Next Article in Journal
Feline Foamy Virus Transmission in Tsushima Leopard Cats (Prionailurus bengalensis euptilurus) on Tsushima Island, Japan
Next Article in Special Issue
Development of a Rapid Epstein–Barr Virus Detection System Based on Recombinase Polymerase Amplification and a Lateral Flow Assay
Previous Article in Journal
Canine Distemper Virus Alters Defense Responses in an Ex Vivo Model of Pulmonary Infection
Previous Article in Special Issue
Epstein–Barr Virus History and Pathogenesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epstein–Barr Virus B Cell Growth Transformation: The Nuclear Events

Department of Medicine, Division of Infectious Diseases, Brigham and Women’s Hospital and Harvard Medical School, 181 Longwood Avenue, Boston, MA 02115, USA
Viruses 2023, 15(4), 832; https://doi.org/10.3390/v15040832
Submission received: 8 February 2023 / Revised: 13 March 2023 / Accepted: 17 March 2023 / Published: 24 March 2023
(This article belongs to the Special Issue Epstein-Barr Virus Replication and Pathogenesis)

Abstract

:
Epstein–Barr virus (EBV) is the first human DNA tumor virus identified from African Burkitt’s lymphoma cells. EBV causes ~200,000 various cancers world-wide each year. EBV-associated cancers express latent EBV proteins, EBV nuclear antigens (EBNAs), and latent membrane proteins (LMPs). EBNA1 tethers EBV episomes to the chromosome during mitosis to ensure episomes are divided evenly between daughter cells. EBNA2 is the major EBV latency transcription activator. It activates the expression of other EBNAs and LMPs. It also activates MYC through enhancers 400–500 kb upstream to provide proliferation signals. EBNALP co-activates with EBNA2. EBNA3A/C represses CDKN2A to prevent senescence. LMP1 activates NF-κB to prevent apoptosis. The coordinated activity of EBV proteins in the nucleus allows efficient transformation of primary resting B lymphocytes into immortalized lymphoblastoid cell lines in vitro.
Keywords:
EBV; EBNA; LMP1

1. Introduction

Epstein–Barr virus (EBV) was discovered from African Burkitt’s lymphoma (BL) cells over fifty years ago [1]. EBV causes many different cancers including B cell lymphomas and Hodgkin’s lymphoma in people with T-cell immune suppression due to HIV infection, transplantation, or advanced age; nasopharyngeal carcinoma; and ~10% of gastric cancers [2]. During primary EBV infection, viruses replicate in oro-pharyngeal epithelial cells and infect resting B-lymphocytes (RBLs). Latency III EBV nuclear antigens, EBNA2, leader protein (EBNALP), EBNA3A, EBNA3C, and Latent Membrane Protein 1 are the principal causes of infected B cell proliferation. While immune responses eliminate most B cells expressing latency III proteins, escaped cells enter lymphatic tissues, where EBV protein expression is down-regulated and only EBNA1 is expressed. EBV infected cells that escape immune destruction are sites of long-term latency persistence, reactivation, and can evolve to EBV+ B cell lymphomas and Hodgkin’s lymphoma [3].
EBV immortalizes RBLs to lymphoblastoid cell lines (LCLs) in vitro. LCLs express type III EBV latency genes. Post-transplant lymphoproliferative disease (PTLD) and AIDS CNS lymphomas express the same latency III EBV genes as LCLs. Therefore, LCLs are a useful and appropriate model for genetic and biochemical investigation of EBV’s roles in converting RBLs to proliferating lymphoblasts in vivo.

2. EBV Transformation from RBL to LCL Timeline

EBV infection of RBLs in vitro results in the establishment of LCLs in 3–4 weeks. Once established, LCLs can be kept in culture and grow continuously. Two days after EBV infection of RBLs, the cell size starts to increase and continues to increase. By day 4, cell size reaches the peak and starts to decrease slowly. Eight days after infection, the cell size is approximately two times the size of RBLs [4,5]. EBNA mRNAs are expressed at LCL level by day 2 and maintain a similar level afterwards. LMP1 mRNA level starts to increase two days after infection and gradually increases to LCL level at day 28 [6,7,8]. EBNA2 protein expression level reaches the peak level two days post infection and then slowly decreases to the LCL expression level [9]. EBNA3A and 3C protein expression start at day two, but only reach high level at day 4 and maintain the expression at that level [9]. LMP1 protein starts to express 4 days post infection and continues to increase to LCL level at day 28 [8,9].

3. EBNA2 Is the Major EBV Transcription Activator

After EBV infection, EBNALP and EBNA2 are the first two EBV genes expressed [10]. EBNALP and EBNA2 mRNAs are generated by differential splicing from the same bicistronic RNA [11,12]. EBNA2 is absolutely required for EBV to immortalize RBLs [4,13,14]. EBNA2 is the major EBV latency transcription activator [15,16] and strongly activates the EBV Cp promoter that drives the expression of all the EBNAs and the LMP1 promoter [16,17]. EBNA2 also upregulates the expression of many cell genes including MYC [18,19,20,21,22,23,24].
There are two types of EBV, type I or II. Type I EBNA2 from B95.8 strain has 487 amino acids (AAs) [25]. The EBNA2 N-terminus has two dimerization/transcription activation domains separated by a stretch of proline residues [26,27,28]. EBNA2 AAs 323–324 bind cell protein RBPJ [29,30]. The EBNA2 C-terminal region contains the major transactivation domain [31]. The RBPJ binding and the transactivation domains are required for EBV transformation [32]. Type II EBNA2 lacks the long stretch of proline repeats at the N-terminus [14]. Recombinant EBV with type II EBNA2 has lower transforming efficiency compared with virus with type I EBNA2 [14]. The reduced transformation activity is caused by S442 in the type II EBNA2 transactivation domain. Substituting the serine residue with aspartic acid that is in the type I EBV AA sequence increases the type II EBNA2 activity [33]. Increased binding to transcription repressor BS69 is also important for EBNA2 transformation activity [33,34,35]. EBNA2 is tethered to host DNA through interactions with host transcription factors (TFs). RBPJ is a major cell protein that tethers EBNA2 to DNA [29,30,36]. In T cells, RBPJ tethers Notch intracellular domain to DNA [37]. RBPJ is a transcription repressor that recruits other repressors to DNA and EBNA2 can mask the repressors to activate gene expression [38,39]. EBNA2 also increases RBPJ DNA binding [40]. EBNA2 interacts with other host TFs that can tether EBNA2 to DNA. Mutation of the SPI1 site or EBF site of the LMP1 promoter greatly reduces EBNA2 activation of the reporter [19,41]. EBNA2 binds to SPI1 and EBF and enhances EBF DNA binding [42,43]. The C-terminal transactivation domain recruits basal transcription factors including EP300/CBP, SNF/SWI complex components, TFIIH subunits, TFIIB, and TFIIE to activate transcription [44,45,46,47,48,49]. In addition, EBNA2 binds to NPM1 and PRMT5 to enhance transcription activation [50,51].
Chromatin Immune Precipitation followed by deep sequencing (ChIP-seq) identifies thousands of EBNA2 binding sites in LCLs or BL cell lines [19,52,53]. EBNA2 mostly binds to enhancers. EBNA2 binding sites overlap extensively with RBPJ binding sites. Furthermore, EBNA2 sites overlap with other B cell TFs, including EBF, ETS family TFs, RUNX3, and NF-κB family TFs in LCLs. The centers of EBNA2 binding sites have much lower H3K4me1, a histone modification marker for active promoters and enhancers, than the regions immediately adjacent to the binding site. Interestingly, these sites in RBLs are in a similar pattern, though at smaller elevation. These data suggest that EBV exploits the preexisting TF binding sites to activate host gene expression [19]. EBNA2 binds to enhancers ~556~428 kb upstream of MYC. Fluorescence in situ hybridization (FISH) and chromatin conformation capture followed by qPCR (3C-qPCR) find that these enhancers interact with MYC TSS by looping out the intervening DNA sequences between enhancers and promoters [19] (Figure 1). EBNA2 conditional inactivation greatly decreases the enhancer–promoter interaction. RNA POLII Chromatin Interaction Analysis by Paired-End Tag Sequencing (ChIA-PET) identifies all genomic interactions mediated by RNA POLII, including enhancer–enhancer, enhancer–promoter, and promoter–promoter interactions. LCL RNA POLII ChIA-PET links all the EBNA2 enhancers to their direct target genes genome-wide [54]. Most importantly, ChIA-PET links EBNA2 enhancers to MYC TSS, in agreement with data generated by capture genome-wide 3C followed by deep sequencing (C Hi-C) and circular 3C followed by deep sequencing (4C-seq) [55]. The EBNA2 enhancer interaction patterns are extremely complicated. These enhancers are found both upstream and downstream of their direct target genes. Some enhancers interact with only one target gene. Some genes interact with multiple enhancers. Some enhancers interact with multiple genes. Some enhancers skip the nearest gene to interact with genes further away. CRISPR disruption of RBPJ sites within the EBNA2 enhancer greatly reduces the mRNA levels of EBNA2-enhancer-linked genes [56].
Super-enhancers (SEs) are clusters of enhancers with extraordinary broad and high active histone marks, including H3K27ac and H3K4me1 or TF ChIP-seq peaks [57]. SEs play critical roles in development, differentiation, and oncogenesis [57]. EBNA2 SEs are enhancers with extraordinary broad and tall EBNA2 ChIP-seq peaks [54]. EBNA2 SEs control the expression of cell genes that are critically important for cell proliferation, including MYC, MAX, RUNX3, BCL2, EBF, BATF, etc. [54,58]. CRISPR deletion of MYC SE using paired sgRNAs targeting the edges of EBNA2 -525 SE greatly reduces MYC expression and cell growth in LCLs [56]. In addition to EBNA2 binding, EBNA2 SEs are also bound by many B cell TFs, including BRD4, EP300, RNA POLII, BATF, EBF, IRF4, etc. [54]. The enrichment of activated and basal TFs in SEs facilitates the phase separation that allows the formation of condensates that robustly increase the transcription activity [59,60].

4. EBNALP in Transcription Regulation

The AA sequence of EBNALP consists of various numbers of identical W repeats and unique Y exons in different EBV isolates. Two copies of W repeats are the minimum for EBNALP to function [61]. Four copies of W repeats are enough for EBNALP to exert optimum activity [62]. A genetic study inserting a stop codon to the end of the last W repeat indicated that EBNALP is important for EBV transformation of RBLs [63]. Mutant EBV with stop codons inserted into every W exon fails to transform naïve B cells from the cord blood [64], indicating that EBNALP is essential for EBV transformation of naïve B cells. This mutant EBV transforms memory B cells with reduced efficiency [4,64].
EBNALP increases the expression of cell cycle progression protein CCND2 [65,66]. EBNALP strongly coactivates EBNA2 transcription [62,67]. EBNALP lacks a transactivation domain in its AA sequence. It regulates transcription through removing transcription repressors such as NCOR and HADC4 from the promoters/enhancers it co-activates with EBNA2 and shuttling Sp100 between different subnuclear localizations [40,68,69]. EBNALP also greatly increases EP300 transcription activity when EP300 is tethered to DNA through the papilloma virus E2 DNA binding domain [70]. EBNALP coactivation is mostly through the EBNA2 N-terminal and C-terminal transactivation domains. EBNALP expressed in bacteria also binds to the EBNA2 transactivation domain [27]. Evolutionary conserved residues in the W repeats of EBNALP are important for coactivation [61,71]. EBNALP also binds to many cell proteins, including HA95, HAX1, HSP72, etc. to affect transcription [72,73,74].
EBNALP does not bind to DNA directly, instead it is tethered to DNA through host proteins. EBNALP ChIP-seq finds that EBNALP binds to thousands of promoter and enhancer sites [75]. Interestingly, ~33% of EBNALP binding sites are at promoters. In contrast, only ~14% of the EBNA2 binding sites are at promoters [19]. EBNALP peaks are also different from EBNA2 peaks that are very sharp and tall. EBNALP peaks tend to be much wider and lower. Integrating the ENCODE GM12878 ChIP-seq data with EBNALP data finds that 82% of the EBNALP peaks overlap with DPF2 ChIP-seq peaks. DPF2 binds to H3K14ac and H4K16ac and can bridge EBNALP to chromatin [76]. Motif analysis identifies that the motifs enrich at the EBNALP ChIP-seq peaks, including CTCF, ETS, IRF4, YY1. A total of 59% of EBNALP peaks overlaps with YY1 peaks in LCLs [75]. YY1 also binds preferentially to promoters, and it bridges promoter–enhancer interactions [77]. The high degree of overlap between EBNALP and YY1 peaks in LCLs suggests that EBNALP may play important role in mediating host genome organization (Figure 1).

5. EBNA3A and EBNA3C

EBNA3A, 3B, and 3C are located in tandem in the EBV genome. They all have similar gene structure with a short exon and a long exon. It is believed that they are the products of gene duplication. They share limited homology in their N-terminal AA sequences that mediate their interactions with host TF RBPJ that tethers EBNA2 to DNA [78,79,80]. Genetic studies indicate that EBNA3B is dispensable while EBNA3A is important and EBNA3C is essential for LCL growth [4,81,82,83,84]. EBNA3A AAs 170–240 and 300–523 are important for LCL growth [85]. EBNA3C AAs 50–400 and 800–900 are essential for LCL growth [86,87].
EBNA3A specifically interacts with RBPJ and prevents RBPJ binding to its cognate DNA in vitro, thus repressing EBNA2-RBPJ transcription activation [78,79,88,89]. EBNA3A-RBPJ interaction is required for EBNA3A to support LCL growth. The EBNA3A mutant, defective for RBPJ binding, fails to rescue EBNA3A inactivation in LCLs [85]. Conditional overexpression of EBNA3A induces LCL growth arrest and reduced EBNA2- RBPJ association [90]. When EBNA3A is fused to a GAL4 DNA binding domain, the fusion protein represses a reporter driven by GAL4 binding sites [91]. EBNA3A also interacts with CTBP and WDR48/USP46/USP12 deubiquitination complexes [92,93]. Transcription profiling of wild-type and EBNA3A mutant LCLs identifies many host genes repressed or activated by EBNA3A [94]. EBNA3A represses the expression of BCL2L11(BIM), CXCL9, CXCL10, MIR221/222, p16INK4A, p14ARF, CDKN2B, CDKN1A, and CDKN2C [95,96,97,98,99,100,101]. shRNA knockdown of p16INK4A and p14ARF allows LCLs to grow in the absence of EBNA3A [100]. EBNA3A represses BCL2L11 expression through CpG methylation and increased H3K27me3 repressive histone modification by polycomb repressive complex (PRC) proteins SUS12 and BMI1 [102,103]. EBNA3A increases H3K27me3 at the p16INK4A and p14ARF loci and the repression requires the interaction of EBNA3A with CTBP [104]. EBNA3A is also required for MCL-1 mitochondrial localization and activates BFL-1 transcription [105]. EBNA3A synergizes with LMP1 to induce B cell lymphoma formation in LMP1 transgenic mice and inhibits differentiation [106]. Genome-wide, EBNA3A binds to thousands of enhancer and promoter sites including MYC, CCND2, p16INK4A, p14ARF, and BCL2 [107,108,109]. Around 50% of EBNA3A sites overlap with EBNA3C sites, and 65% and 63% of EBNA3A sites overlap with BATF and RUNX3 sites. EBNA3A is tethered to DNA through BATF, shown by ChIP-re-ChIP [107].
Repression of p16INK4A is the most important function of EBNA3C. shRNA knockdown of p16INK4A and p14ARF allows LCLs expressing conditional EBNA3C to grow in the non-permissive condition. EBNA3C inactivation does not affect the DNA methylation profiles at the loci but increases H3K27me3 and decreases H3K4me1 [84,100]. EBNA3C null virus can transform B cells from primary B cells from an individual with homozygous deletion of p16INK4a [110]. CRISPR knockout of p16INK4A also allows LCLs to grow in the absence of EBNA3C [111]. EBNA3C binds to p14ARF promoter and recruits Sin3A repressor to the site. The Sin3A recruitment is dependent on EBNA3C expression [112]. EBNA3C interactions with RBPJ, CtBP, WDR48, and USP46 are also important for EBNA3C-mediated p16INK4A repression [92,93,111] (Figure 1).
EBNA3C maintains the LMP1 expression level in growth-arrested Raji cells [113]. EBNA3C co-activates the LMP1 promoter together with EBNA2 [114,115]. EBNA3C upregulates the expression of AID that is important for somatic hypermutation and class switch [116,117]. EBNA3C upregulates CXCL12 and CXCR4 expression in an RBPJ-dependent way [117]. EBNA3C represses the expression of BCL2L11(BIM) by recruiting PRC components SUZ12 and EZH2 to the regulatory elements and enhancer reorganization [55,118]. EBNA3C recruits RBPJ, SUZ12, and BMI1 to repress ADAMDEC1 and ADAM28 [119]. EBNA3C also induces the expression of miRNAs miR-221/222 [99].
Genome wide, EBNA3C binds to thousands of enhancers and promoters in LCLs [108,112]. EBNA3C binding sites coincide with RUNX3, BATF, ATF2, and IRF4. EBNA3C signals are strongest at BATF/IRF4 and SPI1/IRF4 composite sites. EBNA3C binds strongly to the p14ARF promoter through SPI1/IRF4/BATF/RUNX3 host TFs and recruits transcription repressors Sin3A and REST family members in an EBNA3C-dependent manner [112].
EBNA3C specifically interacts with RBPJ and the interaction is required for continuous LCL growth [79,80,86,87,120]. The EBNA3C N-terminal homology domain binds to RBPJ. EBNA3C prevents RBPJ from binding to its cognate sites in vitro, but it can also recruit RBPJ to DNA at some sites in vivo. However, the genome-wide effect of EBNA3C on RBPJ DNA binding is still not known [79,88,119]. EBNA3C interacts with IRF4 and IRF8 through its N-terminus and stabilizes IRF4 [121]. IRF4 is essential for LCL growth [122] and may tether EBNA3C to DNA [112]. EBNA3C promotes proteasome-mediated degradation of IRF8 [121]. EBNA3C specifically modulates cell-cycle-regulating proteins. EBNA3C amino acids 130 to 159 bind CCNA and enhance CCNA-dependent kinase activity [123,124]. These amino acids are required for LCL growth [86,87,125]. EBNA3C stabilizes CCND2 to regulate cell cycle progression [126]. EBNA3C also stabilizes CCND1 through inhibition of its polyubiquitination and EBNA3C enhances the kinase activity of Cyclin D1/CDK6 [127] (Figure 1). EBNA3C amino acids 130–160 bind to two different sites on CCND1. EBNA3C associates with the CCND1/cdk2 complex and enhances the kinase activity. EBNA3C recruits SCFSkp2 activity to CCNA complexes [128]. Through Skp2, EBNA3C also degrades tumor suppressor RB [129]. EBNA3C amino acids 140–149 are important for both the binding and regulation of RB [129]. EBNA3C inhibits the function of p53 through interaction with ING4/5 [130]. EBNA3C can modulate the PIM1 kinase activity [131]. EBNA3C amino acids 365 to 545 are important for EBNA2 coactivation of the LMP1 promoter and interact with SUMO-1 and SUMO-3 [132]. EBNA3C recruitment of histone deacetylase activity and association with the corepressors Sin3A and NCoR are likely to be important for EBNA3C-mediated transcription repression [133].

6. EBNA1

EBNA1 is expressed in all EBV-infected cells. EBNA1 is essential for EBV genome replication, persistence, and transcription [134,135,136,137,138,139]. EBNA1 tethers EBV episomes to the mitotic chromosomes during mitosis to ensure even distribution of episomes into daughter cells [136]. ChIP and microarray identified EBNA1 binding sequences in vitro [140]. ChIP-seq identified EBNA1 binding sites in vivo [141,142]. EBNA1 binding to MEF2B, IL6R, and EBF1 is critical for B cell survival [142]. During EBV transformation of RBLs to LCLs, EBNA1 binds to enhancers linked to genes involved in nucleotide metabolism [143]. EBNA1 forms a DNA protein crosslink with the EBV origin of replication oriP [144]. The protein–DNA crosslinking enables OriP replication termination to promote episome maintenance [144]. A small molecule inhibitor, that selectively inhibits EBNA1 DNA binding, prevents EBV infection and tumor growth [145]. This EBNA1 inhibitor also inhibits gastric cancer growth [146]. Another EBNA1 small molecule has also been reported to perturb episome replication and persistence [147].

7. LMP1

LMP1 activates canonical and non-canonical NF-κB pathways [148]. NF-κB family TFs have five subunits that can form a heterodimer or homodimers and bind to DNA [149]. These proteins are each critical for B cell development and function. NF-κB subunit ChIP-seq identifies a complex NF-κB-binding landscape in LCLs. Nearly one-third of NF-κB-binding sites lack κB motifs and are instead enriched for alternative motifs. The oncogenic forkhead box protein FOXM1 co-occupies nearly half of NF-κB-binding sites and forms protein complexes with NF-κB on DNA. FOXM1 knockdown decreases NF-κB target gene expression and ultimately induces apoptosis, highlighting FOXM1 as a synthetic lethal target in B cell malignancy [149].

8. Future Directions

The expression of viral transcription factors in EBV cancer provides a unique opportunity for targeted therapies. Traditionally, TFs are not ideal drug targets. However, the development of proteolysis targeting chimeras (PROTACs) now allows specific degradation of TFs [150]. A small molecule that binds to TFs is linked to E3 ligase ligand. The molecule then recruits E3 ligase to the TF and degrades the TF. PROTACs have been developed to target many TFs and the degradation of TFs important for cancer cells can inhibit cell growth [151,152,153]. It is therefore now possible to develop EBV-specific PROTACs to degrade EBV TFs to treat EBV-associated cancers.

Funding

This work was supported by NIAID AI123420 and NCI CA047006.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

I thank Elliott Kieff and Ben Gewurz for insightful discussions. I thank Chong Wang for sharing unpublished data. I thank Katelyn Ru-Shin Zhao for generating the figure.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Epstein, M.A.; Achong, B.G.; Barr, Y.M. Virus Particles in Cultured Lymphoblasts from Burkitt’s Lymphoma. Lancet 1964, 1, 702–703. [Google Scholar] [CrossRef] [PubMed]
  2. Cohen, J.I.; Fauci, A.S.; Varmus, H.; Nabel, G.J. Epstein-Barr virus: An important vaccine target for cancer prevention. Sci. Transl. Med. 2011, 3, 107fs7. [Google Scholar] [CrossRef] [Green Version]
  3. Longnecker, R.; Kieff, E.; Cohen, J.I. Epstein-Barr Virus, 8th ed.; Lippincott Williams & Wilkins, a Wolters Kluwer Business: Philadelphia, PA, USA, 2013; Volume 2, pp. 1898–1959. [Google Scholar]
  4. Pich, D.; Mrozek-Gorska, P.; Bouvet, M.; Sugimoto, A.; Akidil, E.; Grundhoff, A.; Hamperl, S.; Ling, P.D.; Hammerschmidt, W. First Days in the Life of Naive Human B Lymphocytes Infected with Epstein-Barr Virus. mBio 2019, 10, e01723-19. [Google Scholar] [CrossRef] [Green Version]
  5. Wang, L.W.; Wang, Z.; Ersing, I.; Nobre, L.; Guo, R.; Jiang, S.; Trudeau, S.; Zhao, B.; Weekes, M.P.; Gewurz, B.E. Epstein-Barr virus subverts mevalonate and fatty acid pathways to promote infected B-cell proliferation and survival. PLoS Pathog. 2019, 15, e1008030. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, C.; Li, D.; Zhang, L.; Jiang, S.; Liang, J.; Narita, Y.; Hou, I.; Zhong, Q.; Zheng, Z.; Xiao, H.; et al. RNA Sequencing Analyses of Gene Expression during Epstein-Barr Virus Infection of Primary B Lymphocytes. J. Virol. 2019, 93, e00226-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Mrozek-Gorska, P.; Buschle, A.; Pich, D.; Schwarzmayr, T.; Fechtner, R.; Scialdone, A.; Hammerschmidt, W. Epstein-Barr virus reprograms human B lymphocytes immediately in the prelatent phase of infection. Proc. Natl. Acad. Sci. USA 2019, 116, 16046–16055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Price, A.M.; Tourigny, J.P.; Forte, E.; Salinas, R.E.; Dave, S.S.; Luftig, M.A. Analysis of Epstein-Barr virus-regulated host gene expression changes through primary B-cell outgrowth reveals delayed kinetics of latent membrane protein 1-mediated NF-kappaB activation. J. Virol. 2012, 86, 11096–11106. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, L.W.; Shen, H.; Nobre, L.; Ersing, I.; Paulo, J.A.; Trudeau, S.; Wang, Z.; Smith, N.A.; Ma, Y.; Reinstadler, B.; et al. Epstein-Barr-Virus-Induced One-Carbon Metabolism Drives B Cell Transformation. Cell Metab. 2019, 30, 539–555.e11. [Google Scholar] [CrossRef] [Green Version]
  10. Alfieri, C.; Birkenbach, M.; Kieff, E. Early events in Epstein-Barr virus infection of human B lymphocytes. Virology 1991, 181, 595–608. [Google Scholar] [CrossRef]
  11. Sample, J.; Hummel, M.; Braun, D.; Birkenbach, M.; Kieff, E. Nucleotide sequences of mRNAs encoding Epstein-Barr virus nuclear proteins: A probable transcriptional initiation site. Proc. Natl. Acad. Sci. USA 1986, 83, 5096–5100. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, F.; Petti, L.; Braun, D.; Seung, S.; Kieff, E. A bicistronic Epstein-Barr virus mRNA encodes two nuclear proteins in latently infected, growth-transformed lymphocytes. J. Virol. 1987, 61, 945–954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hammerschmidt, W.; Sugden, B. Genetic analysis of immortalizing functions of Epstein-Barr virus in human B lymphocytes. Nature 1989, 340, 393–397. [Google Scholar] [CrossRef]
  14. Cohen, J.I.; Wang, F.; Mannick, J.; Kieff, E. Epstein-Barr virus nuclear protein 2 is a key determinant of lymphocyte transformation. Proc. Natl. Acad. Sci. USA 1989, 86, 9558–9562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wang, F.; Gregory, C.D.; Rowe, M.; Rickinson, A.B.; Wang, D.; Birkenbach, M.; Kikutani, H.; Kishimoto, T.; Kieff, E. Epstein-Barr virus nuclear antigen 2 specifically induces expression of the B-cell activation antigen CD23. Proc. Natl. Acad. Sci. USA 1987, 84, 3452–3456. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, F.; Tsang, S.F.; Kurilla, M.G.; Cohen, J.I.; Kieff, E. Epstein-Barr virus nuclear antigen 2 transactivates latent membrane protein LMP1. J. Virol. 1990, 64, 3407–3416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Rooney, C.M.; Brimmell, M.; Buschle, M.; Allan, G.; Farrell, P.J.; Kolman, J.L. Host cell and EBNA-2 regulation of Epstein-Barr virus latent-cycle promoter activity in B lymphocytes. J. Virol. 1992, 66, 496–504. [Google Scholar] [CrossRef] [Green Version]
  18. Kaiser, C.; Laux, G.; Eick, D.; Jochner, N.; Bornkamm, G.W.; Kempkes, B. The proto-oncogene c-myc is a direct target gene of Epstein-Barr virus nuclear antigen 2. J. Virol. 1999, 73, 4481–4484. [Google Scholar] [CrossRef] [Green Version]
  19. Zhao, B.; Zou, J.; Wang, H.; Johannsen, E.; Peng, C.W.; Quackenbush, J.; Mar, J.C.; Morton, C.C.; Freedman, M.L.; Blacklow, S.C.; et al. Epstein-Barr virus exploits intrinsic B-lymphocyte transcription programs to achieve immortal cell growth. Proc. Natl. Acad. Sci. USA 2011, 108, 14902–14907. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, B.; Maruo, S.; Cooper, A.; Chase, M.R.; Johannsen, E.; Kieff, E.; Cahir-McFarland, E. RNAs induced by Epstein-Barr virus nuclear antigen 2 in lymphoblastoid cell lines. Proc. Natl. Acad. Sci. USA 2006, 103, 1900–1905. [Google Scholar] [CrossRef] [Green Version]
  21. Schlee, M.; Krug, T.; Gires, O.; Zeidler, R.; Hammerschmidt, W.; Mailhammer, R.; Laux, G.; Sauer, G.; Lovric, J.; Bornkamm, G.W. Identification of Epstein-Barr virus (EBV) nuclear antigen 2 (EBNA2) target proteins by proteome analysis: Activation of EBNA2 in conditionally immortalized B cells reflects early events after infection of primary B cells by EBV. J. Virol. 2004, 78, 3941–3952. [Google Scholar] [CrossRef] [Green Version]
  22. Kempkes, B.; Spitkovsky, D.; Jansen-Durr, P.; Ellwart, J.W.; Kremmer, E.; Delecluse, H.J.; Rottenberger, C.; Bornkamm, G.W.; Hammerschmidt, W. B-cell proliferation and induction of early G1-regulating proteins by Epstein-Barr virus mutants conditional for EBNA2. EMBO J. 1995, 14, 88–96. [Google Scholar] [CrossRef] [PubMed]
  23. Spender, L.C.; Lucchesi, W.; Bodelon, G.; Bilancio, A.; Karstegl, C.E.; Asano, T.; Dittrich-Breiholz, O.; Kracht, M.; Vanhaesebroeck, B.; Farrell, P.J. Cell target genes of Epstein-Barr virus transcription factor EBNA-2: Induction of the p55alpha regulatory subunit of PI3-kinase and its role in survival of EREB2.5 cells. J. Gen. Virol. 2006, 87, 2859–2867. [Google Scholar] [CrossRef] [PubMed]
  24. Maier, S.; Staffler, G.; Hartmann, A.; Hock, J.; Henning, K.; Grabusic, K.; Mailhammer, R.; Hoffmann, R.; Wilmanns, M.; Lang, R.; et al. Cellular target genes of Epstein-Barr virus nuclear antigen 2. J. Virol. 2006, 80, 9761–9771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Dambaugh, T.; Hennessy, K.; Chamnankit, L.; Kieff, E. U2 region of Epstein-Barr virus DNA may encode Epstein-Barr nuclear antigen 2. Proc. Natl. Acad. Sci. USA 1984, 81, 7632–7636. [Google Scholar] [CrossRef] [Green Version]
  26. Harada, S.; Yalamanchili, R.; Kieff, E. Epstein-Barr virus nuclear protein 2 has at least two N-terminal domains that mediate self-association. J. Virol. 2001, 75, 2482–2487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Peng, C.W.; Zhao, B.; Kieff, E. Four EBNA2 domains are important for EBNALP coactivation. J. Virol. 2004, 78, 11439–11442. [Google Scholar] [CrossRef] [Green Version]
  28. Friberg, A.; Thumann, S.; Hennig, J.; Zou, P.; Nossner, E.; Ling, P.D.; Sattler, M.; Kempkes, B. The EBNA-2 N-Terminal Transactivation Domain Folds into a Dimeric Structure Required for Target Gene Activation. PLoS Pathog. 2015, 11, e1004910. [Google Scholar] [CrossRef] [Green Version]
  29. Ling, P.D.; Rawlins, D.R.; Hayward, S.D. The Epstein-Barr virus immortalizing protein EBNA-2 is targeted to DNA by a cellular enhancer-binding protein. Proc. Natl. Acad. Sci. USA 1993, 90, 9237–9241. [Google Scholar] [CrossRef] [Green Version]
  30. Grossman, S.R.; Johannsen, E.; Tong, X.; Yalamanchili, R.; Kieff, E. The Epstein-Barr virus nuclear antigen 2 transactivator is directed to response elements by the J kappa recombination signal binding protein. Proc. Natl. Acad. Sci. USA 1994, 91, 7568–7572. [Google Scholar] [CrossRef] [Green Version]
  31. Cohen, J.I.; Kieff, E. An Epstein-Barr virus nuclear protein 2 domain essential for transformation is a direct transcriptional activator. J. Virol. 1991, 65, 5880–5885. [Google Scholar] [CrossRef] [Green Version]
  32. Cohen, J.I.; Wang, F.; Kieff, E. Epstein-Barr virus nuclear protein 2 mutations define essential domains for transformation and transactivation. J. Virol. 1991, 65, 2545–2554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Tzellos, S.; Correia, P.B.; Karstegl, C.E.; Cancian, L.; Cano-Flanagan, J.; McClellan, M.J.; West, M.J.; Farrell, P.J. A single amino acid in EBNA-2 determines superior B lymphoblastoid cell line growth maintenance by Epstein-Barr virus type 1 EBNA-2. J. Virol. 2014, 88, 8743–8753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ponnusamy, R.; Khatri, R.; Correia, P.B.; Wood, C.D.; Mancini, E.J.; Farrell, P.J.; West, M.J. Increased association between Epstein-Barr virus EBNA2 from type 2 strains and the transcriptional repressor BS69 restricts EBNA2 activity. PLoS Pathog. 2019, 15, e1007458. [Google Scholar] [CrossRef]
  35. Harter, M.R.; Liu, C.D.; Shen, C.L.; Gonzalez-Hurtado, E.; Zhang, Z.M.; Xu, M.; Martinez, E.; Peng, C.W.; Song, J. BS69/ZMYND11 C-Terminal Domains Bind and Inhibit EBNA2. PLoS Pathog. 2016, 12, e1005414. [Google Scholar] [CrossRef] [Green Version]
  36. Henkel, T.; Ling, P.D.; Hayward, S.D.; Peterson, M.G. Mediation of Epstein-Barr virus EBNA2 transactivation by recombination signal-binding protein J kappa. Science 1994, 265, 92–95. [Google Scholar] [CrossRef]
  37. Wang, H.; Zou, J.; Zhao, B.; Johannsen, E.; Ashworth, T.; Wong, H.; Pear, W.S.; Schug, J.; Blacklow, S.C.; Arnett, K.L.; et al. Genome-wide analysis reveals conserved and divergent features of Notch1/RBPJ binding in human and murine T-lymphoblastic leukemia cells. Proc. Natl. Acad. Sci. USA 2011, 108, 14908–14913. [Google Scholar] [CrossRef] [Green Version]
  38. Zhou, S.; Hayward, S.D. Nuclear localization of CBF1 is regulated by interactions with the SMRT corepressor complex. Mol. Cell. Biol. 2001, 21, 6222–6232. [Google Scholar] [CrossRef] [Green Version]
  39. Hsieh, J.J.; Hayward, S.D. Masking of the CBF1/RBPJ kappa transcriptional repression domain by Epstein-Barr virus EBNA2. Science 1995, 268, 560–563. [Google Scholar] [CrossRef] [PubMed]
  40. Portal, D.; Zhao, B.; Calderwood, M.A.; Sommermann, T.; Johannsen, E.; Kieff, E. EBV nuclear antigen EBNALP dismisses transcription repressors NCoR and RBPJ from enhancers and EBNA2 increases NCoR-deficient RBPJ DNA binding. Proc. Natl. Acad. Sci. USA 2011, 108, 7808–7813. [Google Scholar] [CrossRef] [Green Version]
  41. Johannsen, E.; Koh, E.; Mosialos, G.; Tong, X.; Kieff, E.; Grossman, S.R. Epstein-Barr virus nuclear protein 2 transactivation of the latent membrane protein 1 promoter is mediated by J kappa and PU.1. J. Virol. 1995, 69, 253–262. [Google Scholar] [CrossRef] [Green Version]
  42. Yue, W.; Davenport, M.G.; Shackelford, J.; Pagano, J.S. Mitosis-specific hyperphosphorylation of Epstein-Barr virus nuclear antigen 2 suppresses its function. J. Virol. 2004, 78, 3542–3552. [Google Scholar] [CrossRef] [Green Version]
  43. Beer, S.; Wange, L.E.; Zhang, X.; Kuklik-Roos, C.; Enard, W.; Hammerschmidt, W.; Scialdone, A.; Kempkes, B. EBNA2-EBF1 complexes promote MYC expression and metabolic processes driving S-phase progression of Epstein-Barr virus-infected B cells. Proc. Natl. Acad. Sci. USA 2022, 119, e2200512119. [Google Scholar] [CrossRef]
  44. Tong, X.; Drapkin, R.; Reinberg, D.; Kieff, E. The 62- and 80-kDa subunits of transcription factor IIH mediate the interaction with Epstein-Barr virus nuclear protein 2. Proc. Natl. Acad. Sci. USA 1995, 92, 3259–3263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Tong, X.; Drapkin, R.; Yalamanchili, R.; Mosialos, G.; Kieff, E. The Epstein-Barr virus nuclear protein 2 acidic domain forms a complex with a novel cellular coactivator that can interact with TFIIE. Mol. Cell. Biol. 1995, 15, 4735–4744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Tong, X.; Wang, F.; Thut, C.J.; Kieff, E. The Epstein-Barr virus nuclear protein 2 acidic domain can interact with TFIIB, TAF40, and RPA70 but not with TATA-binding protein. J. Virol. 1995, 69, 585–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Wang, L.; Grossman, S.R.; Kieff, E. Epstein-Barr virus nuclear protein 2 interacts with p300, CBP, and PCAF histone acetyltransferases in activation of the LMP1 promoter. Proc. Natl. Acad. Sci. USA 2000, 97, 430–435. [Google Scholar]
  48. Wu, D.Y.; Kalpana, G.V.; Goff, S.P.; Schubach, W.H. Epstein-Barr virus nuclear protein 2 (EBNA2) binds to a component of the human SNF-SWI complex, hSNF5/Ini1. J. Virol. 1996, 70, 6020–6028. [Google Scholar] [CrossRef] [Green Version]
  49. Chabot, P.R.; Raiola, L.; Lussier-Price, M.; Morse, T.; Arseneault, G.; Archambault, J.; Omichinski, J.G. Structural and functional characterization of a complex between the acidic transactivation domain of EBNA2 and the Tfb1/p62 subunit of TFIIH. PLoS Pathog. 2014, 10, e1004042. [Google Scholar] [CrossRef] [PubMed]
  50. Liu, C.D.; Chen, Y.L.; Min, Y.L.; Zhao, B.; Cheng, C.P.; Kang, M.S.; Chiu, S.J.; Kieff, E.; Peng, C.W. The nuclear chaperone nucleophosmin escorts an Epstein-Barr Virus nuclear antigen to establish transcriptional cascades for latent infection in human B cells. PLoS Pathog. 2012, 8, e1003084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Liu, C.D.; Cheng, C.P.; Fang, J.S.; Chen, L.C.; Zhao, B.; Kieff, E.; Peng, C.W. Modulation of Epstein-Barr virus nuclear antigen 2-dependent transcription by protein arginine methyltransferase 5. Biochem. Biophys. Res. Commun. 2013, 430, 1097–1102. [Google Scholar] [CrossRef] [Green Version]
  52. Lu, F.; Chen, H.S.; Kossenkov, A.V.; DeWispeleare, K.; Won, K.J.; Lieberman, P.M. EBNA2 Drives Formation of New Chromosome Binding Sites and Target Genes for B-Cell Master Regulatory Transcription Factors RBP-jkappa and EBF1. PLoS Pathog. 2016, 12, e1005339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. McClellan, M.J.; Wood, C.D.; Ojeniyi, O.; Cooper, T.J.; Kanhere, A.; Arvey, A.; Webb, H.M.; Palermo, R.D.; Harth-Hertle, M.L.; Kempkes, B.; et al. Modulation of enhancer looping and differential gene targeting by Epstein-Barr virus transcription factors directs cellular reprogramming. PLoS Pathog. 2013, 9, e1003636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhou, H.; Schmidt, S.C.; Jiang, S.; Willox, B.; Bernhardt, K.; Liang, J.; Johannsen, E.C.; Kharchenko, P.; Gewurz, B.E.; Kieff, E.; et al. Epstein-Barr virus oncoprotein super-enhancers control B cell growth. Cell Host Microbe 2015, 17, 205–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Wood, C.D.; Veenstra, H.; Khasnis, S.; Gunnell, A.; Webb, H.M.; Shannon-Lowe, C.; Andrews, S.; Osborne, C.S.; West, M.J. MYC activation and BCL2L11 silencing by a tumour virus through the large-scale reconfiguration of enhancer-promoter hubs. eLIFE 2016, 5, e18270. [Google Scholar] [CrossRef]
  56. Jiang, S.; Zhou, H.; Liang, J.; Gerdt, C.; Wang, C.; Ke, L.; Schmidt, S.C.S.; Narita, Y.; Ma, Y.; Wang, S.; et al. The Epstein-Barr Virus Regulome in Lymphoblastoid Cells. Cell Host Microbe 2017, 22, 561–573.e4. [Google Scholar] [CrossRef] [Green Version]
  57. Whyte, W.A.; Orlando, D.A.; Hnisz, D.; Abraham, B.J.; Lin, C.Y.; Kagey, M.H.; Rahl, P.B.; Lee, T.I.; Young, R.A. Master Transcription Factors and Mediator Establish Super-Enhancers at Key Cell Identity Genes. Cell 2013, 153, 307–319. [Google Scholar] [CrossRef] [Green Version]
  58. Gunnell, A.; Webb, H.M.; Wood, C.D.; McClellan, M.J.; Wichaidit, B.; Kempkes, B.; Jenner, R.G.; Osborne, C.; Farrell, P.J.; West, M.J. RUNX super-enhancer control through the Notch pathway by Epstein-Barr virus transcription factors regulates B cell growth. Nucleic Acids Res. 2016, 44, 4636–4650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Boija, A.; Klein, I.A.; Sabari, B.R.; Dall’Agnese, A.; Coffey, E.L.; Zamudio, A.V.; Li, C.H.; Shrinivas, K.; Manteiga, J.C.; Hannett, N.M.; et al. Transcription Factors Activate Genes through the Phase-Separation Capacity of Their Activation Domains. Cell 2018, 175, 1842–1855.e16. [Google Scholar] [CrossRef] [Green Version]
  60. Peng, Q.; Wang, L.; Qin, Z.; Wang, J.; Zheng, X.; Wei, L.; Zhang, X.; Zhang, X.; Liu, C.; Li, Z.; et al. Phase Separation of Epstein-Barr Virus EBNA2 and Its Coactivator EBNALP Controls Gene Expression. J. Virol. 2020, 94, e01771-19. [Google Scholar] [CrossRef] [PubMed]
  61. McCann, E.M.; Kelly, G.L.; Rickinson, A.B.; Bell, A.I. Genetic analysis of the Epstein-Barr virus-coded leader protein EBNA-LP as a co-activator of EBNA2 function. J. Gen. Virol. 2001, 82, 3067–3079. [Google Scholar] [CrossRef]
  62. Harada, S.; Kieff, E. Epstein-Barr virus nuclear protein LP stimulates EBNA-2 acidic domain-mediated transcriptional activation. J. Virol. 1997, 71, 6611–6618. [Google Scholar] [CrossRef] [Green Version]
  63. Mannick, J.B.; Cohen, J.I.; Birkenbach, M.; Marchini, A.; Kieff, E. The Epstein-Barr virus nuclear protein encoded by the leader of the EBNA RNAs is important in B-lymphocyte transformation. J. Virol. 1991, 65, 6826–6837. [Google Scholar] [CrossRef] [Green Version]
  64. Szymula, A.; Palermo, R.D.; Bayoumy, A.; Groves, I.J.; Ba Abdullah, M.; Holder, B.; White, R.E. Epstein-Barr virus nuclear antigen EBNA-LP is essential for transforming naive B cells, and facilitates recruitment of transcription factors to the viral genome. PLoS Pathog. 2018, 14, e1006890. [Google Scholar] [CrossRef] [Green Version]
  65. Sinclair, A.J.; Palmero, I.; Peters, G.; Farrell, P.J. EBNA-2 and EBNA-LP cooperate to cause G0 to G1 transition during immortalization of resting human B lymphocytes by Epstein-Barr virus. EMBO J. 1994, 13, 3321–3328. [Google Scholar] [CrossRef]
  66. Sinclair, A.J.; Palmero, I.; Holder, A.; Peters, G.; Farrell, P.J. Expression of cyclin D2 in Epstein-Barr virus-positive Burkitt’s lymphoma cell lines is related to methylation status of the gene. J. Virol. 1995, 69, 1292–1295. [Google Scholar] [CrossRef] [Green Version]
  67. Nitsche, F.; Bell, A.; Rickinson, A. Epstein-Barr virus leader protein enhances EBNA-2-mediated transactivation of latent membrane protein 1 expression: A role for the W1W2 repeat domain. J. Virol. 1997, 71, 6619–6628. [Google Scholar] [CrossRef] [Green Version]
  68. Portal, D.; Rosendorff, A.; Kieff, E. Epstein-Barr nuclear antigen leader protein coactivates transcription through interaction with histone deacetylase 4. Proc. Natl. Acad. Sci. USA 2006, 103, 19278–19283. [Google Scholar] [CrossRef] [Green Version]
  69. Ling, P.D.; Peng, R.S.; Nakajima, A.; Yu, J.H.; Tan, J.; Moses, S.M.; Yang, W.H.; Zhao, B.; Kieff, E.; Bloch, K.D.; et al. Mediation of Epstein-Barr virus EBNA-LP transcriptional coactivation by Sp100. EMBO J. 2005, 24, 3565–3575. [Google Scholar] [CrossRef] [Green Version]
  70. Wang, C.; Zhou, H.; Xue, Y.; Liang, J.; Narita, Y.; Gerdt, C.; Zheng, A.Y.; Jiang, R.; Trudeau, S.; Peng, C.W.; et al. Epstein-Barr Virus Nuclear Antigen Leader Protein Coactivates EP300. J. Virol. 2018, 92, e02155-17. [Google Scholar] [CrossRef] [Green Version]
  71. Peng, R.; Tan, J.; Ling, P.D. Conserved regions in the Epstein-Barr virus leader protein define distinct domains required for nuclear localization and transcriptional cooperation with EBNA2. J. Virol. 2000, 74, 9953–9963. [Google Scholar] [CrossRef] [Green Version]
  72. Han, I.; Xue, Y.; Harada, S.; Orstavik, S.; Skalhegg, B.; Kieff, E. Protein kinase A associates with HA95 and affects transcriptional coactivation by Epstein-Barr virus nuclear proteins. Mol. Cell. Biol. 2002, 22, 2136–2146. [Google Scholar] [CrossRef] [Green Version]
  73. Kawaguchi, Y.; Nakajima, K.; Igarashi, M.; Morita, T.; Tanaka, M.; Suzuki, M.; Yokoyama, A.; Matsuda, G.; Kato, K.; Kanamori, M.; et al. Interaction of Epstein-Barr virus nuclear antigen leader protein (EBNA-LP) with HS1-associated protein X-1: Implication of cytoplasmic function of EBNA-LP. J. Virol. 2000, 74, 10104–10111. [Google Scholar] [CrossRef] [Green Version]
  74. Peng, C.W.; Zhao, B.; Chen, H.C.; Chou, M.L.; Lai, C.Y.; Lin, S.Z.; Hsu, H.Y.; Kieff, E. Hsp72 up-regulates Epstein-Barr virus EBNALP coactivation with EBNA2. Blood 2007, 109, 5447–5454. [Google Scholar] [CrossRef]
  75. Portal, D.; Zhou, H.; Zhao, B.; Kharchenko, P.V.; Lowry, E.; Wong, L.; Quackenbush, J.; Holloway, D.; Jiang, S.; Lu, Y.; et al. Epstein-Barr virus nuclear antigen leader protein localizes to promoters and enhancers with cell transcription factors and EBNA2. Proc. Natl. Acad. Sci. USA 2013, 110, 18537–18542. [Google Scholar] [CrossRef] [Green Version]
  76. Huber, F.M.; Greenblatt, S.M.; Davenport, A.M.; Martinez, C.; Xu, Y.; Vu, L.P.; Nimer, S.D.; Hoelz, A. Histone-binding of DPF2 mediates its repressive role in myeloid differentiation. Proc. Natl. Acad. Sci. USA 2017, 114, 6016–6021. [Google Scholar] [CrossRef] [Green Version]
  77. Weintraub, A.S.; Li, C.H.; Zamudio, A.V.; Sigova, A.A.; Hannett, N.M.; Day, D.S.; Abraham, B.J.; Cohen, M.A.; Nabet, B.; Buckley, D.L.; et al. YY1 Is a Structural Regulator of Enhancer-Promoter Loops. Cell 2017, 171, 1573–1588.e28. [Google Scholar] [CrossRef] [Green Version]
  78. Dalbies-Tran, R.; Stigger-Rosser, E.; Dotson, T.; Sample, C.E. Amino acids of Epstein-Barr virus nuclear antigen 3A essential for repression of Jkappa-mediated transcription and their evolutionary conservation. J. Virol. 2001, 75, 90–99. [Google Scholar] [CrossRef] [Green Version]
  79. Zhao, B.; Marshall, D.R.; Sample, C.E. A conserved domain of the Epstein-Barr virus nuclear antigens 3A and 3C binds to a discrete domain of Jkappa. J. Virol. 1996, 70, 4228–4236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Waltzer, L.; Logeat, F.; Brou, C.; Israel, A.; Sergeant, A.; Manet, E. The human J kappa recombination signal sequence binding protein (RBP-J kappa) targets the Epstein-Barr virus EBNA2 protein to its DNA responsive elements. EMBO J. 1994, 13, 5633–5638. [Google Scholar]
  81. Tomkinson, B.; Kieff, E. Second-site homologous recombination in Epstein-Barr virus: Insertion of type 1 EBNA 3 genes in place of type 2 has no effect on in vitro infection. J. Virol. 1992, 66, 780–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Tomkinson, B.; Robertson, E.; Kieff, E. Epstein-Barr virus nuclear proteins EBNA-3A and EBNA-3C are essential for B-lymphocyte growth transformation. J. Virol. 1993, 67, 2014–2025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Maruo, S.; Johannsen, E.; Illanes, D.; Cooper, A.; Kieff, E. Epstein-Barr Virus nuclear protein EBNA3A is critical for maintaining lymphoblastoid cell line growth. J. Virol. 2003, 77, 10437–10447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Maruo, S.; Wu, Y.; Ishikawa, S.; Kanda, T.; Iwakiri, D.; Takada, K. Epstein-Barr virus nuclear protein EBNA3C is required for cell cycle progression and growth maintenance of lymphoblastoid cells. Proc. Natl. Acad. Sci. USA 2006, 103, 19500–19505. [Google Scholar] [CrossRef] [Green Version]
  85. Maruo, S.; Johannsen, E.; Illanes, D.; Cooper, A.; Zhao, B.; Kieff, E. Epstein-Barr virus nuclear protein 3A domains essential for growth of lymphoblasts: Transcriptional regulation through RBP-Jkappa/CBF1 is critical. J. Virol. 2005, 79, 10171–10179. [Google Scholar] [CrossRef] [Green Version]
  86. Maruo, S.; Wu, Y.; Ito, T.; Kanda, T.; Kieff, E.D.; Takada, K. Epstein-Barr virus nuclear protein EBNA3C residues critical for maintaining lymphoblastoid cell growth. Proc. Natl. Acad. Sci. USA 2009, 106, 4419–4424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lee, S.; Sakakibara, S.; Maruo, S.; Zhao, B.; Calderwood, M.A.; Holthaus, A.M.; Lai, C.Y.; Takada, K.; Kieff, E.; Johannsen, E. Epstein-Barr virus nuclear protein 3C domains necessary for lymphoblastoid cell growth: Interaction with RBP-Jkappa regulates TCL1. J. Virol. 2009, 83, 12368–12377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Robertson, E.S.; Lin, J.; Kieff, E. The amino-terminal domains of Epstein-Barr virus nuclear proteins 3A, 3B, and 3C interact with RBPJ(kappa). J. Virol. 1996, 70, 3068–3074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Waltzer, L.; Perricaudet, M.; Sergeant, A.; Manet, E. Epstein-Barr virus EBNA3A and EBNA3C proteins both repress RBP-J kappa-EBNA2-activated transcription by inhibiting the binding of RBP-J kappa to DNA. J. Virol. 1996, 70, 5909–5915. [Google Scholar] [CrossRef] [Green Version]
  90. Cooper, A.; Johannsen, E.; Maruo, S.; Cahir-McFarland, E.; Illanes, D.; Davidson, D.; Kieff, E. EBNA3A association with RBP-Jkappa down-regulates c-myc and Epstein-Barr virus-transformed lymphoblast growth. J. Virol. 2003, 77, 999–1010. [Google Scholar] [CrossRef] [Green Version]
  91. Cludts, I.; Farrell, P.J. Multiple functions within the Epstein-Barr virus EBNA-3A protein. J. Virol. 1998, 72, 1862–1869. [Google Scholar] [CrossRef] [Green Version]
  92. Hickabottom, M.; Parker, G.A.; Freemont, P.; Crook, T.; Allday, M.J. Two nonconsensus sites in the Epstein-Barr virus oncoprotein EBNA3A cooperate to bind the co-repressor carboxyl-terminal-binding protein (CtBP). J. Biol. Chem. 2002, 277, 47197–47204. [Google Scholar] [CrossRef] [Green Version]
  93. Ohashi, M.; Holthaus, A.M.; Calderwood, M.A.; Lai, C.Y.; Krastins, B.; Sarracino, D.; Johannsen, E. The EBNA3 family of Epstein-Barr virus nuclear proteins associates with the USP46/USP12 deubiquitination complexes to regulate lymphoblastoid cell line growth. PLoS Pathog. 2015, 11, e1004822. [Google Scholar] [CrossRef] [PubMed]
  94. Hertle, M.L.; Popp, C.; Petermann, S.; Maier, S.; Kremmer, E.; Lang, R.; Mages, J.; Kempkes, B. Differential gene expression patterns of EBV infected EBNA-3A positive and negative human B lymphocytes. PLoS Pathog. 2009, 5, e1000506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Anderton, E.; Yee, J.; Smith, P.; Crook, T.; White, R.E.; Allday, M.J. Two Epstein-Barr virus (EBV) oncoproteins cooperate to repress expression of the proapoptotic tumour-suppressor Bim: Clues to the pathogenesis of Burkitt’s lymphoma. Oncogene 2008, 27, 421–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Harth-Hertle, M.L.; Scholz, B.A.; Erhard, F.; Glaser, L.V.; Dolken, L.; Zimmer, R.; Kempkes, B. Inactivation of intergenic enhancers by EBNA3A initiates and maintains polycomb signatures across a chromatin domain encoding CXCL10 and CXCL9. PLoS Pathog. 2013, 9, e1003638. [Google Scholar] [CrossRef] [PubMed]
  97. Tursiella, M.L.; Bowman, E.R.; Wanzeck, K.C.; Throm, R.E.; Liao, J.; Zhu, J.; Sample, C.E. Epstein-Barr Virus Nuclear Antigen 3A Promotes Cellular Proliferation by Repression of the Cyclin-Dependent Kinase Inhibitor p21WAF1/CIP1. PLoS Pathog. 2014, 10, e1004415. [Google Scholar] [CrossRef] [Green Version]
  98. Bazot, Q.; Deschamps, T.; Tafforeau, L.; Siouda, M.; Leblanc, P.; Harth-Hertle, M.L.; Rabourdin-Combe, C.; Lotteau, V.; Kempkes, B.; Tommasino, M.; et al. Epstein-Barr virus nuclear antigen 3A protein regulates CDKN2B transcription via interaction with MIZ-1. Nucleic Acids Res. 2014, 42, 9700–9716. [Google Scholar] [CrossRef]
  99. Bazot, Q.; Paschos, K.; Skalska, L.; Kalchschmidt, J.S.; Parker, G.A.; Allday, M.J. Epstein-Barr Virus Proteins EBNA3A and EBNA3C Together Induce Expression of the Oncogenic MicroRNA Cluster miR-221/miR-222 and Ablate Expression of Its Target p57KIP2. PLoS Pathog. 2015, 11, e1005031. [Google Scholar] [CrossRef] [Green Version]
  100. Maruo, S.; Zhao, B.; Johannsen, E.; Kieff, E.; Zou, J.; Takada, K. Epstein-Barr virus nuclear antigens 3C and 3A maintain lymphoblastoid cell growth by repressing p16INK4A and p14ARF expression. Proc. Natl. Acad. Sci. USA 2011, 108, 1919–1924. [Google Scholar] [CrossRef] [Green Version]
  101. Styles, C.T.; Bazot, Q.; Parker, G.A.; White, R.E.; Paschos, K.; Allday, M.J. EBV epigenetically suppresses the B cell-to-plasma cell differentiation pathway while establishing long-term latency. PLoS Biol. 2017, 15, e2001992. [Google Scholar] [CrossRef] [Green Version]
  102. Paschos, K.; Bazot, Q.; Lees, J.; Farrell, P.J.; Allday, M.J. Requirement for PRC1 subunit BMI1 in host gene activation by Epstein-Barr virus protein EBNA3C. Nucleic Acids Res. 2019, 47, 2807–2821. [Google Scholar] [CrossRef]
  103. Paschos, K.; Smith, P.; Anderton, E.; Middeldorp, J.M.; White, R.E.; Allday, M.J. Epstein-barr virus latency in B cells leads to epigenetic repression and CpG methylation of the tumour suppressor gene Bim. PLoS Pathog. 2009, 5, e1000492. [Google Scholar] [CrossRef] [Green Version]
  104. Skalska, L.; White, R.E.; Franz, M.; Ruhmann, M.; Allday, M.J. Epigenetic repression of p16(INK4A) by latent Epstein-Barr virus requires the interaction of EBNA3A and EBNA3C with CtBP. PLoS Pathog. 2010, 6, e1000951. [Google Scholar] [CrossRef] [Green Version]
  105. Price, A.M.; Dai, J.; Bazot, Q.; Patel, L.; Nikitin, P.A.; Djavadian, R.; Winter, P.S.; Salinas, C.A.; Barry, A.P.; Wood, K.C.; et al. Epstein-Barr virus ensures B cell survival by uniquely modulating apoptosis at early and late times after infection. Elife 2017, 6, e22509. [Google Scholar] [CrossRef]
  106. Sommermann, T.; Yasuda, T.; Ronen, J.; Wirtz, T.; Weber, T.; Sack, U.; Caeser, R.; Zhang, J.; Li, X.; Chu, V.T.; et al. Functional interplay of Epstein-Barr virus oncoproteins in a mouse model of B cell lymphomagenesis. Proc. Natl. Acad. Sci. USA 2020, 117, 14421–14432. [Google Scholar] [CrossRef]
  107. Schmidt, S.C.; Jiang, S.; Zhou, H.; Willox, B.; Holthaus, A.M.; Kharchenko, P.V.; Johannsen, E.C.; Kieff, E.; Zhao, B. Epstein-Barr virus nuclear antigen 3A partially coincides with EBNA3C genome-wide and is tethered to DNA through BATF complexes. Proc. Natl. Acad. Sci. USA 2015, 112, 554–559. [Google Scholar] [CrossRef] [Green Version]
  108. Wang, A.; Welch, R.; Zhao, B.; Ta, T.; Keles, S.; Johannsen, E. Epstein-Barr Virus Nuclear Antigen 3 (EBNA3) Proteins Regulate EBNA2 Binding to Distinct RBPJ Genomic Sites. J. Virol. 2015, 90, 2906–2919. [Google Scholar] [CrossRef] [Green Version]
  109. McClellan, M.J.; Khasnis, S.; Wood, C.D.; Palermo, R.D.; Schlick, S.N.; Kanhere, A.S.; Jenner, R.G.; West, M.J. Downregulation of integrin receptor-signaling genes by Epstein-Barr virus EBNA 3C via promoter-proximal and -distal binding elements. J. Virol. 2012, 86, 5165–5178. [Google Scholar] [CrossRef] [Green Version]
  110. Skalska, L.; White, R.E.; Parker, G.A.; Turro, E.; Sinclair, A.J.; Paschos, K.; Allday, M.J. Induction of p16(INK4a) is the major barrier to proliferation when Epstein-Barr virus (EBV) transforms primary B cells into lymphoblastoid cell lines. PLoS Pathog. 2013, 9, e1003187. [Google Scholar] [CrossRef]
  111. Ohashi, M.; Hayes, M.; McChesney, K.; Johannsen, E. Epstein-Barr virus nuclear antigen 3C (EBNA3C) interacts with the metabolism sensing C-terminal binding protein (CtBP) repressor to upregulate host genes. PLoS Pathog. 2021, 17, e1009419. [Google Scholar] [CrossRef]
  112. Jiang, S.; Willox, B.; Zhou, H.; Holthaus, A.M.; Wang, A.; Shi, T.T.; Maruo, S.; Kharchenko, P.V.; Johannsen, E.C.; Kieff, E.; et al. Epstein-Barr virus nuclear antigen 3C binds to BATF/IRF4 or SPI1/IRF4 composite sites and recruits Sin3A to repress CDKN2A. Proc. Natl. Acad. Sci. USA 2014, 111, 421–426. [Google Scholar] [CrossRef] [Green Version]
  113. Allday, M.J.; Farrell, P.J. Epstein-Barr virus nuclear antigen EBNA3C/6 expression maintains the level of latent membrane protein 1 in G1-arrested cells. J. Virol. 1994, 68, 3491–3498. [Google Scholar] [CrossRef] [Green Version]
  114. Lin, J.; Johannsen, E.; Robertson, E.; Kieff, E. Epstein-Barr virus nuclear antigen 3C putative repression domain mediates coactivation of the LMP1 promoter with EBNA-2. J. Virol. 2002, 76, 232–242. [Google Scholar] [CrossRef] [Green Version]
  115. Zhao, B.; Sample, C.E. Epstein-barr virus nuclear antigen 3C activates the latent membrane protein 1 promoter in the presence of Epstein-Barr virus nuclear antigen 2 through sequences encompassing an spi-1/Spi-B binding site. J. Virol. 2000, 74, 5151–5160. [Google Scholar] [CrossRef]
  116. Kalchschmidt, J.S.; Bashford-Rogers, R.; Paschos, K.; Gillman, A.C.; Styles, C.T.; Kellam, P.; Allday, M.J. Epstein-Barr virus nuclear protein EBNA3C directly induces expression of AID and somatic mutations in B cells. J. Exp. Med. 2016, 213, 921–928. [Google Scholar] [CrossRef] [Green Version]
  117. Zhao, B.; Mar, J.C.; Maruo, S.; Lee, S.; Gewurz, B.E.; Johannsen, E.; Holton, K.; Rubio, R.; Takada, K.; Quackenbush, J.; et al. Epstein-Barr virus nuclear antigen 3C regulated genes in lymphoblastoid cell lines. Proc. Natl. Acad. Sci. USA 2011, 108, 337–342. [Google Scholar] [CrossRef] [Green Version]
  118. Paschos, K.; Parker, G.A.; Watanatanasup, E.; White, R.E.; Allday, M.J. BIM promoter directly targeted by EBNA3C in polycomb-mediated repression by EBV. Nucleic Acids Res. 2012, 40, 7233–7246. [Google Scholar] [CrossRef] [Green Version]
  119. Kalchschmidt, J.S.; Gillman, A.C.; Paschos, K.; Bazot, Q.; Kempkes, B.; Allday, M.J. EBNA3C Directs Recruitment of RBPJ (CBF1) to Chromatin during the Process of Gene Repression in EBV Infected B Cells. PLoS Pathog. 2016, 12, e1005383. [Google Scholar] [CrossRef] [Green Version]
  120. Abudayyeh, O.O.; Gootenberg, J.S.; Essletzbichler, P.; Han, S.; Joung, J.; Belanto, J.J.; Verdine, V.; Cox, D.B.T.; Kellner, M.J.; Regev, A.; et al. RNA targeting with CRISPR-Cas13. Nature 2017, 550, 280–284. [Google Scholar] [CrossRef] [Green Version]
  121. Banerjee, S.; Lu, J.; Cai, Q.; Saha, A.; Jha, H.C.; Dzeng, R.K.; Robertson, E.S. The EBV Latent Antigen 3C Inhibits Apoptosis through Targeted Regulation of Interferon Regulatory Factors 4 and 8. PLoS Pathog. 2013, 9, e1003314. [Google Scholar] [CrossRef] [Green Version]
  122. Ma, Y.; Walsh, M.J.; Bernhardt, K.; Ashbaugh, C.W.; Trudeau, S.J.; Ashbaugh, I.Y.; Jiang, S.; Jiang, C.; Zhao, B.; Root, D.E.; et al. CRISPR/Cas9 Screens Reveal Epstein-Barr Virus-Transformed B Cell Host Dependency Factors. Cell Host Microbe 2017, 21, 580–591.e7. [Google Scholar] [CrossRef] [Green Version]
  123. Knight, J.S.; Robertson, E.S. Epstein-Barr virus nuclear antigen 3C regulates cyclin A/p27 complexes and enhances cyclin A-dependent kinase activity. J. Virol. 2004, 78, 1981–1991. [Google Scholar] [CrossRef] [Green Version]
  124. Knight, J.S.; Sharma, N.; Kalman, D.E.; Robertson, E.S. A cyclin-binding motif within the amino-terminal homology domain of EBNA3C binds cyclin A and modulates cyclin A-dependent kinase activity in Epstein-Barr virus-infected cells. J. Virol. 2004, 78, 12857–12867. [Google Scholar] [CrossRef] [Green Version]
  125. Shukla, S.K.; Jha, H.C.; El-Naccache, D.W.; Robertson, E.S. An EBV recombinant deleted for residues 130-159 in EBNA3C can deregulate p53/Mdm2 and Cyclin D1/CDK6 which results in apoptosis and reduced cell proliferation. Oncotarget 2016, 7, 18116–18134. [Google Scholar] [CrossRef] [Green Version]
  126. Pei, Y.; Singh, R.K.; Shukla, S.K.; Lang, F.; Zhang, S.; Robertson, E.S. Epstein-Barr Virus Nuclear Antigen 3C Facilitates Cell Proliferation by Regulating Cyclin D2. J. Virol. 2018, 92, e00663-18. [Google Scholar] [CrossRef] [Green Version]
  127. Saha, A.; Halder, S.; Upadhyay, S.K.; Lu, J.; Kumar, P.; Murakami, M.; Cai, Q.; Robertson, E.S. Epstein-Barr virus nuclear antigen 3C facilitates G1-S transition by stabilizing and enhancing the function of cyclin D1. PLoS Pathog. 2011, 7, e1001275. [Google Scholar] [CrossRef]
  128. Knight, J.S.; Sharma, N.; Robertson, E.S. SCFSkp2 complex targeted by Epstein-Barr virus essential nuclear antigen. Mol. Cell. Biol. 2005, 25, 1749–1763. [Google Scholar] [CrossRef] [Green Version]
  129. Knight, J.S.; Sharma, N.; Robertson, E.S. Epstein-Barr virus latent antigen 3C can mediate the degradation of the retinoblastoma protein through an SCF cellular ubiquitin ligase. Proc. Natl. Acad. Sci. USA 2005, 102, 18562–18566. [Google Scholar] [CrossRef] [Green Version]
  130. Saha, A.; Bamidele, A.; Murakami, M.; Robertson, E.S. EBNA3C attenuates the function of p53 through interaction with inhibitor of growth family proteins 4 and 5. J. Virol. 2011, 85, 2079–2088. [Google Scholar] [CrossRef] [Green Version]
  131. Banerjee, S.; Lu, J.; Cai, Q.; Sun, Z.; Jha, H.C.; Robertson, E.S. EBNA3C augments Pim-1 mediated phosphorylation and degradation of p21 to promote B-cell proliferation. PLoS Pathog. 2014, 10, e1004304. [Google Scholar] [CrossRef] [Green Version]
  132. Rosendorff, A.; Illanes, D.; David, G.; Lin, J.; Kieff, E.; Johannsen, E. EBNA3C coactivation with EBNA2 requires a SUMO homology domain. J. Virol. 2004, 78, 367–377. [Google Scholar] [CrossRef] [Green Version]
  133. Knight, J.S.; Lan, K.; Subramanian, C.; Robertson, E.S. Epstein-Barr virus nuclear antigen 3C recruits histone deacetylase activity and associates with the corepressors mSin3A and NCoR in human B-cell lines. J. Virol. 2003, 77, 4261–4272. [Google Scholar] [CrossRef] [Green Version]
  134. Kang, M.S.; Kieff, E. Epstein-Barr virus latent genes. Exp. Mol. Med. 2015, 47, e131. [Google Scholar] [CrossRef] [Green Version]
  135. Westhoff Smith, D.; Sugden, B. Potential cellular functions of Epstein-Barr Nuclear Antigen 1 (EBNA1) of Epstein-Barr Virus. Viruses 2013, 5, 226–240. [Google Scholar] [CrossRef] [Green Version]
  136. De Leo, A.; Calderon, A.; Lieberman, P.M. Control of Viral Latency by Episome Maintenance Proteins. Trends Microbiol. 2020, 28, 150–162. [Google Scholar] [CrossRef]
  137. Altmann, M.; Pich, D.; Ruiss, R.; Wang, J.; Sugden, B.; Hammerschmidt, W. Transcriptional activation by EBV nuclear antigen 1 is essential for the expression of EBV’s transforming genes. Proc. Natl. Acad. Sci. USA 2006, 103, 14188–14193. [Google Scholar] [CrossRef] [Green Version]
  138. Frappier, L. Ebna1. Curr. Top. Microbiol. Immunol. 2015, 391, 3–34. [Google Scholar] [CrossRef]
  139. Frappier, L. EBNA1 and host factors in Epstein-Barr virus latent DNA replication. Curr. Opin. Virol. 2012, 2, 733–739. [Google Scholar] [CrossRef]
  140. Dresang, L.R.; Vereide, D.T.; Sugden, B. Identifying sites bound by Epstein-Barr virus nuclear antigen 1 (EBNA1) in the human genome: Defining a position-weighted matrix to predict sites bound by EBNA1 in viral genomes. J. Virol. 2009, 83, 2930–2940. [Google Scholar] [CrossRef] [Green Version]
  141. Lu, F.; Wikramasinghe, P.; Norseen, J.; Tsai, K.; Wang, P.; Showe, L.; Davuluri, R.V.; Lieberman, P.M. Genome-wide analysis of host-chromosome binding sites for Epstein-Barr Virus Nuclear Antigen 1 (EBNA1). Virol. J. 2010, 7, 262. [Google Scholar] [CrossRef] [Green Version]
  142. Tempera, I.; De Leo, A.; Kossenkov, A.V.; Cesaroni, M.; Song, H.; Dawany, N.; Showe, L.; Lu, F.; Wikramasinghe, P.; Lieberman, P.M. Identification of MEF2B, EBF1, and IL6R as Direct Gene Targets of Epstein-Barr Virus (EBV) Nuclear Antigen 1 Critical for EBV-Infected B-Lymphocyte Survival. J. Virol. 2016, 90, 345–355. [Google Scholar] [CrossRef] [Green Version]
  143. Lamontagne, R.J.; Soldan, S.S.; Su, C.; Wiedmer, A.; Won, K.J.; Lu, F.; Goldman, A.R.; Wickramasinghe, J.; Tang, H.Y.; Speicher, D.W.; et al. A multi-omics approach to Epstein-Barr virus immortalization of B-cells reveals EBNA1 chromatin pioneering activities targeting nucleotide metabolism. PLoS Pathog. 2021, 17, e1009208. [Google Scholar] [CrossRef]
  144. Dheekollu, J.; Wiedmer, A.; Ayyanathan, K.; Deakyne, J.S.; Messick, T.E.; Lieberman, P.M. Cell-cycle-dependent EBNA1-DNA crosslinking promotes replication termination at oriP and viral episome maintenance. Cell 2021, 184, 643–654.e13. [Google Scholar] [CrossRef]
  145. Messick, T.E.; Smith, G.R.; Soldan, S.S.; McDonnell, M.E.; Deakyne, J.S.; Malecka, K.A.; Tolvinski, L.; van den Heuvel, A.P.J.; Gu, B.W.; Cassel, J.A.; et al. Structure-based design of small-molecule inhibitors of EBNA1 DNA binding blocks Epstein-Barr virus latent infection and tumor growth. Sci. Transl. Med. 2019, 11, eaau5612. [Google Scholar] [CrossRef]
  146. Soldan, S.S.; Anderson, E.M.; Frase, D.M.; Zhang, Y.; Caruso, L.B.; Wang, Y.; Deakyne, J.S.; Gewurz, B.E.; Tempera, I.; Lieberman, P.M.; et al. EBNA1 inhibitors have potent and selective antitumor activity in xenograft models of Epstein-Barr virus-associated gastric cancer. Gastric Cancer 2021, 24, 1076–1088. [Google Scholar] [CrossRef]
  147. Lee, E.K.; Kim, S.Y.; Noh, K.W.; Joo, E.H.; Zhao, B.; Kieff, E.; Kang, M.S. Small molecule inhibition of Epstein-Barr virus nuclear antigen-1 DNA binding activity interferes with replication and persistence of the viral genome. Antivir. Res. 2014, 104, 73–83. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, L.W.; Jiang, S.; Gewurz, B.E. Epstein-Barr Virus LMP1-Mediated Oncogenicity. J. Virol. 2017, 91, e01718-16. [Google Scholar] [CrossRef] [Green Version]
  149. Zhao, B.; Barrera, L.A.; Ersing, I.; Willox, B.; Schmidt, S.C.; Greenfeld, H.; Zhou, H.; Mollo, S.B.; Shi, T.T.; Takasaki, K.; et al. The NF-kappaB genomic landscape in lymphoblastoid B cells. Cell Rep. 2014, 8, 1595–1606. [Google Scholar] [CrossRef] [Green Version]
  150. Henley, M.J.; Koehler, A.N. Advances in targeting ‘undruggable’ transcription factors with small molecules. Nat. Rev. Drug Discov. 2021, 20, 669–688. [Google Scholar] [CrossRef]
  151. Liu, J.; Chen, H.; Kaniskan, H.U.; Xie, L.; Chen, X.; Jin, J.; Wei, W. TF-PROTACs Enable Targeted Degradation of Transcription Factors. J. Am. Chem. Soc. 2021, 143, 8902–8910. [Google Scholar] [CrossRef]
  152. Lu, G.; Middleton, R.E.; Sun, H.; Naniong, M.; Ott, C.J.; Mitsiades, C.S.; Wong, K.K.; Bradner, J.E.; Kaelin, W.G., Jr. The myeloma drug lenalidomide promotes the cereblon-dependent destruction of Ikaros proteins. Science 2014, 343, 305–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Bondeson, D.P.; Mares, A.; Smith, I.E.; Ko, E.; Campos, S.; Miah, A.H.; Mulholland, K.E.; Routly, N.; Buckley, D.L.; Gustafson, J.L.; et al. Catalytic in vivo protein knockdown by small-molecule PROTACs. Nat. Chem. Biol. 2015, 11, 611–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Host enhancer/silencer rewiring following EBV transformation of RBLs into LCLs. In RBLs, MYC and CCND2 are expressed at basal level with few enhancer promoter interactions. p16INK4A is expressed at high level with extensive enhancer–promoter interactions and active enhancer mark H3K27ac. Following EBV transformation, EBNA2 induces enhancers 400–500 kb upstream of MYC to loop to MYC TSS to activate MYC expression. EBNALP recruits YY1 to CCND2 locus and promotes enhancer–promoter interaction to activate CCND2 expression. These loci are now marked by H3K27ac. At the p16INK4A locus, much less enhancer–promoter interactions is seen and EBNA3C recruits transcription repressors Sin3A and polycomb repressive complexes to the locus with increased H3K27me3. EBNA3C stabilizes CCND and increases CDK6 kinase activity.
Figure 1. Host enhancer/silencer rewiring following EBV transformation of RBLs into LCLs. In RBLs, MYC and CCND2 are expressed at basal level with few enhancer promoter interactions. p16INK4A is expressed at high level with extensive enhancer–promoter interactions and active enhancer mark H3K27ac. Following EBV transformation, EBNA2 induces enhancers 400–500 kb upstream of MYC to loop to MYC TSS to activate MYC expression. EBNALP recruits YY1 to CCND2 locus and promotes enhancer–promoter interaction to activate CCND2 expression. These loci are now marked by H3K27ac. At the p16INK4A locus, much less enhancer–promoter interactions is seen and EBNA3C recruits transcription repressors Sin3A and polycomb repressive complexes to the locus with increased H3K27me3. EBNA3C stabilizes CCND and increases CDK6 kinase activity.
Viruses 15 00832 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, B. Epstein–Barr Virus B Cell Growth Transformation: The Nuclear Events. Viruses 2023, 15, 832. https://doi.org/10.3390/v15040832

AMA Style

Zhao B. Epstein–Barr Virus B Cell Growth Transformation: The Nuclear Events. Viruses. 2023; 15(4):832. https://doi.org/10.3390/v15040832

Chicago/Turabian Style

Zhao, Bo. 2023. "Epstein–Barr Virus B Cell Growth Transformation: The Nuclear Events" Viruses 15, no. 4: 832. https://doi.org/10.3390/v15040832

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