect of Differences in the Microbiome of Cyp17a1-Deficient Mice on Atherosclerotic Background
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Recurrent HBV Integration Targets as Potential Drivers in Hepatocellular Carcinoma

1
JBS Science, Inc., Doylestown, PA 18902, USA
2
The Baruch S. Blumberg Research Institute, Doylestown, PA 18902, USA
3
Department of Internal Medicine, National Cheng Kung University Medical College, Tainan 704, Taiwan
4
Department of Surgery, National Cheng Kung University Medical College, Tainan 704, Taiwan
*
Author to whom correspondence should be addressed.
Cells 2021, 10(6), 1294; https://doi.org/10.3390/cells10061294
Submission received: 22 April 2021 / Revised: 15 May 2021 / Accepted: 20 May 2021 / Published: 23 May 2021

Abstract

:
Chronic hepatitis B virus (HBV) infection is the major etiology of hepatocellular carcinoma (HCC), frequently with HBV integrating into the host genome. HBV integration, found in 85% of HBV-associated HCC (HBV–HCC) tissue samples, has been suggested to be oncogenic. Here, we investigated the potential of HBV–HCC driver identification via the characterization of recurrently targeted genes (RTGs). A total of 18,596 HBV integration sites from our in-house study and others were analyzed. RTGs were identified by applying three criteria: at least two HCC subjects, reported by at least two studies, and the number of reporting studies. A total of 396 RTGs were identified. Among the 28 most frequent RTGs, defined as affected in at least 10 HCC patients, 23 (82%) were associated with carcinogenesis and 5 (18%) had no known function. Available breakpoint positions from the three most frequent RTGs, TERT, MLL4/KMT2B, and PLEKHG4B, were analyzed. Mutual exclusivity of TERT promoter mutation and HBV integration into TERT was observed. We present an RTG consensus through comprehensive analysis to enable the potential identification and discovery of HCC drivers for drug development and disease management.

1. Introduction

Hepatocellular carcinoma (HCC) is the second leading cause of cancer deaths worldwide [1,2,3], and its poor prognosis is, in part, due to the lack of effective treatment options. A major etiology of this multifactorial disease is chronic hepatitis B virus (HBV) infection, which is associated with approximately 50% of HCC cases worldwide [4]. Current HBV–HCC screening guidelines vary regionally: the American Association for the Study of Liver Disease (AASLD), Canadian Association for the Study of the Liver (CASL), European Association for the Study of the Liver (EASL), Asian-Pacific Association for the Study of the Liver (APASL), and Latin-American Association for the Study of the Liver (ALEH) all have distinctive recommendations. Irrespective of the region, ultrasound with or without serum AFP screening is recommended every 6 months in HBV populations with risk factors that may include being of Asian descent, >40 years of age, cirrhosis, and PAGE-B score [5].
During infection, HBV can integrate into the host genome. It has been proposed that integration events mostly occur through non-homologous end joining (NHEJ) and micro-homologous recombination [6,7,8,9]. Although HBV DNA integration into the host genome is considered rare—with an estimate of one integration event per 10,000 HBV-infected hepatocytes [10]—integrated viral DNA has been reported in more than 85% of HBV-related HCC (HBV–HCC) cases, suggesting a significant association between HBV integration and hepatocarcinogenesis. Additionally, mechanisms of HBV integration in HCC carcinogenesis potentially vary from patient to patient, including the insertional mutagenesis of HCC-associated genes, induction of chromosomal instability, and continuous expression of viral proteins [11,12]. Therefore, understanding the impact of integrated HBV DNA on carcinogenesis may identify potential HCC driver genes as personalized biomarkers could paving the way for precision disease management in HBV–HCC patients.
With the advent of next-generation sequencing (NGS), thousands of HBV integration sites have been identified across the human genome. Although no host sequence preference or specificity [13,14,15,16,17,18] has been identified, integration can activate known HCC driver genes and has been reported in TERT, CCNE1, and MLL4 [19]. These, and other frequently affected genes, have become known as recurrently targeted genes (RTGs). Similar to the approach of identifying BRAF V600E driver mutations as recurrent hotspot mutations, here, we take advantage of the large number of reported integration sites from the literature and our in-house study to test the hypothesis that characterization of frequent RTGs can be a tool for the identification of HBV–HCC drivers.
In this study, we collected integration sites identified in HCC tumor tissue and defined RTGs. A total of 18,596 HBV integration sites, detected using PCR- and NGS-based approaches [5,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35], were analyzed. By characterizing the most frequent RTGs, we demonstrate the potential of identifying HCC drivers for HCC precision medicine and drug development.

2. Materials and Methods

2.1. Data Mining/Search Strategy

We searched the PubMed (1 January 2000–1 December 2020) database using Medical Subject Heading (MeSH) terms “hepatitis B virus”, “HBV integration”, and “hepatitis B integration sites” to identify studies that have reported HBV integration sites by either NGS- or PCR-based approaches. Additional studies were obtained by cross-referencing from the literature. We included only English-language studies with HCC subjects. We included all such studies that identified HBV integration sites using NGS-based approaches. Among studies using PCR-based methods, we only included those with sample sizes of 10 or more HCC patients. HBV integration sites identified by RNA-seq [7,8,20] were not included, because the expression of integrated sequences can be regulated by many host cellular factors and is not within the scope of this study. We filtered out repeated integration sites to ensure that each integration site was included only once, except for two studies that utilized different methods on overlapping samples [21,22]. A total of 33 reported studies in addition to our study are included, as summarized in Table S1.

2.2. In-House HCC Specimens and HBV Integration Analysis

Archived FFPE tumor tissue DNA (Table S2) [23,24] from stage I–IIIB patients (n = 32) was obtained from the National Cheng-Kung University Medical Center, Taiwan, and collected following the guidelines of the Institutional Review Board. An HBV enrichment NGS assay (JBS Science, Doylestown, PA, USA) was used to capture integration sites. Briefly, NGS libraries were generated and enriched for HBV sequences, and the enriched libraries were sequenced on the Illumina MiSeq platform (Penn State Hershey Genomics Sciences Facility at Penn State College of Medicine, Hershey, PA, USA). The NGS data were analyzed using ChimericSeq [25] to identify HBV–host junction sequences. Reference genomes NC_003977.1 (HBV) and GRCh38.p2 (human) were used. The closest genes within 150 kb of the HBV integration breakpoint were identified with ChimericSeq software. Genes were defined as in NCBI’s RefSeq database. Tailored junction-specific PCR-Sanger sequencing assays were designed to validate each HBV integration site of interest, identified by the HBV enrichment NGS assay.

2.3. TERT Promoter Mutation Analysis by PCR-Sanger Sequencing

HCC tissue DNA was used to amplify a 163 bp region (Chr5:1,295,151–1,295,313) of the TERT promoter by HotStart Plus Taq Polymerase (Qiagen, Valencia, CA, USA) with forward primer 5′-CAGCGCTGCCTGAAACTC-3′ and reverse primer 5′-GTCCTGCCCCTTCACCTT-3′. The PCR products were sequenced at the NAPCore Facility at the Children’s Hospital of Philadelphia (Philadelphia, PA, USA) and analyzed using the ClustalW software [26].

2.4. Identification of RTGs

To identify host genes that may have been affected by HBV DNA integration across all studies, we identified the closest gene within 150 kb of the integration event, the reported distance within which host genes can be impacted by integration [27,28]. For a gene to be classified as an RTG, it must have been identified as an HBV integration locus in two or more independent studies from at least two independent laboratories, each of which included at least two HCC patients with the gene affected in the tumor tissue. The requirement that a gene must have been identified in multiple laboratories served to minimize the issue of potential cross-contamination. The full list of identified RTGs can be provided upon request.

2.5. Gene Functional Enrichment Pathway Analysis

A total of 396 RTGs were subjected to enrichment pathway analysis using Enrichr (http://amp.pharm.mssm.edu/Enrichr (accessed on 30 January 2021)), to identify significantly (p < 0.05) enriched pathways as determined by gene ontology.

3. Results

3.1. Overview of the Studies for RTG Identification

The studies included in RTG identification are summarized in Table 1. Twenty-three studies utilized NGS, including our in-house study (Table S1), and 10 studies utilized PCR-based approaches for HBV integration identification. For each study, the sample size and the number and percentage of HCC tumors that had detectable integration sites are listed. We compiled a total of 18,596 HBV integration sites from tumor tissues of a combined 1310 HCC patients. In all, we found that 76% of tumor tissues (n = 1733) contained detectable integration sites. In the 12 studies that enriched for the whole HBV genome, on average, 72% (range 54–100%) of the tumors examined were found to have integrated HBV DNA (n = 880). In two studies using an HBV DR1-2 enrichment NGS assay, 65% [29] and 91% (our study) of tumors examined were positive for integrated HBV DNA.

3.2. Characteristics of HBV–HCC Patients with Integrated HBV DNA

To compare the general sociodemographic and clinical characteristics of the HBV–HCC population compiled in this study (age, gender, HBV genotype, and cirrhotic etiology (designated as “cirrhotic HCC”)) to those previously reported in the literature [4,58,59], we divided our study population into two groups based on the detectability of integrated HBV DNA (Table 2). Analysis of each parameter was performed where available. Overall, there were no significant differences between the two groups (with integrated DNA detectable vs. not detectable), including in the sex ratio. Of the three reported HBV genotypes, genotype C was the most frequently reported in the integration-positive patient cohort (75%), while the cohort with no detectable integration had only two patients with a genotype reported (both C). Interestingly, only 47% (7/15) of the integration-negative patients had cirrhosis, which is much lower than the 70–90% range previously reported in the general HBV–HCC population [4]. In the integration-positive cohort, 62% of HCC tumors were derived from a cirrhotic liver, which is also below the previously reported range.
Lastly, because it is known that HBV behavior differs in Asian and Western cohorts [60], the geographic region where patient specimens were collected is indicated for each study. The majority of studies (72.7%) collected patient specimens in Asian populations, while only 6.0% examined Western cohorts.

3.3. Recurrent Sites of HBV Integration

Next, we identified RTGs in the compiled HCC cohort using the criteria described in the Materials and Methods and then examined their associations with carcinogenesis from the published literature. Of the 18,596 integration sites examined, 2892 sites were found within 150 kb of gene-coding sequences, and 396 were RTGs. The 396 RTGs were found in 673 HCC patients, which was 51% of the integration-positive patients (n = 1310) or 39% of all HCC patients (n = 1733). To investigate the association of frequent RTGs with carcinogenesis, we analyzed the most frequently recurrent genes (n = 28), defined as found to contain HBV integrants in at least 10 HCC patients (Table 3). As expected, TERT and MLL4 were the two most recurrent genes. Interestingly, these 28 genes either have previously been associated with carcinogenesis (23/28, 82%), or have no known function (5/28, 18%). A full list of biological pathways associated with the 28 RTGs is detailed in Supplementary Table S3.
Next, the 396 RTGs were queried for significantly enriched gene ontology (GO) pathways using Enrichr [61]. The top enriched biological pathways were axon guidance (p = 0.00007) and positive regulation of Ras protein signal transduction (p = 0.0001), suggesting possible links with oncogenesis (Figure 1A). Heparin sulfate-glucosamine 3-sulfotransferase I (HS3ST1) activity was among the top enriched pathways from GO molecular functions (Figure 1B). Sulfotransferases have reported associations with carcinogenic activity, and HS3ST1 in particular has been implicated in inflammation [62]. Finally, a search of the Drug Signatures Database (DSigbDB) identified trichostatin, which selectively inhibits class I and II histone deacetylases (HDACs), as the drug/compound related to most RTGs (115/396; Figure 1C).
Table 3. Most frequently reported genes with HBV integrants in HCC tumors.
Table 3. Most frequently reported genes with HBV integrants in HCC tumors.
RTGFull Gene NameSubjects
(n)
Junctions
(n)
Cancer Associated [Ref]
TERT1Telomerase reverse transcriptase357561Multiple cancers [63]
MLL4 (KMT2B)Lysine methyltransferase 2B130220HCC [52,64], Spindle cell sarcoma [65], Gastric cancer [66]
PLEKHG4BPleckstrin Homology and RhoGEF Domain Containing G4B 38115Neuroblastoma [67]
LOC100288778WAS protein family homolog 8 pseudogene3479SCLC [68]
DDX11L1DEAD/H-Box Helicase 11 Like 13357Function unknown
CCNE1Cyclin E13154Multiple cancers [69]
CCNA2Cyclin A22645Multiple cancers [70]
SNTG1Syntrophin Gamma 12527Lung adenocarcinoma [71]
PGBD2 PiggyBac Transposable Element Derived 2 2251Function unknown
DUX4L4Double homeobox 4 like 4 pseudogene2035DUX4 Ewing’s sarcoma [72], ALL [73]
ROCK1P1Rho Associated Coiled-Coil Containing Protein Kinase 1 Pseudogene 12035Prostate cancer [74]
ANKRD26P1ankyrin repeat domain 26 pseudogene 11972Breast cancer [75]
PARD6GPar-6 Family Cell Polarity Regulator Gamma1942Breast, kidney, liver, lung, ovary, and pancreatic cancers [76]
FN1Fibronectin 11718Multiple cancers [77]
CWH43Cell Wall Biogenesis 43 C-Terminal Homolog1473CRC and TSHomas [78]
TPTETransmembrane Phosphatase with Tensin Homology1330HCC [79], prostate cancer [80]
FAM157AFamily with Sequence Similarity 157 Member A1422Function unknown
LOC728323/LINC01881Long Intergenic Non-Protein Coding RNA 18811322Oral cancer [81]
EMBP1Embigin pseudogene 11227Oropharyngeal carcinoma [82], multiple primary cancers [83]
OR4C6Olfactory Receptor Family 4 Subfamily C Member 61222Pancreatic cancer [84]
PRMT2Protein Arginine Methyltransferase 21215Glioblastoma [85]
ROCK1Rho Associated Coiled-Coil Containing Protein Kinase 11223HCC [86,87,88,89], CRC [90]
ANHXAnomalous Homeobox1116Function unknown
CTNND2Catenin Delta 21114HCC [91,92], prostate cancer [93], lung cancer [94]
DDX11L9DEAD/H-Box Helicase 11 Like 9 (Pseudogene)1116Function unknown
SENP5SUMO Specific Peptidase 51111HCC [95], breast cancer [96]
ZNF595Zinc Finger Protein 5951114Lung cancer [97], Gastric cancer [98]
CDRT7CMT1A Duplicated Region Transcript 71010Glioma [99]
1 The number of reported TERT junctions is slightly skewed, because one of the references [45] only reported the number of TERT promoter integrations. RTG, recurrently targeted gene; HCC, hepatocellular carcinoma; NSCLC, non-small cell lung cancer; SCLC, small cell lung cancer; ALL, acute lymphocytic leukemia; CRC, colorectal cancer; TSHoma, thyrotropin-secreting pituitary adenoma; CLL, chronic lymphocytic leukemia; RCC, renal cell carcinoma.

3.4. Integration Breakpoints in the HBV Genome

To investigate the distribution pattern of the integration breakpoints in the HBV genome, we analyzed HBV breakpoints in tumors (n = 4008) where available. We omitted studies that enriched for HBV DR1-2 sequences to assess the HBV breakpoint distribution in an unbiased manner. Consistent with previous reports [6,39,43], we observed that 39.2% of all breakpoints were within the HBV nt. 1300–1900 region in tumors. This region is at the 3′ end of the HBx gene and contains the initiation site of viral replication/transcription. Additionally, we observed a breakpoint hotspot in the HBV DR1-2 region, representing 29.6% of all HBV breakpoints (Figure 2).

3.5. Genomic Breakpoints of TERT, MLL4 and PLEKHG4B RTGs

HBV integration is believed to be non-sequence-specific; therefore, it was of interest to examine all RTG breakpoint coordinates for similarity to each other. To do so, we plotted the available human and HBV breakpoint coordinates of the three most frequent RTGs identified, TERT, MLL4, and PLEKHG4B (Figure 3).
For TERT, the most frequent RTG, 291 of 561 junctions from 205 HCC patients, had both human and HBV breakpoint coordinates available. As expected, most of these breakpoints were centered between DR2 and DR1 of the viral genome and were highly concentrated at the promoter region of the TERT gene (Figure 3A). Of note, 142 (48.7%) TERT integration junction sequences were unique, i.e., reported from 142 different subjects, whereas the rest (57 sequences) were each found in two or more HCC patients. Of the 477 available breakpoint coordinates in the TERT gene, 361 (75.6%) junctions were located upstream of exon 1 and, of these upstream breakpoints, 242 (50.7%) were located within the TERT promoter region (Chr5:1,295,162–1,296,162).
MLL4 was the second most frequently reported RTG, with 220 junctions identified from 130 HCC patients. Among them, 150 breakpoints from 83 HCC patients had both human and viral coordinates available and are plotted in Figure 3B. As with TERT, most of the breakpoints were clustered between DR2 and DR1 of the viral genome and concentrated within exon 3 of the MLL4 gene. Eleven different recurring breakpoints were found in two or more HCC patients, accounting for 37 of the 150 junctions examined.
The third most reported RTG was PLEKHG4B, with 116 integrations reported in this study. Interestingly, all the 116 breakpoints were centered within a 3 kb region which was 131 kb upstream from the PLEKHG4B coding region. A total of 47 breakpoints from eight HCC patients had both viral and human coordinates available, as shown in Figure 3C. All the breakpoints were found upstream of the transcription start site (Chr5:140,373). Unlike in TERT and MLL4, the viral breakpoints were centered in two HBV regions (nt. 1802–1814 and 2390; 15 and 14 breakpoints, respectively) but at various human coordinates. Further analysis of the human sequences (Chr5:10,000–13,000) in the upstream integration region revealed a 1877 bp segment containing simple repeat sequences and a 1057 bp segment containing satellite sequences. This finding is interesting because HBV has been suggested to have a higher propensity to integrate into repeat regions/retrotransposons, as recently shown by Chauhan et al. [100]. In addition, a motif, TAAACCCTAAC, was discovered, appearing four times in the Chr5:10,000–13,000 region and once in the HBV genome, each with p < 0.0001 by the Student’s t-test. However, querying the GenomeNet database for this motif produced no matches, and analysis of the region for known motifs produced no results. No recurrent breakpoints were identified among the integration sites in the PLEKHG4B gene. Notably, seven of the eight HCC patients with this unique junction pattern, where the same HBV breakpoint was joined to different host chromosome breakpoints, were reported from one study by Yang et al. [43].
TERT hotspot promoter mutations (−124, −146) are the most frequently reported mutations in HCC, found in about 50% of cases [101,102,103,104,105,106]. In HBV–HCC, up-regulation of TERT expression could also be caused by HBV integration at or near the TERT promoter region [28,29,30,32,38,48]. Therefore, we next compared the frequencies of TERT promoter mutation and HBV integration. For our in-house cohort (n = 22), shown in Figure 4A, promoter mutations were found in 6 of 22 samples, and integrations in the TERT gene were found in 5 of 22 samples in a mutually exclusive fashion. Together, TERT alterations were detected in 50% (11/22) of this cohort. To determine if this mutual exclusivity applied to a larger sample size, we examined TERT alterations identified by us and others [40,42,45] together, as summarized in Figure 4B. Of the 347 HBV–HCC patients, 174 (50%) were found to have detectable TERT alterations. Promoter mutations and HBV integrations were mutually exclusive and comprised 40% (70/174) and 60% (104/174) of all alterations, respectively.

4. Discussion

In this study, we compiled and studied a total of 18,596 HBV integration sites from 1733 HCC patients reported in 32 previous studies and our in-house study, to test our hypothesis that frequent HBV RTGs are associated with HCC carcinogenesis and thus may represent HCC driver genes that can reveal potential therapeutic targets and molecular profiling for precision medicine. Using three criteria for RTG selection, we identified 396 RTGs. Encouragingly, the most frequent RTGs (n = 28), reported in at least 10 HCC patients each, either have known involvement in carcinogenesis (23/28; 82%) or have no known function (5/28; 18%). By gene ontology analysis, RTGs were mapped to functions related to carcinogenesis. We describe a potential tool to identify or discover HCC drivers by the characterization of frequent RTGs. More studies are needed to demonstrate the association of carcinogenesis with the frequency of RTGs that have unknown functions.
Three criteria were applied to candidate genes to identify 396 HBV RTGs in this study: (1) location within 150 kb of the breakpoint, the reported distance within which host genes can be impacted by integration [27,28]; (2) the presence of HBV integrants in ≥2 HCC patients; and (3) identification in ≥2 independent laboratories to avoid the possibility of contamination within a laboratory. We are aware that the identification of RTGs across multiple studies is complex due to multi-faceted underlying variables such as integration detection methodologies and patient populations. For instance, some studies do not contain any of the 396 RTGs that we identified [54,55], while others report a high detection rate of a particular RTG, such as MLL4 [52] or cMYC [39]. We are also aware that different methodologies for identifying integrations may have different sensitivities that can result in different integration site profiles. Furthermore, how DNA samples were prepared varied across the studies. Despite these limitations, the detection of highly frequent RTGs as potential HBV–HCC drivers could be clinically useful for HCC patients. Notably, our approach can only identify driver genes involved in HBV–HCC (rather than HCC in general) and is limited to insertional mutagenesis. In addition, HBV integration could contribute to carcinogenesis by causing chromosomal instability, expression of the oncogenic HBV protein X, and inflammation induced by other HBV proteins expressed from the integrated DNA. Although the most frequent RTG, TERT, is also the most frequently mutated gene in HCC in general, other frequently mutated HCC genes, such as TP53 and CTNNB1, were not among the top 28 RTGs. These findings suggest that using RTG analysis as a tool to identify driver genes is limited to HBV–HCC.
By detailed analysis of integrations in the three most frequent RTGs (TERT, MLL4, and PLEKHG4B), we revealed three important features. First, as expected, most of the junction coordinates are different, supporting a non-sequence-specific mechanism of integration in the host genome. Among the junction sequences identified in multiple HCC patients, such as the 57 junctions described in this study, frequent integration in the promoter region of the TERT gene highlights the potential importance of the site in hepatocarcinogenesis. Secondly, an interesting pattern of integration into repetitive sequences was observed in the PLEKH4G4B junctions shown in Figure 3C. Furthermore, a highly repetitive sequence, satellite sequences, and the TAAACCCTAAC motif were identified in these regions, consistent with previous observations that integration occurred frequently in repetitive sequences [100]. However, because these unique integration sequence patterns were reported in only one study and validation in original tissue DNA was not reported, the possibility of an artifact has not been excluded. Lastly, the mutually exclusive detection of TERT promoter mutations and TERT integrations was shown in our small cohort of 22 HCC patients as well as in a larger compiled cohort of 347 HCC patients [40,42,45]. TERT promoter mutations account for 50% of all alterations in this gene, suggesting the importance of identifying TERT integration. This observation also further emphasizes the need for analysis of frequent RTGs to better characterize HCC.
In summary, we identified 396 RTGs from an analysis of 18,596 HBV integration sites. Twenty-eight of the RTGs were identified in 10 or more HCC patients. The RTGs were found to be significantly enriched in several GO pathways. The 28 most frequent RTGs have previously been implicated in carcinogenesis (23/28) or have no known function (5/28). Taken together, our findings demonstrate the potential of RTG identification as a tool for HBV–HCC driver discovery and tumor characterization, to be used in precision medicine and drug development. More studies are needed to further refine the criteria for RTG identification for its applications.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cells10061294/s1, Table S1: Characterization of HBV-JSs identified in in-house HBV–HCC tissue; Table S2: Clinical characteristics of in-house HBV–HCC patient cohort (n = 22); Table S3. Pathways identified by Enrichr GO biological processes of the top 28 RTGs.

Author Contributions

Conceptualization, S.Y.L. and Y.-H.S.; formal analysis, S.Y.L., A.Z., J.L. and J.W.; patient samples/resources, T.-T.C. and Y.-J.L.; data curation, S.Y.L.; writing—original draft preparation, S.Y.L., A.Z., J.W. and Y.-H.S.; writing—review and editing, all authors; supervision, Y.-H.S.; funding acquisition, W.S. and Y.-H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Institutes of Health R43 CA165312 (W.S. and Y.-H.S.), R44 CA165312 (S.Y.L. and W.S.), and R43 CA192507 (W.S.).

Institutional Review Board Statement

Ethical review and approval were waived for this study, because this was a retrospective meta-analysis study and the samples were from the archived and de-identified samples from other studies as described previously [23,24]; thus, the study does not need to be considered as a human subject study.

Informed Consent Statement

Not applicable.

Data Availability Statement

Detailed data are available upon request.

Acknowledgments

The authors thank Dmitry Goryunov for review of the manuscript.

Conflicts of Interest

S.Y.L. and W.S. are employees and shareholders of JBS Science Inc. All other authors declare no conflict of interest.

References

  1. Howlader, N.N.A.; Krapcho, M.; Miller, D.; Bishop, K.; Altekruse, S.F.; Kosary, C.L.; Yu, M.; Ruhl, J.; Tatalovich, Z.; Mariotto, A. (Eds.) SEER Cancer Statistics Review, 1975–2013; Based on November 2015 SEER Data Submission, Posted to the SEER Web Site; National Cancer Institute: Bethesda, MD, USA, April 2016. Available online: http://seer.cancer.gov/csr/1975_2013/ (accessed on 5 December 2020).
  2. Shinozaki, M.; Hoon, D.S.; Giuliano, A.E.; Hansen, N.M.; Wang, H.-J.; Turner, R.; Taback, B. Distinct Hypermethylation Profile of Primary Breast Cancer Is Associated with Sentinel Lymph Node Metastasis. Clin. Cancer Res. 2005, 11, 2156–2162. [Google Scholar] [CrossRef] [Green Version]
  3. Ferlay, J.S.I.; Ervik, M.; Dikshit, R.; Eser, S.; Mathers, C.; Rebelo, M.; Parkin, D.M.; Forman, D.; Bray, F. GLOBOCAN 2012 v1.0, Cancer Incidence and Mortality Worldwide: IARC CancerBase No. 11; International Agency for Research on Cancer: Lyon, France, 2013; Available online: http://globocan.iarc.fr (accessed on 14 March 2017).
  4. El-Serag, H.B. Epidemiology of Viral Hepatitis and Hepatocellular Carcinoma. Gastroenterology 2012, 142, 1264–1273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Anugwom, C.M.; Allaire, M.; Akbar, S.M.F.; Sultan, A.; Bollipo, S.; Mattos, A.Z.; Debes, J.D. Hepatitis B-related hepatocellular carcinoma: Surveillance strategy directed by immune-epidemiology. Hepatoma Res. 2021, 7. [Google Scholar] [CrossRef]
  6. Zhao, L.-H.; Liu, X.; Yan, H.-X.; Li, W.-Y.; Zeng, X.; Yang, Y.; Zhao, J.; Liu, S.-P.; Zhuang, X.-H.; Lin, C.; et al. Genomic and oncogenic preference of HBV integration in hepatocellular carcinoma. Nat. Commun. 2016, 7, 12992. [Google Scholar] [CrossRef]
  7. Lau, C.-C.; Sun, T.; Ching, A.K.; He, M.; Li, J.-W.; Wong, A.M.; Co, N.N.; Chan, A.W.; Li, P.-S.; Lung, R.W.; et al. Viral-Human Chimeric Transcript Predisposes Risk to Liver Cancer Development and Progression. Cancer Cell 2014, 25, 335–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Yoo, S.; Wang, W.; Wang, Q.; Fiel, M.I.; Lee, E.; Hiotis, S.P.; Zhu, J. A pilot systematic genomic comparison of recurrence risks of hepatitis B virus-associated hepatocellular carcinoma with low- and high-degree liver fibrosis. BMC Med. 2017, 15, 214. [Google Scholar] [CrossRef] [Green Version]
  9. Chauhan, R.; Churchill, N.D.; Mulrooney-Cousins, P.M.; I Michalak, T. Initial sites of hepadnavirus integration into host genome in human hepatocytes and in the woodchuck model of hepatitis B-associated hepatocellular carcinoma. Oncogenesis 2017, 6, e317. [Google Scholar] [CrossRef] [PubMed]
  10. Bill, C.A.; Summers, J. Genomic DNA double-strand breaks are targets for hepadnaviral DNA integration. Proc. Natl. Acad. Sci. USA 2004, 101, 11135–11140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Budzinska, M.A.; Shackel, N.A.; Urban, S.; Tu, T. Sequence analysis of integrated hepatitis B virus DNA during HBeAg-seroconversion. Emerg. Microbes Infect. 2018, 7, 1–12. [Google Scholar] [CrossRef]
  12. Lindh, M.; Rydell, G.E.; Larsson, S.B. Impact of integrated viral DNA on the goal to clear hepatitis B surface antigen with different therapeutic strategies. Curr. Opin. Virol. 2018, 30, 24–31. [Google Scholar] [CrossRef]
  13. Scotto, J.; Hadchouel, M.; Hery, C.; Alvarez, F.; Yvart, J.; Tiollais, P.; Bernard, O.; Brechot, C. Hepatitis B virus DNA in children’s liver diseases: Detection by blot hybridisation in liver and serum. Gut 1983, 24, 618–624. [Google Scholar] [CrossRef]
  14. Huang, H.-P.; Tsuei, D.-J.; Wang, K.-J.; Chen, Y.-L.; Ni, Y.-H.; Jeng, Y.-M.; Chen, H.-L.; Hsu, H.-Y.; Chang, M.-H. Differential integration rates of hepatitis B virus DNA in the liver of children with chronic hepatitis B virus infection and hepatocellular carcinoma. J. Gastroenterol. Hepatol. 2005, 20, 1206–1214. [Google Scholar] [CrossRef]
  15. Shafritz, D.A.; Shouval, D.; Sherman, H.I.; Hadziyannis, S.J.; Kew, M.C. Integration of Hepatitis B Virus DNA into the Genome of Liver Cells in Chronic Liver Disease and Hepatocellular Carcinoma. N. Engl. J. Med. 1981, 305, 1067–1073. [Google Scholar] [CrossRef]
  16. Koshy, R.; Maupas, P.; Muller, R.; Hofschneider, P.H. Detection of Hepatitis B Virus-specific DNA in the Genomes of Human Hepatocellular Carcinoma and Liver Cirrhosis Tissues. J. Gen. Virol. 1981, 57, 95–102. [Google Scholar] [CrossRef]
  17. Takada, S.; Gotoh, Y.; Hayashi, S.; Yoshida, M.; Koike, K. Structural rearrangement of integrated hepatitis B virus DNA as well as cellular flanking DNA is present in chronically infected hepatic tissues. J. Virol. 1990, 64, 822–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Tu, T.; Budzinska, M.A.; Vondran, F.W.R.; Shackel, N.A.; Urban, S. Hepatitis B Virus DNA Integration Occurs Early in the Viral Life Cycle in anIn VitroInfection Model via Sodium Taurocholate Cotransporting Polypeptide-Dependent Uptake of Enveloped Virus Particles. J. Virol. 2018, 92, e02007–e02017. [Google Scholar] [CrossRef] [Green Version]
  19. Hai, H. Role of hepatitis B virus DNA integration in human hepatocarcinogenesis. World J. Gastroenterol. 2014, 20, 6236–6243. [Google Scholar] [CrossRef] [PubMed]
  20. Dong, H.; Zhang, L.; Qian, Z.; Zhu, X.; Zhu, G.; Chen, Y.; Xie, X.; Ye, Q.; Zang, J.; Ren, Z.; et al. Identification of HBV-MLL4 Integration and Its Molecular Basis in Chinese Hepatocellular Carcinoma. PLoS ONE 2015, 10, e0123175. [Google Scholar] [CrossRef] [Green Version]
  21. Jiang, Z.; Jhunjhunwala, S.; Liu, J.; Haverty, P.M.; Kennemer, M.I.; Guan, Y.; Lee, W.; Carnevali, P.; Stinson, J.; Johnson, S.; et al. The effects of hepatitis B virus integration into the genomes of hepatocellular carcinoma patients. Genome Res. 2012, 22, 593–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Jhunjhunwala, S.; Jiang, Z.; Stawiski, E.W.; Gnad, F.; Liu, J.; Mayba, O.; Du, P.; Diao, J.; Johnson, S.; Wong, K.-F. Diverse modes of genomic alterations in hepatocellular carcinoma. Genome Biol. 2014, 15, 436. [Google Scholar] [PubMed] [Green Version]
  23. Lin, S.Y.; Dhillon, V.; Jain, S.; Chang, T.-T.; Hu, C.-T.; Lin, Y.-J.; Chen, S.-H.; Chang, K.-C.; Song, W.; Yu, L.; et al. A Locked Nucleic Acid Clamp-Mediated PCR Assay for Detection of a p53 Codon 249 Hotspot Mutation in Urine. J. Mol. Diagn. 2011, 13, 474–484. [Google Scholar] [CrossRef]
  24. Jain, S.; Xie, L.; Boldbaatar, B.; Lin, S.Y.; Hamilton, J.P.; Meltzer, S.J.; Chen, S.-H.; Hu, C.-T.; Block, T.M.; Song, W.; et al. Differential methylation of the promoter and first exon of the RASSF1A gene in hepatocarcinogenesis. Hepatol. Res. 2014, 45, 1110–1123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Shieh, F.-S.; Jongeneel, P.; Steffen, J.D.; Lin, S.; Jain, S.; Song, W.; Su, Y.-H. ChimericSeq: An open-source, user-friendly interface for analyzing NGS data to identify and characterize viral-host chimeric sequences. PLoS ONE 2017, 12, e0182843. [Google Scholar] [CrossRef] [Green Version]
  26. Madeira, F.; Park, Y.M.; Lee, J.; Buso, N.; Gur, T.; Madhusoodanan, N.; Basutkar, P.; Tivey, A.R.N.; Potter, S.C.; Finn, R.D.; et al. The EMBL-EBI search and sequence analysis tools APIs. Nucleic Acids Res. 2019, 47, W636–W641. [Google Scholar] [CrossRef] [Green Version]
  27. Shamay, M.; Agami, R.; Shaul, Y. HBV integrants of hepatocellular carcinoma cell lines contain an active enhancer. Oncogene 2001, 20, 6811–6819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Horikawa, I.; Barrett, J.C. cis-Activation of the Human Telomerase Gene (hTERT) by the Hepatitis B Virus Genome. J. Natl. Cancer Inst. 2001, 93, 1171–1173. [Google Scholar] [CrossRef] [PubMed]
  29. Ding, D.; Lou, X.; Hua, D.; Yu, W.; Li, L.; Wang, J.; Gao, F.; Zhao, N.; Ren, G.; Li, L.; et al. Recurrent Targeted Genes of Hepatitis B Virus in the Liver Cancer Genomes Identified by a Next-Generation Sequencing–Based Approach. PLoS Genet. 2012, 8, e1003065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Fujimoto, A.; Totoki, Y.; Abe, T.; Boroevich, K.A.; Hosoda, F.; Nguyen, H.H.; Aoki, M.; Hosono, N.; Kubo, M.; Miya, F.; et al. Whole-genome sequencing of liver cancers identifies etiological influences on mutation patterns and recurrent mutations in chromatin regulators. Nat. Genet. 2012, 44, 760–764. [Google Scholar] [CrossRef]
  31. Fujimoto, A.; Furuta, M.; Totoki, Y.; Tsunoda, T.; Kato, M.; Shiraishi, Y.; Tanaka, H.; Taniguchi, H.; Kawakami, Y.; Ueno, M.; et al. Whole-genome mutational landscape and characterization of noncoding and structural mutations in liver cancer. Nat. Genet. 2016, 48, 500–509. [Google Scholar] [CrossRef]
  32. Sung, W.-K.; Zheng, H.; Li, S.; Chen, R.; Liu, X.; Li, Y.; Lee, N.P.; Lee, W.H.; Ariyaratne, P.N.; Tennakoon, C.; et al. Genome-wide survey of recurrent HBV integration in hepatocellular carcinoma. Nat. Genet. 2012, 44, 765–769. [Google Scholar] [CrossRef]
  33. Li, W.; Zeng, X.; Lee, N.P.; Liu, X.; Chen, S.; Guo, B.; Yi, S.; Zhuang, X.; Chen, F.; Wang, G.; et al. HIVID: An efficient method to detect HBV integration using low coverage sequencing. Genomics 2013, 102, 338–344. [Google Scholar] [CrossRef] [PubMed]
  34. Miao, R.; Luo, H.; Zhou, H.; Li, G.; Bu, D.; Yang, X.; Zhao, X.; Zhang, H.; Liu, S.; Zhong, Y.; et al. Identification of prognostic biomarkers in hepatitis B virus-related hepatocellular carcinoma and stratification by integrative multi-omics analysis. J. Hepatol. 2014, 61, 840–849. [Google Scholar] [CrossRef] [PubMed]
  35. Hama, N.; Totoki, Y.; Miura, F.; Tatsuno, K.; Saito-Adachi, M.; Nakamura, H.; Arai, Y.; Hosoda, F.; Urushidate, T.; Ohashi, S.; et al. Epigenetic landscape influences the liver cancer genome architecture. Nat. Commun. 2018, 9, 1643. [Google Scholar] [CrossRef]
  36. Duan, M.; Hao, J.; Cui, S.; Worthley, D.L.; Zhang, S.; Wang, Z.; Shi, J.; Liu, L.; Wang, X.; Ke, A.; et al. Diverse modes of clonal evolution in HBV-related hepatocellular carcinoma revealed by single-cell genome sequencing. Cell Res. 2018, 28, 359–373. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, W.; Zhang, K.; Dong, P.; Fanning, G.; Tao, C.; Zhang, H.; Guo, S.; Wang, Z.; Hong, Y.; Yang, X.; et al. Noninvasive chimeric DNA profiling identifies tumor-originated HBV integrants contributing to viral antigen expression in liver cancer. Hepatol. Int. 2020, 14, 326–337. [Google Scholar] [CrossRef]
  38. Toh, S.T.; Jin, Y.; Liu, L.; Wang, J.; Babrzadeh, F.; Gharizadeh, B.; Ronaghi, M.; Toh, H.C.; Chow, P.K.-H.; Chung, A.Y. Deep sequencing of the hepatitis B virus in hepatocellular carcinoma patients reveals enriched integration events, structural alterations and sequence variations. Carcinogenesis 2013, 34, 787–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Yan, H.; Yang, Y.; Zhang, L.; Tang, G.; Wang, Y.; Xue, G.; Zhou, W.; Sun, S. Characterization of the genotype and integration patterns of hepatitis B virus in early-and late-onset hepatocellular carcinoma. Hepatology 2015, 61, 1821–1831. [Google Scholar] [CrossRef]
  40. Kawai-Kitahata, F.; Asahina, Y.; Tanaka, S.; Kakinuma, S.; Murakawa, M.; Nitta, S.; Watanabe, T.; Otani, S.; Taniguchi, M.; Goto, F.; et al. Comprehensive analyses of mutations and hepatitis B virus integration in hepatocellular carcinoma with clinicopathological features. J. Gastroenterol. 2015, 51, 473–486. [Google Scholar] [CrossRef]
  41. Furuta, M.; Tanaka, H.; Shiraishi, Y.; Uchida, T.; Imamura, M.; Fujimoto, A.; Fujita, M.; Sasaki-Oku, A.; Maejima, K.; Nakano, K.; et al. Characterization of HBV integration patterns and timing in liver cancer and HBV-infected livers. Oncotarget 2018, 9, 25075–25088. [Google Scholar] [CrossRef] [Green Version]
  42. Li, C.L.; Li, C.Y.; Lin, Y.Y.; Ho, M.C.; Chen, D.S.; Yeh, S.H.; Chen, P.J. Androgen Receptor Enhances Hepatic TERT Transcription after Hepatitis B Virus Integration or Point Mutation in Promoter Region. Hepatology 2019, 69, 498–512. [Google Scholar] [CrossRef]
  43. Yang, L.; Ye, S.; Zhao, X.; Ji, L.; Zhang, Y.; Zhou, P.; Sun, J.; Guan, Y.; Han, Y.; Ni, C.; et al. Molecular Characterization of HBV DNA Integration in Patients with Hepatitis and Hepatocellular Carcinoma. J. Cancer 2018, 9, 3225–3235. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, X.-P.; Long, X.; Jia, W.-L.; Wu, H.-J.; Zhao, J.; Liang, H.-F.; Laurence, A.; Zhu, J.; Dong, D.; Chen, Y. Viral integration drives multifocal HCC during the occult HBV infection. J. Exp. Clin. Cancer Res. 2019, 38, 1–11. [Google Scholar]
  45. Sze, K.M.; Ho, D.W.; Chiu, Y.; Tsui, Y.; Chan, L.; Lee, J.M.; Chok, K.S.; Chan, A.C.; Tang, C.; Tang, V.W.; et al. Hepatitis B Virus–Telomerase Reverse Transcriptase Promoter Integration Harnesses Host ELF4, Resulting in Telomerase Reverse Transcriptase Gene Transcription in Hepatocellular Carcinoma. Hepatology 2021, 73, 23–40. [Google Scholar] [CrossRef] [Green Version]
  46. Tatsuno, K.; Midorikawa, Y.; Takayama, T.; Yamamoto, S.; Nagae, G.; Moriyama, M.; Nakagawa, H.; Koike, K.; Moriya, K.; Aburatani, H. Impact of AAV2 and Hepatitis B Virus Integration into Genome on Development of Hepatocellular Carcinoma in Patients with Prior Hepatitis B Virus Infection. Clin. Cancer Res. 2019, 25, 6217–6227. [Google Scholar] [CrossRef]
  47. Li, C.; Ho, M.; Lin, Y.; Tzeng, S.; Chen, Y.; Pai, H.; Wang, Y.; Chen, C.; Lee, Y.; Chen, D.; et al. Cell-Free Virus-Host Chimera DNA From Hepatitis B Virus Integration Sites as a Circulating Biomarker of Hepatocellular Cancer. Hepatology 2020, 72, 2063–2076. [Google Scholar] [CrossRef]
  48. Ferber, M.J.; Montoya, D.P.; Yu, C.; Aderca, I.; McGee, A.; Thorland, E.C.; Nagorney, D.M.; Gostout, B.S.; Burgart, L.J.; Boix, L.; et al. Integrations of the hepatitis B virus (HBV) and human papillomavirus (HPV) into the human telomerase reverse transcriptase (hTERT) gene in liver and cervical cancers. Oncogene 2003, 22, 3813–3820. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, Y.; Lau, S.H.; Sham, J.S.-T.; Wu, M.-C.; Wang, T.; Guan, X.-Y. Characterization of HBV integrants in 14 hepatocellular carcinomas: Association of truncated X gene and hepatocellular carcinogenesis. Oncogene 2004, 23, 142–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Tamori, A.; Yamanishi, Y.; Kawashima, S.; Kanehisa, M.; Enomoto, M.; Tanaka, H.; Kubo, S.; Shiomi, S.; Nishiguchi, S. Alteration of Gene Expression in Human Hepatocellular Carcinoma with Integrated Hepatitis B Virus DNA. Clin. Cancer Res. 2005, 11, 5821–5826. [Google Scholar] [CrossRef] [Green Version]
  51. Murakami, Y.; Saigo, K.; Takashima, H.; Minami, M.; Okanoue, T.; Bréchot, C.; Paterlini-Bréchot, P. Large scaled analysis of hepatitis B virus (HBV) DNA integration in HBV related hepatocellular carcinomas. Gut 2005, 54, 1162–1168. [Google Scholar] [CrossRef] [PubMed]
  52. Saigo, K.; Yoshida, K.; Ikeda, R.; Sakamoto, Y.; Murakami, Y.; Urashima, T.; Asano, T.; Kenmochi, T.; Inoue, I. Integration of hepatitis B virus DNA into the myeloid/lymphoid or mixed-lineage leukemia (MLL4) gene and rearrangements of MLL4 in human hepatocellular carcinoma. Hum. Mutat. 2008, 29, 703–708. [Google Scholar] [CrossRef]
  53. Jiang, S.; Yang, Z.; Li, W.; Li, X.; Wang, Y.; Zhang, J.; Xu, C.; Chen, P.-J.; Hou, J.; McCrae, M.A.; et al. Re-evaluation of the Carcinogenic Significance of Hepatitis B Virus Integration in Hepatocarcinogenesis. PLoS ONE 2012, 7, e40363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Saitta, C.; Tripodi, G.; Barbera, A.; Bertuccio, A.; Smedile, A.; Ciancio, A.; Raffa, G.; Sangiovanni, A.; Navarra, G.; Raimondo, G.; et al. Hepatitis B virus (HBV) DNA integration in patients with occult HBV infection and hepatocellular carcinoma. Liver Int. 2015, 35, 2311–2317. [Google Scholar] [CrossRef] [Green Version]
  55. Fang, X.; Wu, H.-H.; Ren, J.-J.; Liu, H.-Z.; Li, K.-Z.; Li, J.-L.; Tang, Y.-P.; Xiao, C.-C.; Huang, T.-R.; Deng, W. Associations between serum HBX quasispecies and their integration in hepatocellular carcinoma. Int. J. Clin. Exp. Pathol. 2017, 10, 11857–11866. [Google Scholar]
  56. Ruan, P.; Dai, X.; Sun, J.; He, C.; Huang, C.; Zhou, R.; Cao, Z.; Ye, L. Different types of viral-host junction found in HBV integration breakpoints in HBV-infected patients. Mol. Med. Rep. 2018, 19, 1410–1416. [Google Scholar] [CrossRef] [PubMed]
  57. He, Z.; Zhu, J.; Mo, J.; Zhao, H.; Chen, Q. HBV DNA integrates into upregulated ZBTB20 in patients with hepatocellular carcinoma. Mol. Med. Rep. 2020, 22, 380–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Yang, J.D.; Kim, W.R.; Coelho, R.; Mettler, T.A.; Benson, J.T.; Sanderson, S.O.; Therneau, T.M.; Kim, B.; Roberts, L.R. Cirrhosis Is Present in Most Patients with Hepatitis B and Hepatocellular Carcinoma. Clin. Gastroenterol. Hepatol. 2011, 9, 64–70. [Google Scholar] [CrossRef] [Green Version]
  59. El–Serag, H.B.; Rudolph, K.L. Hepatocellular Carcinoma: Epidemiology and Molecular Carcinogenesis. Gastroenterology 2007, 132, 2557–2576. [Google Scholar] [CrossRef]
  60. Mueller-Breckenridge, A.J.; Garcia-Alcalde, F.; Wildum, S.; Smits, S.L.; De Man, R.A.; Van Campenhout, M.J.H.; Brouwer, W.P.; Niu, J.; Young, J.A.T.; Najera, I.; et al. Machine-learning based patient classification using Hepatitis B virus full-length genome quasispecies from Asian and European cohorts. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef] [Green Version]
  61. Kuleshov, M.V.; Jones, M.R.; Rouillard, A.D.; Fernandez, N.F.; Duan, Q.; Wang, Z.; Koplev, S.; Jenkins, S.L.; Jagodnik, K.M.; Lachmann, A.; et al. Enrichr: A comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res. 2016, 44, W90–W97. [Google Scholar] [CrossRef] [Green Version]
  62. Smits, N.C.; Kobayashi, T.; Srivastava, P.K.; Skopelja, S.; Ivy, J.A.; Elwood, D.J.; Stan, R.V.; Tsongalis, G.J.; Sellke, F.W.; Gross, P.L. HS3ST1 genotype regulates antithrombin’s inflammomodulatory tone and associates with atherosclerosis. Matrix Biol. 2017, 63, 69–90. [Google Scholar] [CrossRef]
  63. Heidenreich, B.; Rachakonda, P.S.; Hemminki, K.; Kumar, R. TERT promoter mutations in cancer development. Curr. Opin. Genet. Dev. 2014, 24, 30–37. [Google Scholar] [CrossRef] [PubMed]
  64. Tamori, A.; Nishiguchi, S.; Shiomi, S.; Hayashi, T.; Kobayashi, S.; Habu, D.; Takeda, T.; Seki, S.; Hirohashi, K.; Tanaka, H.; et al. Hepatitis B Virus DNA Integration in Hepatocellular Carcinoma After Interferon-Induced Disappearance of Hepatitis C Virus. Am. J. Gastroenterol. 2005, 100, 1748–1753. [Google Scholar] [CrossRef]
  65. O’Meara, E.; Stack, D.; Phelan, S.; McDonagh, N.; Kelly, L.; Sciot, R.; Debiec-Rychter, M.; Morris, T.; Cochrane, D.; Sorensen, P.; et al. Identification of anMLL4-GPS2fusion as an oncogenic driver of undifferentiated spindle cell sarcoma in a child. Genes Chromosom. Cancer 2014, 53, 991–998. [Google Scholar] [CrossRef]
  66. Pan, X.; Ji, X.; Zhang, R.; Zhou, Z.; Zhong, Y.; Peng, W.; Sun, N.; Xu, X.; Xia, L.; Li, P.; et al. Landscape of somatic mutations in gastric cancer assessed using next-generation sequencing analysis. Oncol. Lett. 2018, 16, 4863–4870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Chicard, M.; Boyault, S.; Daage, L.C.; Richer, W.; Gentien, D.; Pierron, G.; Lapouble, E.; Bellini, A.; Clement, N.; Iacono, I.; et al. Genomic Copy Number Profiling Using Circulating Free Tumor DNA Highlights Heterogeneity in Neuroblastoma. Clin. Cancer Res. 2016, 22, 5564–5573. [Google Scholar] [CrossRef] [Green Version]
  68. Li, J.; Wang, J.; Chen, Y.; Yang, L.; Chen, S. A prognostic 4-gene expression signature for squamous cell lung carcinoma. J. Cell. Physiol. 2017, 232, 3702–3713. [Google Scholar] [CrossRef] [PubMed]
  69. Donnellan, R.; Chetty, R. Cyclin E in human cancers. FASEB J. 1999, 13, 773–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Otto, T.; Sicinski, T.O.P. Cell cycle proteins as promising targets in cancer therapy. Nat. Rev. Cancer 2017, 17, 93–115. [Google Scholar] [CrossRef] [Green Version]
  71. Meng, F.; Zhang, L.; Ren, Y.; Ma, Q. The genomic alterations of lung adenocarcinoma and lung squamous cell carcinoma can explain the differences of their overall survival rates. J. Cell. Physiol. 2019, 234, 10918–10925. [Google Scholar] [CrossRef]
  72. Yoshimoto, T.; Tanaka, M.; Homme, M.; Yamazaki, Y.; Takazawa, Y.; Antonescu, C.R.; Nakamura, T. CIC-DUX4 Induces Small Round Cell Sarcomas Distinct from Ewing Sarcoma. Cancer Res. 2017, 77, 2927–2937. [Google Scholar] [CrossRef] [Green Version]
  73. Yasuda, T.; Tsuzuki, S.; Kawazu, M.; Hayakawa, F.; Kojima, S.; Ueno, T.; Imoto, N.; Kohsaka, S.; Kunita, A.; Doi, K.; et al. Recurrent DUX4 fusions in B cell acute lymphoblastic leukemia of adolescents and young adults. Nat. Genet. 2016, 48, 569–574. [Google Scholar] [CrossRef] [PubMed]
  74. Luo, Y.; Jiang, Q.-W.; Wu, J.-Y.; Qiu, J.-G.; Zhang, W.-J.; Mei, X.-L.; Shi, Z.; Di, J.-M. Regulation of migration and invasion by Toll-like receptor-9 signaling network in prostate cancer. Oncotarget 2015, 6, 22564–22574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Krepischi, A.C.V.; Achatz, M.I.W.; Santos, E.M.M.; Costa, S.S.; Lisboa, B.C.G.; Brentani, H.; Santos, T.M.; Goncalves, A.; Nobrega, A.F.; Pearson, P.L.; et al. Germline DNA copy number variation in familial and early-onset breast cancer. Breast Cancer Res. 2012, 14, R24. [Google Scholar] [CrossRef]
  76. Marqués, E.; Englund, J.I.; A Tervonen, T.; Virkunen, E.; Laakso, M.; Myllynen, M.; Mäkelä, A.; Ahvenainen, M.; Lepikhova, T.; Monni, O.; et al. Par6G suppresses cell proliferation and is targeted by loss-of-function mutations in multiple cancers. Oncogene 2015, 35, 1386–1398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Hynes, R.O. Fibronectins. Sci. Am. 1986, 254, 42–51. [Google Scholar] [CrossRef]
  78. Chu, C.-M.; Yao, C.-T.; Chang, Y.-T.; Chou, H.-L.; Chou, Y.-C.; Chen, K.-H.; Terng, H.-J.; Huang, C.-S.; Lee, C.-C.; Su, S.-L.; et al. Gene Expression Profiling of Colorectal Tumors and Normal Mucosa by Microarrays Meta-Analysis Using Prediction Analysis of Microarray, Artificial Neural Network, Classification, and Regression Trees. Dis. Markers 2014, 2014, 1–11. [Google Scholar] [CrossRef] [Green Version]
  79. Dong, X.-Y.; Su, Y.-R.; Qian, X.-P.; Yang, X.-A.; Pang, X.-W.; Wu, H.-Y.; Chen, W.-F. Identification of two novel CT antigens and their capacity to elicit antibody response in hepatocellular carcinoma patients. Br. J. Cancer 2003, 89, 291–297. [Google Scholar] [CrossRef] [Green Version]
  80. Singh, A.P.; Bafna, S.; Chaudhary, K.; Venkatraman, G.; Smith, L.; Eudy, J.D.; Johansson, S.L.; Lin, M.-F.; Batra, S.K. Genome-wide expression profiling reveals transcriptomic variation and perturbed gene networks in androgen-dependent and androgen-independent prostate cancer cells. Cancer Lett. 2008, 259, 28–38. [Google Scholar] [CrossRef] [Green Version]
  81. Ambatipudi, S.; Gerstung, M.; Gowda, R.; Pai, P.; Borges, A.M.; Schäffer, A.A.; Beerenwinkel, N.; Mahimkar, M.B. Genomic Profiling of Advanced-Stage Oral Cancers Reveals Chromosome 11q Alterations as Markers of Poor Clinical Outcome. PLoS ONE 2011, 6, e17250. [Google Scholar] [CrossRef] [Green Version]
  82. Ren, S.; Gaykalova, D.; Wang, J.; Guo, T.; Danilova, L.; Favorov, A.; Fertig, E.; Bishop, J.; Khan, Z.; Flam, E.; et al. Discovery and development of differentially methylated regions in human papillomavirus-related oropharyngeal squamous cell carcinoma. Int. J. Cancer 2018, 143, 2425–2436. [Google Scholar] [CrossRef] [Green Version]
  83. Park, S.L.; Caberto, C.P.; Lin, Y.; Goodloe, R.J.; Dumitrescu, L.; Love, S.-A.; Matise, T.C.; Hindorff, L.A.; Fowke, J.H.; Schumacher, F.R.; et al. Association of Cancer Susceptibility Variants with Risk of Multiple Primary Cancers: The Population Architecture using Genomics and Epidemiology Study. Cancer Epidemiol. Biomark. Prev. 2014, 23, 2568–2578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Weber, L.; Maßberg, D.; Becker, C.; Altmüller, J.; Ubrig, B.; Bonatz, G.; Wölk, G.; Philippou, S.; Tannapfel, A.; Hatt, H.; et al. Olfactory Receptors as Biomarkers in Human Breast Carcinoma Tissues. Front. Oncol. 2018, 8, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Dong, F.; Li, Q.; Yang, C.; Huo, D.; Wang, X.; Ai, C.; Kong, Y.; Sun, X.; Wang, W.; Zhou, Y.; et al. PRMT2 links histone H3R8 asymmetric dimethylation to oncogenic activation and tumorigenesis of glioblastoma. Nat. Commun. 2018, 9, 4552. [Google Scholar] [CrossRef] [Green Version]
  86. Tang, J.; Liu, C.; Xu, B.; Wang, D.; Ma, Z.; Chang, X. ARHGEF10L contributes to liver tumorigenesis through RhoA-ROCK1 signaling and the epithelial-mesenchymal transition. Exp. Cell Res. 2019, 374, 46–68. [Google Scholar] [CrossRef] [PubMed]
  87. Liu, W.; Zhang, Q.; Tang, Q.; Hu, C.; Huang, J.; Liu, Y.; Lu, Y.; Wang, Q.; Li, G.; Zhang, R. Lycorine inhibits cell proliferation and migration by inhibiting ROCK1/cofilin-induced actin dynamics in HepG2 hepatoblastoma cells. Oncol. Rep. 2018, 40, 2298–2306. [Google Scholar] [CrossRef]
  88. Ding, W.; Tan, H.; Zhao, C.; Li, X.; Li, Z.; Jiang, C.; Zhang, Y.; Wang, L. MiR-145 suppresses cell proliferation and motility by inhibiting ROCK1 in hepatocellular carcinoma. Tumor Biol. 2016, 37, 6255–6260. [Google Scholar] [CrossRef] [PubMed]
  89. Deng, Q.; Xie, L.; Li, H. MiR-506 suppresses cell proliferation and tumor growth by targeting Rho-associated protein kinase 1 in hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2015, 467, 921–927. [Google Scholar] [CrossRef]
  90. Song, G.-L.; Jin, C.-C.; Zhao, W.; Tang, Y.; Wang, Y.-L.; Li, M.; Xiao, M.; Li, X.; Li, Q.-S.; Lin, X.; et al. Regulation of the RhoA/ROCK/AKT/β-catenin pathway by arginine-specific ADP-ribosytransferases 1 promotes migration and epithelial-mesenchymal transition in colon carcinoma. Int. J. Oncol. 2016, 49, 646–656. [Google Scholar] [CrossRef] [Green Version]
  91. Chen, R.; Dong, Y.; Xie, X.; Chen, J.; Gao, D.; Liu, Y.; Ren, Z.; Cui, J. Screening candidate metastasis-associated genes in three-dimensional HCC spheroids with different metastasis potential. Int. J. Clin. Exp. Pathol. 2014, 7, 2527–2535. [Google Scholar]
  92. Huang, F.; Chen, J.; Lan, R.; Wang, Z.; Chen, R.; Lin, J.; Fu, L. Hypoxia induced δ-Catenin to enhance mice hepatocellular carcinoma progression via Wnt signaling. Exp. Cell Res. 2019, 374, 94–103. [Google Scholar] [CrossRef]
  93. Zhang, P.; Schaefer-Klein, J.; Cheville, J.C.; Vasmatzis, G.; Kovtun, I.V. Frequently rearranged and overexpressed δ-catenin is responsible for low sensitivity of prostate cancer cells to androgen receptor and β-catenin antagonists. Oncotarget 2018, 9, 24428–24442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Huang, F.; Chen, J.; Wang, Z.; Lan, R.; Fu, L.; Zhang, L. δ-Catenin promotes tumorigenesis and metastasis of lung adenocarcinoma. Oncol. Rep. 2017, 39, 809–817. [Google Scholar] [CrossRef]
  95. Jin, Z.-L.; Pei, H.; Xu, Y.-H.; Yu, J.; Deng, T. The SUMO-specific protease SENP5 controls DNA damage response and promotes tumorigenesis in hepatocellular carcinoma. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 3566–3573. [Google Scholar]
  96. Cashman, R.; Cohen, H.; Ben-Hamo, R.; Zilberberg, A.; Efroni, S. SENP5 mediates breast cancer invasion via a TGFβRI SUMOylation cascade. Oncotarget 2014, 5, 1071–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kanwal, M.; Ding, X.-J.; Ma, Z.-H.; Li, L.-W.; Wang, P.; Chen, Y.; Huang, Y.-C.; Cao, Y. Characterization of germline mutations in familial lung cancer from the Chinese population. Gene 2018, 641, 94–104. [Google Scholar] [CrossRef]
  98. Cui, J.; Yin, Y.; Ma, Q.; Wang, G.; Olman, V.; Zhang, Y.; Chou, W.-C.; Hong, C.S.; Zhang, C.; Cao, S.; et al. Comprehensive characterization of the genomic alterations in human gastric cancer. Int. J. Cancer 2014, 137, 86–95. [Google Scholar] [CrossRef] [Green Version]
  99. Hu, C.; Zhou, Y.; Liu, C.; Kang, Y. Risk assessment model constructed by differentially expressed lncRNAs for the prognosis of glioma. Oncol. Rep. 2018, 40, 2467–2476. [Google Scholar] [CrossRef] [PubMed]
  100. Chauhan, R.; Shimizu, Y.; Watashi, K.; Wakita, T.; Fukasawa, M.; Michalak, T.I. Retrotransposon elements among initial sites of hepatitis B virus integration into human genome in the HepG2-NTCP cell infection model. Cancer Genet. 2019, 235–236, 39–56. [Google Scholar] [CrossRef]
  101. Nault, J.C.; Calderaro, J.; Di Tommaso, L.; Balabaud, C.; Zafrani, E.S.; Bioulac-Sage, P.; Roncalli, M.; Zucman-Rossi, J. Telomerase reverse transcriptase promoter mutation is an early somatic genetic alteration in the transformation of premalignant nodules in hepatocellular carcinoma on cirrhosis. Hepatology 2014, 60, 1983–1992. [Google Scholar] [CrossRef]
  102. Nault, J.C.; Mallet, M.; Pilati, C.; Calderaro, J.; Bioulac-Sage, P.; Laurent, C.; Laurent, A.; Cherqui, D.; Balabaud, C.; Zucman-Rossi, J. High frequency of telomerase reverse-transcriptase promoter somatic mutations in hepatocellular carcinoma and preneoplastic lesions. Nat. Commun. 2013, 4, 2218. [Google Scholar] [CrossRef] [Green Version]
  103. Nault, J.-C.; Zucman-Rossi, J. Genetics of hepatocellular carcinoma: The next generation. J. Hepatol. 2014, 60, 224–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Pinyol, R.; Tovar, V.; Llovet, J.M. TERT promoter mutations: Gatekeeper and driver of hepatocellular carcinoma. J. Hepatol. 2014, 61, 685–687. [Google Scholar] [CrossRef] [Green Version]
  105. Quaas, A.; Oldopp, T.; Tharun, L.; Klingenfeld, C.; Krech, T.; Sauter, G.; Grob, T.J. Frequency of TERT promoter mutations in primary tumors of the liver. Virchows Arch. 2014, 465, 673–677. [Google Scholar] [CrossRef] [PubMed]
  106. Totoki, Y.; Tatsuno, K.; Covington, K.R.; Ueda, H.; Creighton, C.J.; Kato, M.; Tsuji, S.; A Donehower, L.; Slagle, B.L.; Nakamura, H.; et al. Trans-ancestry mutational landscape of hepatocellular carcinoma genomes. Nat. Genet. 2014, 46, 1267–1273. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The top five significantly enriched gene ontology terms associated with RTGs as determined by the EnrichR software. (A) Biological processes, (B) molecular functions, and (C) Drug Signatures Database (DSigDB). Pathways are presented based on the combined EnrichR score. DSigDB relates drugs/compounds to their target genes.
Figure 1. The top five significantly enriched gene ontology terms associated with RTGs as determined by the EnrichR software. (A) Biological processes, (B) molecular functions, and (C) Drug Signatures Database (DSigDB). Pathways are presented based on the combined EnrichR score. DSigDB relates drugs/compounds to their target genes.
Cells 10 01294 g001
Figure 2. Distribution of integration breakpoints in the HBV genome in HCC tumor samples. A total of 3052 HBV breakpoints were plotted. The histogram represents the frequency of integration breakpoints at different loci in the HBV genome (nt. 1–3215; bin size 25) as numbered in the outer ring.
Figure 2. Distribution of integration breakpoints in the HBV genome in HCC tumor samples. A total of 3052 HBV breakpoints were plotted. The histogram represents the frequency of integration breakpoints at different loci in the HBV genome (nt. 1–3215; bin size 25) as numbered in the outer ring.
Cells 10 01294 g002
Figure 3. Mapping of TERT, MLL4, and PLEKH4G4B HBV integration breakpoints along the human and HBV genomes. (A) TERT breakpoints. A total of 219 TERT integration breakpoints derived from 161 patients are plotted. The y-axis coordinates decrease from 1320 kb to 1260 kb to represent the direction of the transcriptional start site from a 5′–3′ orientation. The expanded view of the region with the most integration sites is shown for the human genome position 1297 kb to 1294 kb and the HBV nt. 1500–2000. (B) MLL4 breakpoints. A total of 115 MLL4 integration breakpoints, derived from 64 patients, are plotted. Blue squares denoting exon regions are shown. (C) PLEKH4G4B breakpoints. A total of 47 of the 116 reported PLEKHG4B breakpoints plotted are derived from 8 unique HCC patients. Colored dots correspond to each unique patient. Each dot represents the mapped locations of the integration sites where the human gene breakpoints (GRCh37) are located on the y-axis, and HBV breakpoints are located on the x-axis, in accordance with the reported locations.
Figure 3. Mapping of TERT, MLL4, and PLEKH4G4B HBV integration breakpoints along the human and HBV genomes. (A) TERT breakpoints. A total of 219 TERT integration breakpoints derived from 161 patients are plotted. The y-axis coordinates decrease from 1320 kb to 1260 kb to represent the direction of the transcriptional start site from a 5′–3′ orientation. The expanded view of the region with the most integration sites is shown for the human genome position 1297 kb to 1294 kb and the HBV nt. 1500–2000. (B) MLL4 breakpoints. A total of 115 MLL4 integration breakpoints, derived from 64 patients, are plotted. Blue squares denoting exon regions are shown. (C) PLEKH4G4B breakpoints. A total of 47 of the 116 reported PLEKHG4B breakpoints plotted are derived from 8 unique HCC patients. Colored dots correspond to each unique patient. Each dot represents the mapped locations of the integration sites where the human gene breakpoints (GRCh37) are located on the y-axis, and HBV breakpoints are located on the x-axis, in accordance with the reported locations.
Cells 10 01294 g003
Figure 4. TERT gene alterations identified in HBV–HCC tissues. (A) In-house cohort (n = 22). (B) Compiled HBV–HCC cohort. Patients are derived from our in-house study (n = 22) and from the literature (n = 129) [40,42,45]. The number of HCC patients is indicated in parentheses.
Figure 4. TERT gene alterations identified in HBV–HCC tissues. (A) In-house cohort (n = 22). (B) Compiled HBV–HCC cohort. Patients are derived from our in-house study (n = 22) and from the literature (n = 129) [40,42,45]. The number of HCC patients is indicated in parentheses.
Cells 10 01294 g004
Table 1. Summary of HBV integration studies included in this analysis.
Table 1. Summary of HBV integration studies included in this analysis.
StudyHCC Patients
(n)
# of Subjects with Integrated DNA * (% of total)# of Junctions * Identified in SubjectsJunction SequenceClinical VariablesGeographic Region
NGS-basedWGS[21]33 (100%)15YesYesNA 7
[30,31]91 164 (45%)223YesYesJapan
[32,33]8176 (94%)344YesYesHong Kong
[34]22 (100%)5YesYesChina
[22]5 1,25 (100%)92Yes-NA 7
[35]55 (100%)21YesYesNA
[36]32 (67%)11YesYesChina
Whole HBV Genome[37]2016 (80%)424YesYesChina
[38]4826 (54%)57--Singapore
[39]6051 (85%)156YesYesChina
[6]426344 (81%)3486Yes3-China
[40]4928 (57%)121YesYesJapan
[41]4035 (90%)257YesYesNA
[42]10194 (93%)510YesYesTaiwan
[43]5454 (100%)2870YesYesChina
[44]1 51 (100%)2YesYesNA
[45]9585 (89%)1563YesYesHong Kong
[46]24372 (30%)4347YesYesJapan
[47]5044 (88%)289YesYesTaiwan
DR1-2[29]4026 (65%)42YesYesChina
Our study2220 (91%)27YesYesTaiwan
PCR-based[48]132 (15%)2Yes-USA
[49]1414 (100%)14YesYesChina
[50]1515 (100%)15-YesNA
[51]6055 (92%)60-YesNA
[52]107 (70%)8YesYesJapan
[53]6041 (68%)101YesYesChina
[54]59 445 (76%)45--Italy
[55]159 (60%)9--China
[56]1818 (100%)2083-YesChina
[57]3021 (70%)1397 6--China
Total17331310 (76%)18,596
1 HBV (+) HCC cohorts only. 2 Three patients previously reported in Jiang 2012 [21] were removed. 3 Only human chromosome sequence position provided. 4 Cohorts of HBsAg (−)/occult (+) and HBsAg (+) HCC patients. 5 HBV capture sequencing of a single occult HBV infection (OBI) patient. 6 Five of the thirty HCC tissues were analyzed by HBV-targeted NGS. 7 Specimens were purchased commercially from vendors such as Seracare LifeSciences, ProteoGenex, or Indivumed. * HBV DNA integration sites. WGS, whole-genome next-generation sequencing; Whole HBV genome, whole HBV genome enrichment performed prior to NGS; DR1-2, HBV DR1-2 integration hotspot region enriched prior to NGS; -, not available.
Table 2. Major sociodemographic and clinical features of the study subjects with and without detectable integrated HBV DNA in tumor tissue.
Table 2. Major sociodemographic and clinical features of the study subjects with and without detectable integrated HBV DNA in tumor tissue.
Reported HBV–HCC Population 1Integrated HBV DNA
Not Detectable
(n = 423)
Detectable
(n = 1310)
Age (years)
Range
Avg. ± SD
NA
55–65 ± NA
33–83
59.4 ± 13.0 (n = 40)
11–85
54.9 ± 11.6 (n = 359)
Gender (Total) (n = 66)(n = 635)
MaleNA40466
Female15169
Genotype (Total) (n = 2)(n = 84)
BNA024 2
C276
D01
Cirrhosis %70–90%
(n = NA)
46.7%
(n = 7/15)
62.3%
(n = 105/279)
1 Characteristics of the HBV–HCC subjects obtained from previous reports [4,58,59]. 2 One patient contained a mix of HBV B and C genotypes. NA, not available; n, number of patients available for analysis; Avg. ± SD, average ± standard deviation.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lin, S.Y.; Zhang, A.; Lian, J.; Wang, J.; Chang, T.-T.; Lin, Y.-J.; Song, W.; Su, Y.-H. Recurrent HBV Integration Targets as Potential Drivers in Hepatocellular Carcinoma. Cells 2021, 10, 1294. https://doi.org/10.3390/cells10061294

AMA Style

Lin SY, Zhang A, Lian J, Wang J, Chang T-T, Lin Y-J, Song W, Su Y-H. Recurrent HBV Integration Targets as Potential Drivers in Hepatocellular Carcinoma. Cells. 2021; 10(6):1294. https://doi.org/10.3390/cells10061294

Chicago/Turabian Style

Lin, Selena Y., Adam Zhang, Jessica Lian, Jeremy Wang, Ting-Tsung Chang, Yih-Jyh Lin, Wei Song, and Ying-Hsiu Su. 2021. "Recurrent HBV Integration Targets as Potential Drivers in Hepatocellular Carcinoma" Cells 10, no. 6: 1294. https://doi.org/10.3390/cells10061294

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop