当前位置: 首页 > 医学版 > 期刊论文 > 基础医学 > 病菌学杂志 > 2005年 > 第15期 > 正文
编号:11202904
Marked Variation in Response of Consensus Binding
     Departments of Molecular Biophysics and Biochemistry

    Pediatrics

    Epidemiology and Public Health, Yale University School of Medicine, New Haven, Connecticut

    Department of Tumour Virology, German Cancer Research Center, Im Neuenheimer Feld 242, Heidelberg, Germany

    ABSTRACT

    The R transactivator (Rta) protein activates Epstein-Barr virus (EBV) lytic-cycle genes by several distinct mechanisms that include direct binding to viral promoters, synergy with BamHI Z EBV replication activator (ZEBRA), and activation of cellular signaling pathways. In the direct and synergistic mechanisms of action, Rta binds to specific DNA sequences that are present in the promoters of responsive genes. It has been difficult to demonstrate the capacity of Rta expressed in mammalian cells to bind DNA in vitro in order to study the relative affinities of Rta binding elements. We discovered that a short C-terminal region of Rta inhibits the ability of Rta to bind DNA in vitro. C-terminally truncated versions of Rta bind DNA efficiently and thus facilitate a comparison of consensus Rta binding elements (CRBEs) found in promoters of five Rta-responsive genes: BMLF1, BHLF1, BMRF1, BaRF1, and BLRF2. All CRBEs in the promoters of the five genes conform to the proposed recognition sequence GNCCN9GGNG, where N is any nucleotide and N9 represents a sequence of nine nucleotides. Nonetheless, CRBEs varied markedly in their abilities to bind Rta in electrophoretic mobility shift assays. Not all CRBEs bound or responded to Rta. Binding affinities of the CRBEs and the capacity to be activated by Rta in reporter assays were strongly correlated. The CRBEs from the BMLF1 and BHLF1 promoters conferred the greatest response. The response of the BMRF1, BaRF1, and BLRF2 CRBEs was less robust. By creation of chimeras, inversions, and point mutations, differences in binding affinities and transcriptional activation levels could be attributed to N9 sequence variation. The length of N9 was also critical for a maximal response. In Raji and BZLF1-knockout cells, the mRNAs of the five Rta-responsive lytic-cycle genes differed dramatically in kinetics of expression, abundance, and synergistic responses to ZEBRA and Rta. Affinities of Rta response elements for Rta are likely to play an important role in temporal regulation and the level of lytic-cycle EBV gene expression.

    INTRODUCTION

    The oncogenic human gammaherpesviruses, Epstein-Barr virus (EBV) and Kaposi's sarcoma-associated herpesvirus (KSHV), exist in two states in infected cells. During latency, the viruses express a limited gene program devoted to strategies for viral genome persistence, alterations in signal transduction, modification of cell cycle controls, inhibition of apoptosis, and modulation of immune recognition. After the switch into lytic replication, the viruses express products that address many of the same biologic functions, but lytic-cycle products are also involved in the essential tasks of DNA replication and packaging of the viral genome, leading to production of progeny viruses. ZEBRA (BamHI Z EBV replication activator) and Rta (R transactivator), the products of two immediate-early genes of EBV, BZLF1 and BRLF1, control the switch between latency and the lytic cycle. ZEBRA was discovered through analysis of regions of defective EBV DNA that were capable of disrupting latency (12, 13). Rta was discovered by showing that a region of EBV DNA carried on a cosmid that by itself did not express early antigen could transactivate the restricted early antigen encoded by BHRF1 (26).

    ZEBRA is sufficient to activate the entire EBV lytic cycle in all cell backgrounds (13, 57, 64). Rta is regularly competent to activate expression of a limited set of early and late viral genes but drives the lytic cycle to completion in only some cell backgrounds (53, 69). The capacity of Rta to initiate the full cascade of lytic-cycle viral gene expression appears to be contingent on its ability to activate ZEBRA and stimulate its own synthesis (53). By using mutant EBV genomes in which the BZLF1 and BRLF1 genes have been disrupted, it has been shown that both immediate-early products are absolutely required for full viral replication and production of mature virions (17). Homologues of Rta are present among all gammaherpesviruses (15, 26, 40, 61, 66). In all these viruses, the Rta proteins play essential roles in stimulating lytic-cycle viral gene expression and DNA replication (53, 61, 66, 67).

    EBV Rta is known to carry out many biologic functions. Rta plays a crucial role in initiation of the lytic cycle by reciprocal activation of ZEBRA and by autostimulation (53). Rta and ZEBRA may be expressed from a bicistronic transcript (6, 42); Rta promotes the translation of ZEBRA from this transcript (6). Rta activates transcription of early- and late-lytic-cycle viral genes. Some late-lytic-cycle genes, including gp350 (BLLF1) and BLRF2, are directly activated by Rta in the absence of DNA replication (17, 55). Rta is essential for viral lytic DNA replication, although, unlike ZEBRA, Rta has not been shown to interact directly with the lytic origin of DNA replication. Rta stimulates cells to enter S phase (63).

    EBV lytic-cycle genes can be subgrouped into different classes on the basis of their responses to ZEBRA and Rta. One group of genes is activated by ZEBRA, acting alone, and another by Rta. A third group of viral genes is activated synergistically by Rta and ZEBRA. Synergistic activation requires that both ZEBRA and Rta bind DNA (52). A fourth group of viral genes may be activated by Rta and repressed by ZEBRA at early times in the lytic cycle before lytic viral DNA synthesis (55). The repressive action of ZEBRA has recently been shown to be dependent on phosphorylation of ZEBRA by casein kinase 2 (16).

    Viral genes are activated by Rta by at least two distinct general modes of action. One group of genes, including those that are stimulated by Rta acting in synergy with ZEBRA, are activated by the binding of Rta to specific DNA sequences, known as Rta response elements (RREs), found in the promoters. Another group of genes that lack RREs in their promoters are activated by indirect mechanisms. At least three examples of the indirect mode of action are known. Rta activates the promoter of the BZLF1 gene (Zp) by activating mitogen-activated protein kinases and phosphoinositol kinases. These kinases in turn lead to activation of the ATF2 protein that binds to the ZII (cyclic AMP response element) site in Zp (1). Rta autostimulates the promoter of BRLF1 (Rp) via an Sp1/Sp3 element (54). Rta activates the EBV DNA polymerase (BALF5) gene via USF and E2F binding sites (39).

    Rta is a 605-amino-acid (aa) protein with no known cellular homologues. The N terminus contains an overlapping DNA binding (aa 1 to 280) and dimerization (aa 1 to 232) domain that does not correspond to any previously described DNA binding motif (43). Rta homodimerizes in the absence of DNA. The transcriptional activation domain is found in the C-terminal region of the protein. An obligatory acidic activation domain (aa 520 to 605) contains positionally conserved hydrophobic residues that are predicted to form alpha helices (43). A weaker accessory activating domain contains two proline-rich subregions (aa 352 to 410 and 450 to 500). The activation domain of Rta contacts the TATA binding protein and TFIID in vitro (41). Rta interacts with CREB binding protein (CBP) at multiple sites on both proteins (62). This interaction, and presumably the histone acetyltransferase activity of CBP, is required for Rta's ability to activate some promoters (e.g., BMLF1) but not others (e.g., BZLF1) (62). Rta also interacts with retinoblastoma protein, causing a release of E2F (63). This interaction is thought to have two functional consequences: activation of certain viral promoters such as BALF5 by E2F and activation of host cells to progress through the S phase of the cell cycle (63). Rta is posttranslationally modified by SUMO-1 at several lysines (5). Modification by SUMO-1 is thought to enhance transcriptional activation of RREs by Rta (5).

    The first RREs were identified in a region of EBV DNA that contains a bidirectional promoter/enhancer that controls expression of the BHRF1 gene (bcl2 homologue) and the BHLF1 gene whose function remains unknown (22). This region overlaps the origin of lytic viral DNA replication (orilyt) (25). Two contiguous RREs were identified in a stretch of about 70 nucleotides (nt). Both RREs were contacted in vitro by a fragment of Rta protein containing the DNA binding domain (aa 1 to 355). On the basis of guanine methylation interference studies, it was surmised that an Rta dimer contacts two GC-rich core sequences separated by 6 or 7 nt (22). A second RRE was identified in the promoter of the BMLF1 gene that encodes a protein that regulates mRNA expression at the posttranscriptional level (4, 34). This single RRE, located at –374 to –391 relative to the transcription start site of BMLF1, was found to confer a response of BMLF1 linked to ? globin (21). This RRE contains GC-rich sequences separated by 8 bp. A consensus Rta binding element (CRBE), 5'GNCCN9GGNG3', where N is any nucleotide and N9 represents a sequence of nine nucleotides, was defined by the cyclic amplification and selection of targets technique (23). There are more than 300 such sites in the EBV genome; about 30% are near TATA elements and could function as enhancers. In the cyclic amplification and selection of targets experiment, Rta was found to bind with comparable affinities to 30 oligonucleotides that fit the consensus and all sites equally mediated Rta-induced transcriptional activation (23).

    In this report we reexamine the nature of the CRBE, using Rta protein that has been expressed in human cells. Until now it has not been possible to study DNA binding by Rta protein expressed in mammalian cells. Previous studies of the protein's ability to bind DNA used Rta that was expressed in Escherichia coli or translated in vitro. Using extracts of human cells, we found that deletions of the C terminus of Rta markedly enhanced its DNA binding capacity. This finding enabled us to compare the relative affinities of the CRBEs found in the promoters of five Rta-responsive genes, BMLF1, BHLF1, BMRF1, BaRF1, and BLRF2. Use of the C-terminally deleted Rta mutant also facilitated a broad strategy of mutagenesis to delineate components of the CRBE that influence binding and response to Rta. Finally, we compared the responses of these five genes to Rta in Raji cells, in which Rta does not activate ZEBRA, and in BZLF1-knockout (BZKO) cells that harbor an EBV genome in which the ZEBRA gene is disrupted.

    MATERIALS AND METHODS

    Cell cultures. HKB5/B5 is an EBV-negative cell line formed by the fusion of HH514-16 cells with 293T cells (8). BJAB cells are derived from an EBV-negative B-cell lymphoma (45). Raji, a human B-cell line derived from a Burkitt's lymphoma, contains an EBV strain that is defective for DNA replication and late gene expression (51). BZKO cells are 293 cells carrying an EBV genome with an inactivated BZLF1 gene (17). Lymphoid cells were grown in RPM1 1640 medium plus fetal calf serum; BZKO cells were cultured in Dulbecco modified Eagle medium plus fetal calf serum in the presence of 100 μg/ml hygromycin.

    Transfection and chemical induction. HKB5/B5 and BZKO cells were transfected using the DMRIE-C reagent (Invitrogen) (8). BJAB and Raji cells were transfected by electroporation (58). Cells were subcultured 2 to 3 days prior to transfection. To induce the lytic cycle, Raji cells were treated with 10 ng/ml of tetradecanoylphorbol-13-acetate (TPA) and 3 mM sodium butyrate.

    Plasmid construction. pRTS, pRTS/Rta, pE4CAT, and pCMV/Z(S186A) have been described previously (53, 55, 61). To make vectors that expressed C-terminal deletion mutants of Rta, various regions of the Rta gene were amplified by PCR and cloned into pRTS at the XbaI and BglII sites. In some constructs, a DNA fragment encoding the VP16 activation domain (aa 413 to 490) was ligated downstream of portions of the Rta gene corresponding to C-terminal truncations. To make reporter constructs, double-stranded annealed oligonucleotides encompassing nucleotides –401 to –365 of the BMLF1 promoter, –766 to –730 and –788 to –752 of the BHLF1 promoter, –182 to –146 of the BMRF1 promoter, –131 to –95 and –232 to –196 of the BaRF1 promoter, and –261 to –225 of the BLRF2 promoter were cloned into pE4CAT digested with HindIII and XbaI. Double-stranded annealed oligonucleotides with different mutations encompassing nucleotides –401 to –365 of the BMLF1 promoter (see Fig. 3, 4, 5, and 6) were cloned into pE4CAT digested with HindIII and XbaI.

    Cell extracts for electrophoretic mobility shift assays (EMSAs) and Western blot analysis. Aliquots of 2 x 106 HKB5/B5 cells were transfected with 4 μg of pRTS, pRTS/Rta, pR550+VP, pR550, or pR575. After 48 h, 107 cells were washed once with phosphate-buffered saline, and cell protein extracts were prepared by the method of Mosser et al. (49). The cell pellets were resuspended in 100 μl of lysis buffer (0.42 M NaCl, 20 mM HEPES [pH 7.5], 25% glycerol, 1.5 mM MgCl2, 0.2 mM EDTA, 1 mM dithiothreitol, 1 mM phenylmethylsulfonyl fluoride, 2 μg aprotinin per ml). Lysates were spun at 90,000 rpm at 4°C for 15 min in a benchtop ultracentrifuge (Beckmann Optima TLX), and the supernatants were aliquoted, flash frozen, and stored at –70°C. Protein concentrations were determined by the Bradford method.

    EMSAs. Annealed double-stranded oligonucleotides were end labeled with 32P by using T4 polynucleotide kinase (Boehringer Mannheim). Binding reaction mixtures contained 15 μg of cell protein in a solution containing 10 mM HEPES (pH 7.5), 50 mM NaCl, 2 mM MgCl2, 2.5 μM ZnSO4, 0.5 mM EDTA, 1 mM dithiothreitol, 15% glycerol, and 0.5 μg poly(dI-dC) in a total volume of 20 μl. After incubation for 5 min at room temperature, 30,000 to 50,000 cpm of labeled oligonucleotide was added in each reaction. For supershift reactions, antisera were added 10 min following the addition of the probe, and incubation at room temperature continued for 10 min. Antibodies to VP16 (sc-7545; Santa Cruz) and YY1 (sc-7341; Santa Cruz) were obtained commercially. Anti-Rta is a polyclonal rabbit antiserum raised against an N-terminal fragment (aa 1 to 320) of Rta (53). For EMSA competition reactions, increasing amounts (50x and 100x) of nonradioactive competitor DNA were added into the initial reaction mix. The reaction mixtures were loaded onto a 4% native polyacrylamide gel in 0.5x Tris-borate-EDTA buffer and electrophoresed at 200 V. Gels were dried on Whatman 3MM paper under a vacuum and exposed to autoradiography film.

    Western blot analysis. Cell extracts were mixed with sodium dodecyl sulfate (SDS) sample buffer and heated to 100°C for 5 min before separation by electrophoresis in a 10% polyacrylamide-SDS gel. Following electrophoresis, the proteins were transferred onto nitrocellulose membranes by electroblotting and blocked in 5% nonfat dry milk overnight at 4°C. The blots were incubated with antiserum at 25°C for 2 h, washed twice for 20 min in 10 mM Tris-HCl (pH 7.5)-200 mM NaCl-5% Tween 20, incubated with 125I-labeled protein A for 1 h, and washed again. The membranes were exposed overnight with intensifying screens to Kodak XAR-5 film at –70°C.

    CAT assays. BJAB cells (1.5 x 107) were transfected with 5 μg of reporter DNA plus 5 μg of expression vector pRTS/Rta or pRTS vector. Chloramphenicol acetyltransferase (CAT) assays were performed 48 h following transfection (58). Activation was calculated as percent acetylation of chloramphenicol in the presence of an activator divided by percent acetylation in the presence of a vector and normalized to the activity of the wild-type BMLF1 RRE cloned into pE4CAT. Results presented are averages of results from at least two separate transfections.

    Northern blot analysis. Raji cells (1.2 x 107) were transfected with 10 μg of pRTS vector or pRTS/Rta plasmid DNA by electroporation or were chemically induced with TPA and butyrate. BZKO cells (5 x 106) were transfected with 6 μg of plasmid DNA by using the DMRIE-C reagent. Total cellular RNA, prepared using an RNeasy mini kit (QIAGEN), was electrophoresed through a 1% agarose formaldehyde gel and transferred onto nylon membranes (Hybond-N+; Amersham Pharmacia Biotech). All probes were labeled by the random-primed method (55). A 531-bp excised EagI fragment from EBV BamHI-M was used to detect the BMRF1 and BaRF1 mRNA, a 1.3-kb EcoRI/BamHI fragment from BamHI-M was used to detect BMLF1 mRNA, and a 268-bp fragment (nt 79,537 to 79,805) amplified by PCR was used to detect BaRF1 mRNA. The BHLF1 probe consisted of a 1.941-kb HincII/SacI fragment from EBV BamHI-H. A 670-bp XbaI/EcoRI fragment from pCMV/BLRF2 was used to detect BLRF2 mRNA. A radiolabeled 370-bp NcoI/PstI fragment of the H1 component of RNase P was used as a probe to control for RNA loading (3). Hybridization was carried out in 6x SSC (1x SSC is 0.15 M NaCl plus 0.015 M sodium citrate)-5x Denhardt's solution-0.5% SDS-100 μg of salmon sperm DNA per ml at 60°C overnight. Membranes were washed in 2x SSC-0.5% SDS once for 20 min and in 0.1x SSC-0.5% SDS three times for 20 min at 60°C.

    RESULTS

    Development of a system to study the DNA binding activity of Rta expressed in human cells. Previous experiments that investigated the interactions between Rta and its binding elements employed Rta that had been purified from E. coli or translated in vitro (21, 23, 52). Since the DNA binding capacity of Rta in mammalian cells might be affected by posttranslational modifications or by protein-protein interactions, we explored the use, in DNA binding assays, of extracts of mammalian cells that expressed Rta. When extracts of HKB5/B5 cells transfected with pRTS/Rta were used in EMSAs, the association between full-length Rta protein and the RRE from the BMLF1 promoter was very weak and detectable only after prolonged exposure of the autoradiograph (Fig. 1A, lane 3, and data not shown). However, two mutant Rta proteins with 55- and 30-aa deletions at the C termini, R550 (aa 1 to 550) and R575 (aa 1 to 575), expressed in the same cells displayed strong DNA binding activity (Fig. 1A, lanes 5 and 6). The specific interaction between the truncated Rta proteins and the BMLF1 promoter RRE was confirmed by supershift with antibody to Rta (aa 1 to 320) (Fig. 1A, lanes 10 and 11). Differences in DNA binding activities of the full-length and truncated Rta proteins were not due to their expression levels since all Rta constructs were expressed equally in transfected cells (Fig. 1B).

    The C-terminal deletion mutations removed a portion of the transcriptional activation domain of the Rta protein. We considered the possibility that a strong transcriptional activation domain, such as that present in the Rta protein, might inhibit DNA binding (50). Since the C-terminal region of Rta is functionally similar to the activation domain of the herpes simplex virus VP16 protein (27), we fused R550 with the VP16 activation domain and tested the DNA binding function of the construct. R550+VP still exhibited markedly enhanced DNA binding activity compared to wild-type Rta (Fig. 1A, lane 4), although R550+VP was reproducibly less active in DNA binding than R550 (Fig. 1A, compare lanes 4 and 5, and data not shown). These data suggested that the C-terminal region of Rta, encompassing aa 576 to 605, inhibits DNA binding in vitro. In further studies to be presented separately, we found that the DNA binding inhibitory region was located in the C-terminal 10 amino acids (L.-W. Chen, P.-J. Chang, and G. Miller, unpublished data). In addition to the Rta/DNA complex, five complexes attributable to cellular proteins binding to the BMLF1 promoter RRE were observed in our assays (Fig. 1A). Four of these complexes were faint, but complex c was abundant. These complexes were equally abundant in extracts of cells expressing Rta and in extracts of cells transfected with empty vector. Complex c was identified as YY1 (see below).

    Heterogeneity of the transactivation function of Rta and DNA binding of R550 on different CRBEs. A proposed Rta binding element, with the consensus sequence GNCCN9GGNG, had been found in the promoters of three EBV lytic-cycle genes that are known targets of Rta: BMLF1, a posttranscriptional activator (4, 34); BMRF1, the DNA polymerase processivity factor (35, 36); and BHLF1, a gene whose function is not known (37). The CRBEs are located at –374 to –390 in the BMLF1 promoter and at –155 to –171 in the BMRF1 promoter. The BHLF1 promoter contains two CRBEs at –740 to –756 (BHLF1-I) and –760 to –776 (BHLF1-II). Two additional EBV lytic-cycle genes that are primarily responsive to Rta have been identified previously: BLRF2, encoding a tegument component (58), and BaRF1, encoding a subunit of ribonucleotide reductase (18, 19). These genes are activated following overexpression of Rta in Raji cells (55). Sequences consistent with the Rta consensus binding site were found in the promoters of these genes as well. However, it had not been determined whether the CRBEs in the BLRF2 and BaRF1 promoters were functional. In the promoter of the BLRF2 gene, the CRBE is located at –234 to –250. Two CRBEs are present in the BaRF1 promoter at –104 to –120 (BaRF1-I) and –205 to –221 (BaRF1-II).

    To analyze and compare their levels of responsiveness to Rta, the seven CRBEs were cloned into a CAT reporter plasmid upstream of the adenovirus E4 minimal promoter. Transcriptional response to Rta was assessed by CAT assays with the EBV-negative B-lymphoma cell line BJAB. The CRBEs differed dramatically in their levels of transcriptional activation by Rta (Fig. 2A). Three CRBEs, one from the BMLF1 promoter and two from the BHLF1 promoter, responded strongly. Three CRBEs, from the BMRF1, BLRF2, and BaRF1-I promoters, responded weakly, at about 20% maximal activity. The response to Rta of the BaRF1-II CRBE did not differ from the low-level activation of the E4-CAT plasmid which lacks a CRBE. These data showed that not all DNA sequences that fit the consensus for an Rta binding element were equivalently responsive to Rta.

    To determine whether the transcriptional response of a CRBE to Rta correlated with its capacity to bind Rta, the interactions between the seven CRBEs and Rta were studied by using EMSA and comparing extracts of cells that had been transfected with empty vector and those that had been transfected with Rta expression plasmids (Fig. 2B, C, and D). The BMLF1, BHLF1-I, and BHLF1-II CRBEs, which mediated high levels of transcription stimulation by Rta, all bound avidly to Rta (aa 1 to 550) expressed in mammalian cells (Fig. 2B) and thus can be defined as strong RREs. The BMRF1, BLRF2, and BaRF1-I CRBEs, which responded 13 to 25% as well as the BMLF1 CRBE in transcriptional activation assays, bound Rta less well than the BMLF1 CRBE in EMSAs and thus represent weak RREs (Fig. 2C and D). We observed no detectable binding of R550 to the BaRF1-II CRBE, which was also inactive in transcriptional activation assays (Fig. 2D). The results of these experiments indicated that there is a tight correlation between in vitro DNA binding activity and in vivo transcriptional activation of CRBEs present in reporter plasmids. Since all the CRBEs fit the proposed consensus sequence, the data additionally suggested that the proposed consensus sequence is not adequate to account for the observed variations in transcriptional activation or DNA binding.

    Importance of the N9 internal sequence of the CRBE. The next group of experiments systematically explored different regions of the CRBE in an effort to account for the heterogeneity of responses to Rta. The proposed CRBE, GNCCN9GGNG, can be considered to consist of three essential components: a 5' core sequence, a 3' core sequence, and 9 nt of internal sequence which has previously been thought to function only as a spacer region between the two GC-rich core sequences (Fig. 2A). The BaRF1-I CRBE, a weak responder, and the BMLF1 CRBE, a strong responder, share identical 5' and 3' core sequences, but they differ in N9 internal and flanking sequences (Fig. 2A). This observation suggested that, in addition to the core sequences, either the N9 or flanking sequences might account for the marked variations in response. To analyze the relative contribution of the different components of the CRBE, we constructed chimeric mutants (Fig. 3A). One set of four hybrid CRBEs contained the flanking and core sequences derived from the BMLF1 RRE with N9 sequences from different CRBEs inserted into this background. An additional two chimeras contained the N9 sequence of the BMLF1 RRE substituted for the N9 sequences of the BaRF1-I and BaRF1-II CRBEs.

    Substitution of N9 derived from the two highly active RREs from the BHLF1 promoter into the background of core and flanking sequences of the BMLF1 CRBE, as in the chimeric mutants ML/HL-I and ML/HL-II, did not significantly reduce transcriptional response or DNA binding (Fig. 3A). Both mutants functioned as strong competitors for binding of R550 to the wild-type BMLF1 RRE (Fig. 3B, lanes 10 to 13). Chimeras containing the N9 sequence from the BaRF1-II CRBE that was deficient in activation and binding of Rta produced intermediate results. The construct ML/aR-II, with the N9 sequence from the BaRF1-II CRBE, was diminished in binding (Fig. 3B, lanes 8 and 9). The construct aR-II/ML, with the N9 sequence from the BMLF1 RRE inserted into the core and flanking sequences of BaRF1-II, was also impaired (Fig. 3C, lanes 12 and 13). These results suggested that the N9 sequence in the BaRF1-II CRBE, as well as core and flanking sequences, could contribute to a reduced response. Both core sequences of the BaRF1-II CRBE, 5'GGCC and 3'GGGG, differ from all the other core sequences. In the 5' core of BaRF1-II, there is a G at position 2, whereas in the other six RREs, there is a T at this position. In the 3' core there is a G at position 3 which is a C or T in the other RREs.

    The most dramatic results that implicated the N9 element in the response to Rta were observed with chimeric mutants that contained components of the CRBE from BaRF1-I. Substitution of the N9 element of the weakly responsive BaRF1-I RRE into the BMLF1 RRE background (as in mutant ML/aR-I) abolished the transcriptional response to Rta and eliminated DNA binding by Rta. ML/aR-I failed to compete for binding by R550 in EMSAs (Fig. 3B, lanes 6 and 7). Conversely, substitution of the N9 sequence from the BMLF1 RRE into the core and flanking sequences of the BaRF-I background, as in mutant aR-I/ML, completely restored the transcriptional response to Rta and the capacity to compete for DNA binding by Rta (Fig. 3A). The aR-I/ML RRE competed for binding by R550 as well as wild-type oligonucleotide competitors (Fig. 3C, lanes 10 and 11). These data revealed the essential importance of the N9 nucleotide sequence in Rta binding and activation.

    Sequence specificity of the N9 element. The N9 sequence of the highly active BMLF1 RRE, CTCTATCAT, differs from the N9 sequence of the weakly responsive BaRF1-I RRE, CATGGAACC, at eight of nine positions. Only C at position 1 is conserved. Constructs with point mutations were generated in order to evaluate which of the remaining positions in N9 might account for the poor response and lack of binding of R550 by the chimera ML/aR-I. Two or three nucleotide point mutations were installed into the BMLF1 RRE in order to mimic the N9 sequence of BaRF1-I. Mutation of nucleotides at positions 4 to 6, as in mutant two (mt2), did not affect transcriptional stimulation or DNA binding by Rta (Fig. 4A and B, lanes 8 and 9). mt3, with alterations at positions 7 to 9, showed about 30% reduction in activation and binding. However, mt1, with changes at positions 2 and 3, reduced the response to Rta by more than 90% and eliminated the capacity of the CRBE to compete with the wild-type BMLF1 RRE for binding of Rta (Fig. 4B, lanes 6 and 7). mt1 behaved similarly to mt4, in which both core elements of the BMLF1 RRE were altered while the N9 and flanking sequences were unchanged.

    The results of the experiments described above suggested that a specific nucleotide sequence in N9 was essential for optimal binding and response to Rta. However, the experiments could not exclude the possibility that the ratio of purines to pyrimidines affected the structure of the N9 element and consequently its response to Rta. When the N9 sequences of two highly active RREs were installed in an inverted and complementary orientation, as in the constructs BMLF1 (IN) and BHLF1-I (IN), stimulation by Rta in a reporter assay and binding by R550 were considerably reduced, although not eliminated (Fig. 5). Both inverted constructs contained the sequence TG at positions 2 and 3 of N9. T is present at position 2 in the wild-type BMLF1 RRE and G at position 3 in the wild-type BHLF1-I RRE. This may explain the retention of some activity by the inverted N9 sequences.

    N9 spacing is optimal. The preceding experiments (Fig. 3 to 5) with chimeras and constructs with CRBE point and inversion mutations strongly suggested that the N9 nucleotide sequence per se played a role in direct recognition by Rta. However, the contribution of the length of the N9 sequence to the Rta response needed to be evaluated. While recognition of the CRBE might depend on sequence, it might also require the proper spacer length between the core elements. Alteration of the CRBE by addition or subtraction of a single nucleotide markedly affected the response to Rta (Fig. 6A). The effect of removal of one nucleotide was more deleterious than the consequences of adding one nucleotide. A mutated BMLF1 RRE with one added nucleotide still retained a weak response and a slight ability to bind R550 (Fig. 6B, lane 7, and C, lane 9). However, removal of a single nucleotide from N9 eliminated both activity and binding. Insertion of 2 or 4 nt into the N9 sequence also abolished the transcriptional response and binding by Rta (Fig. 6A and C, lanes 10 to 13). Of interest, alteration of the N9 spacing, by elimination of position 5 of N9 or by insertion of nucleotides after position 5, maneuvers that drastically affected binding by Rta, did not impinge on binding of the cellular protein(s) which formed the prominent complex c. This result suggested that binding of Rta to the RRE was independent of binding of complex c.

    Identification of complex c as YY1. The RRE present in the BMLF1 promoter contains an inverted consensus YY1 binding site, CGCCATNTT, that overlaps the 3' core element and positions 5 through 9 of the N9 element (Fig. 7). To determine whether YY1 could bind to the BMLF1 RRE, EMSA competitions were conducted using duplex oligonucleotides containing canonical or mutated YY1 recognition sites (Fig. 7). Complex c was eliminated by the wild-type YY1 competitor (Fig. 7B, lanes 3, 4, 9, and 10) but not by the mutated YY1 site (Fig. 7B, lanes 5, 6, 11, and 12). Antibody to YY1 also altered the electrophoretic mobility of complex c (Fig. 7B, lanes 7 and 13).

    The BaRF1-I RRE also reproducibly formed a complex that comigrated with complex c (Fig. 2D, lanes 6 to 8). The BaRF1-I RRE contains a degenerate YY1 site, CCCATGGAA, that overlaps the 5' core and N9 sequence. Complex c formed by the BaRF1-I RRE could be altered in mobility by antibody to YY1 and was eliminated by competition with oligonucleotides containing a wild-type YY1 consensus sequence and not by oligonucleotides with a mutated YY1 site (data not shown). None of the other five CRBEs contained consensus YY1 recognition sites.

    Heterogeneity of expression of EBV lytic-cycle genes containing RREs of different affinities. In Raji cells, Rta activates some lytic-cycle genes but does not detectably cross-stimulate the expression of ZEBRA (55). Therefore, this cell background permits analysis of Rta-responsive genes. We examined Rta-dependent expression of the five lytic-cycle genes in the promoters of which RREs of heterogeneous response had been identified (Fig. 2). After Raji cells were transfected with Rta expression plasmid, the abundance of lytic-cycle mRNAs was examined at intervals from 12 to 48 h after transfection. The level of lytic-cycle mRNA observed following transfection with Rta was compared to that in cells that had been transfected with control vector and in cells that had been treated with TPA and sodium butyrate, a stimulus that activates expression of both Rta and ZEBRA (20, 70).

    Two remarkable findings, illustrated in Fig. 8A, were evident from this experiment. The kinetics of expression of the Rta-responsive genes varied markedly. Near-maximal levels of the BMLF1 and BLRF2 mRNAs were present 12 h after transfection with Rta; maximal expression of the other three genes was delayed to 24 h or later. The expression levels following TPA/sodium butyrate treatment, relative to expression following transfection with Rta, varied among the five genes. The expression of two genes, BaRF1 and BMLF1, was equally high after Rta transfection and after TPA/sodium butyrate treatment. The expression of two genes, BHLF1 and BMRF1, that are known to be synergistic targets of Rta and ZEBRA (55) was greater following chemical induction than following transfection with Rta. Expression of a third class of Rta-responsive genes, represented by BLRF2, was inhibited by TPA/sodium butyrate treatment. Rta-dependent activation of BLRF2 is known to be repressed by ZEBRA (55). These experiments defined three distinct classes of Rta-responsive genes.

    Related experiments were conducted with BZKO cells in an effort to determine whether an Rta response could be observed in the complete absence of ZEBRA and whether the Rta response was synergistic with ZEBRA (Fig. 8B). Transfection of BZKO cells with an Rta expression vector resulted in weak activation of all five lytic-cycle genes (Fig. 8B, lane 2). This component of the experiment showed that low-level transcription of these genes can be activated in the total absence of ZEBRA. Cotransfection with Z(S186A), a mutant ZEBRA protein that is competent to synergize with Rta but by itself manifests no activity, resulted in synergistic activation of BMRF1 and BHLF1 but did not affect the expression level of BaRF1, BMLF1, or BLRF2. The same two of the five Rta-responsive genes that were shown to be synergistic targets in Raji cells, by comparing the response to Rta with the response to TPA/butyrate, were found to be synergistic targets by combined expression of Rta and Z(S186A) in BZKO cells.

    DISCUSSION

    In this report, we provide several novel findings that should advance understanding of the mechanism by which Rta protein enhances EBV lytic-cycle activation. Rta contains amino acids at its C terminus that inhibit the capacity of the protein to bind DNA in vitro (Fig. 1). Elimination of this inhibitory region allowed us to develop the first system to study the DNA binding activity of EBV Rta protein that was expressed in human cells. We found that seven CRBEs from five EBV lytic-cycle promoters, all of which fit the previously defined consensus sequence, GNCCN9GGNG, varied markedly in their capacities to bind to Rta in vitro and to confer activation by Rta in reporter assays in transient transfection assays with human cells (Fig. 2). A significant component of this variation in binding and activation responses was attributable to the N9 sequences, which were previously thought merely to provide a spacer function (Fig. 3 to 6). On the basis of these studies, we propose an optimal sequence for an RRE (Fig. 9). We also identified YY1 as a predominant cellular factor that binds to the BMLF1 RRE (Fig. 7). However, binding of YY1 could not account for variation in responses among the seven CRBEs that were studied. We showed that kinetics of expression, abundance, and synergy with ZEBRA of Rta-induced transcripts of the five lytic-cycle genes were highly variable (Fig. 8), a result that implied that there are further temporal and possibly stability controls imposed on Rta-induced transcripts. These new findings represent previously undescribed components of a puzzle that attempts to account for how a single viral transcription factor selectively regulates expression of different classes of viral genes.

    Regulation of the DNA binding activity of Rta. Deletion of the C-terminal 55 aa of Rta markedly enhanced the capacity of the protein to bind to the RRE from the BMLF1 promoter (Fig. 1) and to RREs from other promoters (Fig. 2). In further studies (L.-W. Chen et al., unpublished), the DNA binding inhibitory region was mapped to the C-terminal 10 aa of Rta. The identification of a DNA binding inhibitory domain in the C terminus of the EBV Rta protein is not only a technical advance that has proved useful in analysis of DNA binding in vitro by Rta, it is also likely to elucidate an important regulatory mechanism. If the default position of Rta is a non-DNA binding form, then the unanswered question is, "How does Rta access DNA in vivo?" One possibility is that the C-terminal DNA binding inhibitory domain can be transiently inactivated, through either protein-protein interaction or modification.

    Since these deletions of the DNA binding inhibitory region eliminated a portion of the transcriptional activation domain of Rta, it was possible that competition between interactions with components of the transcription machinery and interaction of Rta with DNA may account for enhanced binding of C-terminally truncated mutants to DNA. However, C-terminally truncated mutants of Rta to which the powerful activation domain of herpes simplex virus VP16 protein had been appended still bound DNA with higher affinity than wild-type Rta (Fig. 1A, lane 4). Thus, it is unlikely that elimination of the transcriptional activation domain alone accounts for restoration of the DNA binding activity measurable by EMSA.

    Several cellular transcription factors, including Ets-1, p53, androgen receptor, and interferon regulatory factor 3, also contain autoinhibitory domains that block the capacity of the proteins to bind DNA (24, 30, 32, 38, 50). Recently, a DNA binding inhibitory region has been identified between aa 520 and 535 of ORF50 protein, the Rta homologue in KSHV (7). Basic amino acids in this inhibitory region of ORF50 that play an essential role in blocking DNA binding also affect ORF50 protein stability (7). There are no obvious amino acid similarities between the autoinhibitory regions of KSHV ORF50 and EBV Rta. Furthermore, elimination of amino acids that inhibit the ability of EBV Rta to bind DNA did not affect protein stability (Fig. 1B and data not shown).

    At least two mechanisms can be invoked to account for autoinhibition of DNA binding. First is an intramolecular mechanism: the inhibitory region may interact with the DNA binding domain of the protein. In the folded structure of the protein, the inhibitory region may sterically interfere with the interaction of the DNA binding domain with DNA. Second is an intermolecular mechanism: the inhibitory region may interact with another cellular protein that in turn alters conformation or blocks access to DNA. Studies with purified Rta protein will be required to distinguish between these alternatives.

    Since Rta operates by several distinct mechanisms, including direct interaction with viral promoters, synergistic interaction with ZEBRA, and activation of signal transduction cascades, it is likely that the DNA binding activity of Rta protein is tightly regulated. The DNA binding inhibitory region that we have identified likely plays an important role in this regulation.

    Heterogeneity of CRBEs. The seven CRBEs could be divided into three groups, based on their affinities for Rta in vitro and on their capacities to be activated by Rta in reporter assays. There was close correlation between binding affinity for Rta, as assessed by EMSA, and the capacity of the CRBE to confer a transcriptional response to Rta. The RREs from BMLF1 and BHLF1 (I and II) were of high affinity; the BMRF1, BLRF2, and BaRF1-I RREs were of low affinity; and the BaRF1-II CRBE did not function as an RRE (Fig. 2).

    Extensive mutagenesis experiments showed that a significant component of the variation in responses was attributable to the 9 nt which separated the 5' and 3' core elements of the CRBE. This was most convincingly demonstrated by chimeric constructs. For example, ML/aR-I, a chimera which contained the core and flanking sequences from the high-avidity BMLF1 CRBE and the N9 sequence from the low-affinity BaRF1-I CRBE, lost binding activity and could not be transactivated by Rta. Nonetheless, the N9 sequences were not sufficient to confer a response and needed to be accompanied by appropriate core and flanking sequences. For example, the N9 sequences from the BMLF1 RRE could only partially rescue the nonfunctional BaRF1-II CRBE in the chimera aR-II/ML.

    Point mutations within N9 showed that specific nucleotides at positions 2 and 3 and to a lesser extent those at positions 7 to 9 were required for maximal response. Total purine and pyrimidine content within N9 was not sufficient to maintain maximal binding or activation since inversion of the N9 sequences of the highly reactive CRBEs from BMLF1 and BHLF1-I significantly reduced activity (Fig. 5). All of these data indicate that the internal 9 nt of the CRBE serve a sequence-specific function and not only act as a spacer. Nonetheless, the spacer function is also critical, since adding or subtracting a single nucleotide from the N9 sequence of the BMLF1 CRBE reduced its response by more than 80%.

    Guanine methylation studies have suggested that a dimer of Rta contacts both GC-rich core sequences separated by 9 nt. This result suggests that each Rta monomer binds to adjacent major grooves of DNA (23). Several other transcription factors appear to have a mode of interaction with DNA in which core-contacting sequences are separated by spacer nucleotides. For example, the Saccharomyces cerevisiae transcriptional activator, GAL4, binds to conserved CCG triplets at each end of a 17-bp site (44). There is an 11-nt spacer. The binding site for human papillomavirus type 16 E2 protein consists of two palindromic triplets, ACC and GGT, separated by 6-nt spacer sequences (28). The binding site for E. coli cyclic AMP receptor protein has two symmetrically related inverted recognition elements, TGTGA and TCACA, separated by a spacer that is either 6 or 8 nt long (2, 31). The response element of the androgen receptor contains inverted repeats of TGTTCT with a 3-nt spacer (11, 59). In view of our results, it will be important to reexamine the binding sites for these transcription factors for a sequence-specific function of the spacer nucleotides.

    Our results showing that the N9 component of the CRBE provides not only a crucial spacer function but also a sequence-specific function could have one of two interpretations: the crucial bases within N9 may also be bound by monomers of Rta or they may be bound by cellular proteins. Experiments in which one nucleotide was added or subtracted from the N9 sequence (Fig. 6) showed that cellular complex b is exquisitely sensitive to the 9-nt length. This observation favors the idea that the N9 sequences attract a specific cellular protein. An attractive possibility that would merge the two hypotheses is that the N9 sequences bind the HMGB1 protein which has been found in vitro to bend RRE DNA and to enhance the formation of complexes between Rta and DNA (46).

    YY1 is one cellular factor that binds to some CRBEs. Complex c was reproducibly the most abundant of several cellular factors that were found to bind to the BMLF1 CRBE. Several lines of evidence proved that complex c contains YY1 (Fig. 7). A consensus YY1 site is embedded in the CRBE from BMLF1; the YY1 element overlaps nt 8 and 9 of N9 and nt 1 and 2 of the 3' core sequence. Mutations in mt3 and mt4 which destroyed the YY1 consensus sequence abolished formation of complex c. Complex c could be eliminated by competition with a wild-type YY1 site but not by competition with a mutant YY1 oligonucleotide. Complex c could be supershifted by antibody to YY1.

    Our experiments suggested that YY1 and EBV Rta bind independently to the CRBE of the BMLF1 promoter. Elimination of YY1 binding by competition with YY1 oligonucleotide did not abolish the binding of Rta (Fig. 7B, lanes 9 and 10). Mutants that altered the length of N9 maintained binding of YY1 but markedly reduced or eliminated binding of Rta (Fig. 6B, lanes 7 and 11). We never observed a ternary complex in which the CRBE was occupied by YY1 and Rta. While such a large complex could fail to enter the gel, a more likely explanation is that YY1 and Rta do not directly interact on DNA since the binding sites for the two proteins overlap.

    YY1, a zinc finger protein, can repress or activate transcription (48, 56, 60, 65). YY1 has been reported to act as a repressor in the regulation of the EBV BZLF1 and BRLF1 genes (47, 68). Our data do not yet provide conclusive evidence that YY1 either represses or enhances activation of CRBEs by Rta. Both the BMLF1 CRBE and the chimeric aR-I/ML CRBE strongly bind YY1, yet Rta also powerfully activates these constructs (Fig. 3). Thus, YY1 is not likely to function as a repressor in the transactivation assays but may possibly do so in vivo. The two CRBEs from the BHLF1 promoter lack a YY1 consensus binding site (Fig. 2B); these CRBEs bind Rta avidly, and the CRBEs are stimulated strongly by Rta. Thus, binding of YY1 is not required for activation by Rta. Mutant CRBEs that cannot be activated by Rta by virtue of alterations in the length of the N9 sequence nonetheless bind YY1 (Fig. 6). Thus, YY1 binding alone is insufficient to lead to transcriptional activation of an RRE. Conversely, as shown in Fig. 3B, CRBEs with mutant YY1 sites that do not bind YY1 remain maximally responsive to Rta.

    Of the seven CRBEs that we studied, YY1 was found to bind strongly to only one high-affinity RRE, namely, that in BMLF1, and weakly to the low-affinity RRE from BaRF1. Since both high- and low-affinity RREs bound or did not bind YY1, our data do not support an essential universal role for YY1 in regulating Rta-mediated activation of gene expression. Nonetheless, YY1 could function as a repressor of basal activity of the BMLF1 promoter and possibly contribute to temporal regulation of this promoter.

    Behavior of the Rta-responsive genes in a biologic context. We attempted to relate the binding affinities of the RREs in vitro and the transcriptional responses of reporter constructs to the biologic outcome of lytic gene expression following transfection with Rta in two cell backgrounds, Raji and BZKO (Fig. 8). In Raji cells, where Rta does not detectably activate ZEBRA or stimulate its own expression, we compared expression levels of the target genes after transfection with Rta with their expression levels after treatment with TPA and sodium butyrate, which activate both ZEBRA and Rta (55). All five genes whose promoters contained identifiable RREs (Fig. 1) responded to transfection with Rta but with striking differences in levels of transcript abundance, kinetics of expression, and the synergistic roles of additional expression of ZEBRA. It was notable that the most abundant transcripts were observed in Raji cells for the BMLF1 and BHLF1 genes, whose promoters contain the highest-affinity RREs. Transcript abundance was considerably less for the genes with low-affinity RREs. More quantitative studies of transcript abundance are needed before this correlation can be established. However, the kinetics of expression of Rta-induced genes did not correlate with RRE affinity. Near-maximal activity of BMLF1, with a high-affinity RRE, and BLRF2, with a low-affinity RRE, was observed at 12 h, while expression of the BHLF1 and BMRF1 genes was delayed.

    The identification of RREs of different affinities is only the beginning of an analysis that would attempt to understand such complex phenomena as transcript abundance and temporal regulation of the Rta-responsive genes. Our analysis was limited to a reporter assay that measured the response of the RRE fused to a heterologous minimal promoter. Additional positive or negative regulating elements undoubtedly exist in the full-length natural promoters. These promoters are also likely to be regulated by epigenetic mechanisms, including DNA methylation and chromatin modifications. Moreover, our analysis did not examine a number of other important components of mRNA regulation, such as transcriptional initiation rate, elongation rate, control of RNA processing, RNA transport, and RNA stability.

    In related studies, it was demonstrated that all of the five lytic-cycle genes were activated by transfection of BZKO cells with Rta. BZKO cells are 293 cells containing an EBV bacmid in which the BZLF1 gene has been insertionally inactivated (17). In these cells, transcripts of the genes with high-affinity RREs, namely, BHLF1 and BMLF1, were more abundant than mRNAs from genes with low-affinity RREs, BaRF1, BMRF1, and BLRF2, following expression of Rta alone. BMRF1 and BHLF1 could be shown to be synergistic targets on the basis of markedly enhanced expression following coexpression of Rta and Z(S186A) (Fig. 8B). Since the activity of Z(S186A) can be revealed only in the presence of Rta, it is useful in defining synergistic targets. The other three genes did not respond synergistically to Z(S186A). All five genes were activated by expression of ZEBRA in BZKO cells (data not shown) because ZEBRA activates Rta in this background (17). BLRF2 was not repressed by ZEBRA in BZKO cells because in this cell background ZEBRA stimulates cell DNA synthesis and BLRF2 is repressed only at early times (16, 55). Otherwise, studies with the two-cell background produced identical results. BMLF1 and BaRF1 are dominantly controlled by expression of Rta. BHLF1 and BMRF1 are synergistic targets of Rta and ZEBRA. This classification of Rta-responsive genes is similar to those in previous studies using reporter constructs (9, 10, 14, 29, 33, 52) and is supported by experiments in which the effects of transfection of Raji cells with Rta alone and with Rta plus the Z(S186A) mutant were compared (55). In summary, in both cell backgrounds, RRE affinity correlated with transcript abundance, but in neither cell background could RRE affinity predict whether the responding gene was a dominant target of Rta or a synergistic target of Rta and ZEBRA.

    In conclusion, this is the first report to study the DNA binding function of Rta expressed in mammalian cells. The system we describe has permitted us to elucidate differences in affinities of the RREs and to further define the nature of the RREs and has enabled us to begin to delineate interactions among RREs, Rta, and cellular proteins. An ultimate goal will be to understand how the differences in the RREs that we have described influence biologic parameters of the EBV lytic cycle, such as transcript kinetics, abundance, and cell type-specific controls on the target genes of Rta.

    ACKNOWLEDGMENTS

    This work was supported by NIH grants CA16038, CA12055, and CA70036.

    We thank Jill Countryman and Ayman El-Guindy for helpful comments on the manuscript.

    REFERENCES

    Adamson, A. L., D. Darr, E. Holley-Guthrie, R. A. Johnson, A. Mauser, J. Swenson, and S. Kenney. 2000. Epstein-Barr virus immediate-early proteins BZLF1 and BRLF1 activate the ATF2 transcription factor by increasing the levels of phosphorylated p38 and c-Jun N-terminal kinases. J. Virol. 74:1224-1233.

    Barber, A. M., V. B. Zhurkin, and S. Adhya. 1993. CRP-binding sites: evidence for two structural classes with 6-bp and 8-bp spacers. Gene 130:1-8.

    Bartkiewicz, M., H. Gold, and S. Altman. 1989. Identification and characterization of an RNA molecule that copurifies with RNase P activity from HeLa cells. Genes Dev. 3:488-499.

    Buisson, M., E. Manet, M. C. Trescol-Biemont, H. Gruffat, B. Durand, and A. Sergeant. 1989. The Epstein-Barr virus (EBV) early protein EB2 is a posttranscriptional activator expressed under the control of EBV transcription factors EB1 and R. J. Virol. 63:5276-5284.

    Chang, L. K., Y. H. Lee, T. S. Cheng, Y. R. Hong, P. J. Lu, J. J. Wang, W. H. Wang, C. W. Kuo, S. S. Li, and S. T. Liu. 2004. Post-translational modification of Rta of Epstein-Barr virus by SUMO-1. J. Biol. Chem. 279:38803-38812.

    Chang, P. J., Y. S. Chang, and S. T. Liu. 1998. Role of Rta in the translation of bicistronic BZLF1 of Epstein-Barr virus. J. Virol. 72:5128-5136.

    Chang, P. J., and G. Miller. 2004. Autoregulation of DNA binding and protein stability of Kaposi's sarcoma-associated herpesvirus ORF50 protein. J. Virol. 78:10657-10673.

    Chang, P. J., D. Shedd, L. Gradoville, M. S. Cho, L. W. Chen, J. Chang, and G. Miller. 2002. Open reading frame 50 protein of Kaposi's sarcoma-associated herpesvirus directly activates the viral PAN and K12 genes by binding to related response elements. J. Virol. 76:3168-3178.

    Chavrier, P., H. Gruffat, A. Chevallier-Greco, M. Buisson, and A. Sergeant. 1989. The Epstein-Barr virus (EBV) early promoter DR contains a cis-acting element responsive to the EBV transactivator EB1 and an enhancer with constitutive and inducible activities. J. Virol. 63:607-614.

    Chevallier-Greco, A., H. Gruffat, E. Manet, A. Calender, and A. Sergeant. 1989. The Epstein-Barr virus (EBV) DR enhancer contains two functionally different domains: domain A is constitutive and cell specific, domain B is transactivated by the EBV early protein R. J. Virol. 63:615-623.

    Claessens, F., G. Verrijdt, E. Schoenmakers, A. Haelens, B. Peeters, G. Verhoeven, and W. Rombauts. 2001. Selective DNA binding by the androgen receptor as a mechanism for hormone-specific gene regulation. J. Steroid Biochem. Mol. Biol. 76:23-30.

    Countryman, J., H. Jenson, R. Seibl, H. Wolf, and G. Miller. 1987. Polymorphic proteins encoded within BZLF1 of defective and standard Epstein-Barr viruses disrupt latency. J. Virol. 61:3672-3679.

    Countryman, J., and G. Miller. 1985. Activation of expression of latent Epstein-Barr herpesvirus after gene transfer with a small cloned subfragment of heterogeneous viral DNA. Proc. Natl. Acad. Sci. USA 82:4085-4089.

    Cox, M. A., J. Leahy, and J. M. Hardwick. 1990. An enhancer within the divergent promoter of Epstein-Barr virus responds synergistically to the R and Z transactivators. J. Virol. 64:313-321.

    DeWire, S. M., M. A. McVoy, and B. Damania. 2002. Kinetics of expression of rhesus monkey rhadinovirus (RRV) and identification and characterization of a polycistronic transcript encoding the RRV Orf50/Rta, RRV R8, and R8.1 genes. J. Virol. 76:9819-9831.

    El-Guindy, A. S., and G. Miller. 2004. Phosphorylation of Epstein-Barr virus ZEBRA protein at its casein kinase 2 sites mediates its ability to repress activation of a viral lytic cycle late gene by Rta. J. Virol. 78:7634-7644.

    Feederle, R., M. Kost, M. Baumann, A. Janz, E. Drouet, W. Hammerschmidt, and H. J. Delecluse. 2000. The Epstein-Barr virus lytic program is controlled by the co-operative functions of two transactivators. EMBO J. 19:3080-3089.

    Gibson, T., P. Stockwell, M. Ginsburg, and B. Barrell. 1984. Homology between two EBV early genes and HSV ribonucleotide reductase and 38K genes. Nucleic Acids Res. 12:5087-5099.

    Gibson, T. J., B. G. Barrell, and P. J. Farrell. 1986. Coding content and expression of the EBV B95-8 genome in the region from base 62,248 to base 82,920. Virology 152:136-148.

    Gradoville, L., D. Kwa, A. El-Guindy, and G. Miller. 2002. Protein kinase C-independent activation of the Epstein-Barr virus lytic cycle. J. Virol. 76:5612-5626.

    Gruffat, H., N. Duran, M. Buisson, F. Wild, R. Buckland, and A. Sergeant. 1992. Characterization of an R-binding site mediating the R-induced activation of the Epstein-Barr virus BMLF1 promoter. J. Virol. 66:46-52.

    Gruffat, H., E. Manet, A. Rigolet, and A. Sergeant. 1990. The enhancer factor R of Epstein-Barr virus (EBV) is a sequence-specific DNA binding protein. Nucleic Acids Res. 18:6835-6843.

    Gruffat, H., and A. Sergeant. 1994. Characterization of the DNA-binding site repertoire for the Epstein-Barr virus transcription factor R. Nucleic Acids Res. 22:1172-1178.

    Hagman, J., and R. Grosschedl. 1992. An inhibitory carboxyl-terminal domain in Ets-1 and Ets-2 mediates differential binding of ETS family factors to promoter sequences of the mb-1 gene. Proc. Natl. Acad. Sci. USA 89:8889-8893.

    Hammerschmidt, W., and B. Sugden. 1988. Identification and characterization of oriLyt, a lytic origin of DNA replication of Epstein-Barr virus. Cell 55:427-433.

    Hardwick, J. M., P. M. Lieberman, and S. D. Hayward. 1988. A new Epstein-Barr virus transactivator, R, induces expression of a cytoplasmic early antigen. J. Virol. 62:2274-2284.

    Hardwick, J. M., L. Tse, N. Applegren, J. Nicholas, and M. A. Veliuona. 1992. The Epstein-Barr virus R transactivator (Rta) contains a complex, potent activation domain with properties different from those of VP16. J. Virol. 66:5500-5508.

    Hegde, R. S., and E. J. Androphy. 1998. Crystal structure of the E2 DNA-binding domain from human papillomavirus type 16: implications for its DNA binding-site selection mechanism. J. Mol. Biol. 284:1479-1489.

    Holley-Guthrie, E. A., E. B. Quinlivan, E. C. Mar, and S. Kenney. 1990. The Epstein-Barr virus (EBV) BMRF1 promoter for early antigen (EA-D) is regulated by the EBV transactivators, BRLF1 and BZLF1, in a cell-specific manner. J. Virol. 64:3753-3759.

    Hupp, T. R., A. Sparks, and D. P. Lane. 1995. Small peptides activate the latent sequence-specific DNA binding function of p53. Cell 83:237-245.

    Ivanov, V. I., L. E. Minchenkova, B. K. Chernov, P. McPhie, S. Ryu, S. Garges, A. M. Barber, V. B. Zhurkin, and S. Adhya. 1995. CRP-DNA complexes: inducing the A-like form in the binding sites with an extended central spacer. J. Mol. Biol. 245:228-240.

    Jayaraman, J., and C. Prives. 1995. Activation of p53 sequence-specific DNA binding by short single strands of DNA requires the p53 C-terminus. Cell 81:1021-1029.

    Kenney, S., E. Holley-Guthrie, E. C. Mar, and M. Smith. 1989. The Epstein-Barr virus BMLF1 promoter contains an enhancer element that is responsive to the BZLF1 and BRLF1 transactivators. J. Virol. 63:3878-3883.

    Kenney, S., J. Kamine, E. Holley-Guthrie, E. C. Mar, J. C. Lin, D. Markovitz, and J. Pagano. 1989. The Epstein-Barr virus immediate-early gene product, BMLF1, acts in trans by a posttranscriptional mechanism which is reporter gene dependent. J. Virol. 63:3870-3877.

    Kiehl, A., and D. I. Dorsky. 1991. Cooperation of EBV DNA polymerase and EA-D(BMRF1) in vitro and colocalization in nuclei of infected cells. Virology 184:330-340.

    Li, J. S., B. S. Zhou, G. E. Dutschman, S. P. Grill, R. S. Tan, and Y. C. Cheng. 1987. Association of Epstein-Barr virus early antigen diffuse component and virus-specified DNA polymerase activity. J. Virol. 61:2947-2949.

    Lieberman, P. M., J. M. Hardwick, and S. D. Hayward. 1989. Responsiveness of the Epstein-Barr virus NotI repeat promoter to the Z transactivator is mediated in a cell-type-specific manner by two independent signal regions. J. Virol. 63:3040-3050.

    Lin, R., Y. Mamane, and J. Hiscott. 1999. Structural and functional analysis of interferon regulatory factor 3: localization of the transactivation and autoinhibitory domains. Mol. Cell. Biol. 19:2465-2474.

    Liu, C., N. D. Sista, and J. S. Pagano. 1996. Activation of the Epstein-Barr virus DNA polymerase promoter by the BRLF1 immediate-early protein is mediated through USF and E2F. J. Virol. 70:2545-2555.

    Lukac, D. M., R. Renne, J. R. Kirshner, and D. Ganem. 1998. Reactivation of Kaposi's sarcoma-associated herpesvirus infection from latency by expression of the ORF 50 transactivator, a homolog of the EBV R protein. Virology 252:304-312.

    Manet, E., C. Allera, H. Gruffat, I. Mikaelian, A. Rigolet, and A. Sergeant. 1993. The acidic activation domain of the Epstein-Barr virus transcription factor R interacts in vitro with both TBP and TFIIB and is cell-specifically potentiated by a proline-rich region. Gene Expr. 3:49-59.

    Manet, E., H. Gruffat, M. C. Trescol-Biemont, N. Moreno, P. Chambard, J. F. Giot, and A. Sergeant. 1989. Epstein-Barr virus bicistronic mRNAs generated by facultative splicing code for two transcriptional trans-activators. EMBO J. 8:1819-1826.

    Manet, E., A. Rigolet, H. Gruffat, J. F. Giot, and A. Sergeant. 1991. Domains of the Epstein-Barr virus (EBV) transcription factor R required for dimerization, DNA binding and activation. Nucleic Acids Res. 19:2661-2667.

    Marmorstein, R., M. Carey, M. Ptashne, and S. C. Harrison. 1992. DNA recognition by GAL4: structure of a protein-DNA complex. Nature 356:408-414.

    Menezes, J., W. Leibold, G. Klein, and G. Clements. 1975. Establishment and characterization of an Epstein-Barr virus (EBC)-negative lymphoblastoid B cell line (BJA-B) from an exceptional, EBV-genome-negative African Burkitt's lymphoma. Biomedicine 22:276-284.

    Mitsouras, K., B. Wong, C. Arayata, R. C. Johnson, and M. Carey. 2002. The DNA architectural protein HMGB1 displays two distinct modes of action that promote enhanceosome assembly. Mol. Cell. Biol. 22:4390-4401.

    Montalvo, E. A., M. Cottam, S. Hill, and Y. J. Wang. 1995. YY1 binds to and regulates cis-acting negative elements in the Epstein-Barr virus BZLF1 promoter. J. Virol. 69:4158-4165.

    Moriuchi, M., and H. Moriuchi. 2003. YY1 transcription factor down-regulates expression of CCR5, a major coreceptor for HIV-1. J. Biol. Chem. 278:13003-13007.

    Mosser, D. D., N. G. Theodorakis, and R. I. Morimoto. 1988. Coordinate changes in heat shock element-binding activity and HSP70 gene transcription rates in human cells. Mol. Cell. Biol. 8:4736-4744.

    Pufall, M. A., and B. J. Graves. 2002. Autoinhibitory domains: modular effectors of cellular regulation. Annu. Rev. Cell Dev. Biol. 18:421-462.

    Pulvertaft, R. J. V. 1964. Cytology of Burkitt's tumor (African lymphoma). Lancet i:238-240.

    Quinlivan, E. B., E. A. Holley-Guthrie, M. Norris, D. Gutsch, S. L. Bachenheimer, and S. C. Kenney. 1993. Direct BRLF1 binding is required for cooperative BZLF1/BRLF1 activation of the Epstein-Barr virus early promoter, BMRF1. Nucleic Acids Res. 21:1999-2007.

    Ragoczy, T., L. Heston, and G. Miller. 1998. The Epstein-Barr virus Rta protein activates lytic cycle genes and can disrupt latency in B lymphocytes. J. Virol. 72:7978-7984.

    Ragoczy, T., and G. Miller. 2001. Autostimulation of the Epstein-Barr virus BRLF1 promoter is mediated through consensus Sp1 and Sp3 binding sites. J. Virol. 75:5240-5251.

    Ragoczy, T., and G. Miller. 1999. Role of the Epstein-Barr virus RTA protein in activation of distinct classes of viral lytic cycle genes. J. Virol. 73:9858-9866.

    Riquet, F. B., L. Tan, B. K. Choy, M. Osaki, G. Karsenty, T. F. Osborne, P. E. Auron, and M. B. Goldring. 2001. YY1 is a positive regulator of transcription of the Col1a1 gene. J. Biol. Chem. 276:38665-38672.

    Rooney, C., N. Taylor, J. Countryman, H. Jenson, J. Kolman, and G. Miller. 1988. Genome rearrangements activate the Epstein-Barr virus gene whose product disrupts latency. Proc. Natl. Acad. Sci. USA 85:9801-9805.

    Serio, T. R., A. Angeloni, J. L. Kolman, L. Gradoville, R. Sun, D. A. Katz, W. Van Grunsven, J. Middeldorp, and G. Miller. 1996. Two 21-kilodalton components of the Epstein-Barr virus capsid antigen complex and their relationship to ZEBRA-associated protein p21 (ZAP21). J. Virol. 70:8047-8054.

    Shaffer, P. L., A. Jivan, D. E. Dollins, F. Claessens, and D. T. Gewirth. 2004. Structural basis of androgen receptor binding to selective androgen response elements. Proc. Natl. Acad. Sci. USA 101:4758-4763.

    Sui, G., E. B. Affar, Y. Shi, C. Brignone, N. R. Wall, P. Yin, M. Donohoe, M. P. Luke, D. Calvo, and S. R. Grossman. 2004. Yin Yang 1 is a negative regulator of p53. Cell 117:859-872.

    Sun, R., S. F. Lin, L. Gradoville, Y. Yuan, F. Zhu, and G. Miller. 1998. A viral gene that activates lytic cycle expression of Kaposi's sarcoma-associated herpesvirus. Proc. Natl. Acad. Sci. USA 95:10866-10871.

    Swenson, J. J., E. Holley-Guthrie, and S. C. Kenney. 2001. Epstein-Barr virus immediate-early protein BRLF1 interacts with CBP, promoting enhanced BRLF1 transactivation. J. Virol. 75:6228-6234.

    Swenson, J. J., A. E. Mauser, W. K. Kaufmann, and S. C. Kenney. 1999. The Epstein-Barr virus protein BRLF1 activates S phase entry through E2F1 induction. J. Virol. 73:6540-6550.

    Takada, K., N. Shimizu, S. Sakuma, and Y. Ono. 1986. trans activation of the latent Epstein-Barr virus (EBV) genome after transfection of the EBV DNA fragment. J. Virol. 57:1016-1022.

    Tan, D. P., K. Nonaka, G. H. Nuckolls, Y. H. Liu, R. E. Maxson, H. C. Slavkin, and L. Shum. 2002. YY1 activates Msx2 gene independent of bone morphogenetic protein signaling. Nucleic Acids Res. 30:1213-1223.

    Whitehouse, A., I. M. Carr, J. C. Griffiths, and D. M. Meredith. 1997. The herpesvirus saimiri ORF50 gene, encoding a transcriptional activator homologous to the Epstein-Barr virus R protein, is transcribed from two distinct promoters of different temporal phases. J. Virol. 71:2550-2554.

    Wu, T. T., E. J. Usherwood, J. P. Stewart, A. A. Nash, and R. Sun. 2000. Rta of murine gammaherpesvirus 68 reactivates the complete lytic cycle from latency. J. Virol. 74:3659-3667.

    Zalani, S., A. Coppage, E. Holley-Guthrie, and S. Kenney. 1997. The cellular YY1 transcription factor binds a cis-acting, negatively regulating element in the Epstein-Barr virus BRLF1 promoter. J. Virol. 71:3268-3274.

    Zalani, S., E. Holley-Guthrie, and S. Kenney. 1996. Epstein-Barr viral latency is disrupted by the immediate-early BRLF1 protein through a cell-specific mechanism. Proc. Natl. Acad. Sci. USA 93:9194-9199.

    zur Hausen, H., F. J. O'Neill, U. K. Freese, and E. Hecker. 1978. Persisting oncogenic herpesvirus induced by the tumour promotor TPA. Nature 272:373-375.(Lee-Wen Chen, Pey-Jium Ch)