Skip to main content
  • ASM
    • Antimicrobial Agents and Chemotherapy
    • Applied and Environmental Microbiology
    • Clinical Microbiology Reviews
    • Clinical and Vaccine Immunology
    • EcoSal Plus
    • Infection and Immunity
    • Journal of Bacteriology
    • Journal of Clinical Microbiology
    • Journal of Microbiology & Biology Education
    • Journal of Virology
    • mBio
    • Microbiology and Molecular Biology Reviews
    • Microbiology Resource Announcements
    • Microbiology Spectrum
    • Molecular and Cellular Biology
    • mSphere
    • mSystems
  • Log in
  • My alerts
  • My Cart

Main menu

  • Home
  • Articles
    • Current Issue
    • Accepted Manuscripts
    • COVID-19 Special Collection
    • Minireviews
    • JVI Classic Spotlights
    • Archive
  • For Authors
    • Submit a Manuscript
    • Scope
    • Editorial Policy
    • Submission, Review, & Publication Processes
    • Organization and Format
    • Errata, Author Corrections, Retractions
    • Illustrations and Tables
    • Nomenclature
    • Abbreviations and Conventions
    • Publication Fees
    • Ethics Resources and Policies
  • About the Journal
    • About JVI
    • Editor in Chief
    • Editorial Board
    • For Reviewers
    • For the Media
    • For Librarians
    • For Advertisers
    • Alerts
    • RSS
    • FAQ
  • Subscribe
    • Members
    • Institutions
  • ASM
    • Antimicrobial Agents and Chemotherapy
    • Applied and Environmental Microbiology
    • Clinical Microbiology Reviews
    • Clinical and Vaccine Immunology
    • EcoSal Plus
    • Infection and Immunity
    • Journal of Bacteriology
    • Journal of Clinical Microbiology
    • Journal of Microbiology & Biology Education
    • Journal of Virology
    • mBio
    • Microbiology and Molecular Biology Reviews
    • Microbiology Resource Announcements
    • Microbiology Spectrum
    • Molecular and Cellular Biology
    • mSphere
    • mSystems

User menu

  • Log in
  • My alerts
  • My Cart

Search

  • Advanced search
Journal of Virology
publisher-logosite-logo

Advanced Search

  • Home
  • Articles
    • Current Issue
    • Accepted Manuscripts
    • COVID-19 Special Collection
    • Minireviews
    • JVI Classic Spotlights
    • Archive
  • For Authors
    • Submit a Manuscript
    • Scope
    • Editorial Policy
    • Submission, Review, & Publication Processes
    • Organization and Format
    • Errata, Author Corrections, Retractions
    • Illustrations and Tables
    • Nomenclature
    • Abbreviations and Conventions
    • Publication Fees
    • Ethics Resources and Policies
  • About the Journal
    • About JVI
    • Editor in Chief
    • Editorial Board
    • For Reviewers
    • For the Media
    • For Librarians
    • For Advertisers
    • Alerts
    • RSS
    • FAQ
  • Subscribe
    • Members
    • Institutions
Genome Replication and Regulation of Viral Gene Expression

The Herpesvirus Saimiri Replication and Transcription Activator Acts Synergistically with CCAAT Enhancer Binding Protein Alpha To Activate the DNA Polymerase Promoter

Louise Wakenshaw, Matthew S. Walters, Adrian Whitehouse
Louise Wakenshaw
1Institute of Molecular and Cellular Biology, Faculty of Biological Sciences
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Matthew S. Walters
1Institute of Molecular and Cellular Biology, Faculty of Biological Sciences
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Adrian Whitehouse
1Institute of Molecular and Cellular Biology, Faculty of Biological Sciences
2Astbury Centre for Structural Molecular Biology, University of Leeds, Leeds LS2 9JT, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: a.whitehouse@leeds.ac.uk
DOI: 10.1128/JVI.79.21.13548-13560.2005
  • Article
  • Figures & Data
  • Info & Metrics
  • PDF
Loading

ABSTRACT

The open reading frame (ORF) 50 gene product, also known as the replication and transcription activator (Rta), is an immediate-early gene which is well conserved among all gamma-2 herpesviruses and plays a pivotal role in regulating the latent-lytic switch. Herpesvirus saimiri (HVS) ORF 50a functions as a sequence-specific transactivator capable of activating delayed-early (DE) gene expression via binding directly to an ORF 50 response element (RE) within the respective promoter. Analysis of the ORF 50 REs have identified two distinct types within HVS gene promoters. The first comprises a consensus sequence motif, CCN9GG, the second an AT-rich sequence. Here we demonstrate that ORF 50a is capable of transactivating the DE ORF 9 promoter which encodes the DNA polymerase. Deletion analysis of the ORF 9 promoter mapped the ORF 50 RE to a 95-bp region situated 126 bp upstream of the initiation codon. Gel retardation analysis further mapped the RE to a 28-bp fragment, which was able to confer ORF 50 responsiveness on an enhancerless simian virus 40 minimal promoter. Furthermore, sequence analysis identified multiple CCAAT enhancer binding protein alpha (C/EBPα) binding sites within the ORF 9 promoter and specifically two within the close vicinity of the AT-rich ORF 50 RE. Analysis demonstrated that the HVS ORF 50a and C/EBPα proteins associate with the ORF 9 promoter in vivo, interact directly, and synergistically activate the ORF 9 promoter by binding to adjacent binding motifs. Overall, these data suggest a cooperative interaction between HVS ORF 50a and C/EBPα proteins to activate the DNA polymerase promoter during early stages of the lytic replication cycle.

Herpesvirus saimiri (HVS) is the prototype gammaherpesvirus of the Rhadinovirus genus (8, 9). HVS causes a persistent asymptomatic infection in its natural host, the squirrel monkey (Saimiri sciureus); however, upon experimental transfer to other New World primates, HVS infection results in fulminant polyclonal T-cell lymphomas and lymphoproliferative diseases (8, 9). In addition, HVS subgroups A and C possess the ability to immortalize common marmoset T lymphocytes to interleukin-2-independent proliferation (6, 34). Moreover, subgroup C viruses are capable of transforming human, rabbit, and rhesus monkey lymphocytes in vitro (3, 4). HVS can also stably transduce a variety of human cell lines, where they harbor the viral genome as a multiple high-copy-number nonintegrating circular episome without production of virus particles (12, 16, 28-30).

The genome of HVS (strain A11) comprises a unique internal low-G+C DNA segment (L-DNA) of approximately 110 kbp, flanked by a variable number of 1,444-bp high-G+C tandem repetitions (H-DNA) (1). Sequence analysis indicates HVS shares significant homology with other herpesviruses of oncogenic potential, including Epstein-Barr virus (EBV), Kaposi's sarcoma-associated herpesvirus (or human herpesvirus 8) (KSHV), and murine gammaherpesvirus 68 (MHV-68) (1, 22, 27, 35). Their genomes are predominantly colinear with homologous genes in approximately the same locations and orientation. However, conserved gene blocks are separated by genes unique to each respective virus. Like all Rhadinoviruses, the HVS genome encodes a number of cellular homologues whose products may have a role in transformation, immune evasion, and long-term persistence of the viral episome (1, 8).

The open reading frame (ORF) 50 gene product, also known as the replication and transcription activator (Rta), is an immediate-early gene which is well conserved among all gamma-2 herpesviruses (5, 7, 40). The KSHV, MHV-68, and HVS ORF 50 homologues have all been shown to play a pivotal role in regulating the latent-lytic switch in gamma-2 herpesviruses (10, 11, 20, 33, 43). For example, overexpression of the larger spliced ORF 50a gene in a HVS-persistently infected A549 cell line, where the HVS DNA is stably maintained as nonintegrated circular episomes, was sufficient to reactivate the entire lytic replication cycle producing infectious virus particles (10).

The ORF 50 homologues function as sequence-specific transactivators (5). Analysis of the major HVS ORF 50a gene product has shown that it activates transcription following direct interactions with specific response elements (REs) within HVS promoters (23, 36, 37, 41, 42). Recently, we have identified a DNA binding domain within ORF 50a which has a high degree of homology to the DNA binding domain encoded by mammalian high-mobility group A (HMGA) chromosomal proteins with an AT-hook. Deletion and site-directed mutagenesis have demonstrated that this domain, encompassing residues 399 to 412, is required for ORF 50-mediated transactivation and binding to the ORF 50 RE. Moreover, ORF50ΔAT-hook is capable of functioning as a trans-dominant mutant, leading to a reduction in virus replication (36). Once bound, HVS ORF 50a recruits and interacts with the TATA binding protein (17), suggesting that it recruits components of the TFIID complex, allowing transcription initiation by RNA polymerase II.

Further analysis of the ORF 50 REs have identified two distinct types within HVS gene promoters. The first, contained within the major DNA binding protein (ORF 6) and ORF 57 promoters comprises a consensus-specific sequence motif, CCN9GG, (41, 42) Analysis of the HVS genome suggests that at least 10 HVS gene promoters contain similar CCN9GG REs. This type of RE has significant homology to the EBV Rta RE consensus sequence, GNCCN9GGNG (13-15). It has been shown by guanine methylation studies that the CCN9GG motif is essential for EBV Rta binding and suggests that Rta binds to adjacent major grooves of the DNA (13-15). In contrast, HVS ORF 50a is capable of stimulating its own promoter via direct binding to an alternative AT-rich sequence RE within its promoter (37). Sequence analysis of the RE demonstrates that it has no direct homology with the CCN9GG motif containing REs (37). Interestingly, the alternative ORF 50 REs both confer ORF 50 responsiveness to an enhancerless promoter; however, stimulation levels vary, which may suggest a further regulatory step of the lytic temporal cascade (37).

In this report, we aimed to determine whether HVS ORF 50a could transactivate DE genes which do not contain the consensus CCN9GG motif within their gene promoter. One such example is the DE ORF 9 promoter encoding the HVS DNA polymerase. Here we demonstrate that the ORF 9 promoter is responsive to the ORF 50a protein. We show that the ORF 50a protein binds directly to a nonconsensus 28-bp AT-rich RE. Moreover, the ORF 9 promoter contains multiple CCAAT enhancer binding protein alpha (C/EBPα) elements and coexpression of ORF 50a and C/EBPα results in a synergistic effect on the activation of the ORF 9 promoter.

MATERIALS AND METHODS

Plasmid constructs.In order to generate an ORF 9 promoter reporter gene construct, the ORF 9 gene promoter was PCR amplified with primers 5′-CGG GGT ACC GTT AAT GTA GCA AGC GGC GC and 5′-CTA GCT AGC GTC AAG ACA GCA ACT CAG. These oligonucleotides incorporated KpnI and NheI restriction sites, respectively, for convenient cloning of the PCR product. PCR (30 cycles [1 cycle consists of 1 min at 92°C, 1 min at 58°C, and 2 min at 72°C]) was performed with 4 units of Pfu (Stratagene). This fragment was inserted upstream of the luciferase coding region in pGL3-basic (Promega) to derive pORF9-Luc. The 5′ ORF 9 promoter deletion series was constructed by PCR amplification using the forward primer Δ1 (5′-CGG GGT ACC GTT CAC ATT TCT ATT AAT AGG), Δ2 (5′-CGG GGT ACC CCC AGA TGT TGA GAA ATC), Δ3 (5′-CGG GGT ACC GGA ATG CAT AAC ATG CAG), Δ3.5 (5′-CGG GGT ACC GAG CTG CTA GAC CGT C), and Δ4 (5′-CGG GGT ACC GCG TAA GCG ATC TGG TTA); these oligonucleotides incorporated KpnI restriction sites for the convenient cloning of the PCR products. Each fragment was inserted into the luciferase reporter vector, pGL3-basic to derive pORF9Δ1 to pORF9Δ4 (see Fig. 2a). All constructs were confirmed by DNA sequencing (data not shown).

In order to further assess the putative ORF 50 RE within the ORF 9 promoter comprising 20307 to 20335 bp of the published sequence, oligonucleotides encompassing this region, 5′-AAG CAT ATA AGA AAA AGA AGA ACA AAG A and 5′-TCT TTG TTC TTC TTT TTC TTA TAT GCT T, were cloned into pGL3-promoter (Promega), which contains the luciferase reporter gene under the control of a minimal simian virus 40 (SV40) promoter. The oligonucleotides were annealed to form blunt-ended fragments and then cloned directly in either orientation into pGL3-promoter previously digested with SmaI, to derive pGL3-9RE1 and pGL3-9RE2.

In order to generate an ORF 9 promoter construct lacking the putative C/EBPα response element, a deletion construct lacking the sequence between 20286 and 20314 bp was constructed by a PCR-based method. The ORF 9 promoter was PCR amplified in two fragments, thereby removing the C/EBPα binding site, using the following sets of primers: for fragment 1, primer A (5′-CGG GGT ACC GTT AAT GTA GCA AGC GGC GC) and primer B (TTC CCC GGG CCA AGC AAA ATT TGT TTT AAT TG); for fragment 2, primer C (5′ TTC CCC GGG GCA AGA AGC ATA TAA GAA AAA AAG) and primer D (5′-CTA GCT AGC GTC AAG ACA GCA ACT CAG). Primer A and D oligonucleotides incorporated KpnI and NheI and primer B and C incorporated SmaI restriction sites, respectively, for convenient cloning of the PCR product. PCR (30 cycles [1 cycle consists of 1 min at 92°C, 1 min at 50°C, and 2 min at 72°C]) was performed with 4 units of Pfu (Stratagene). These fragments were inserted by triple-way ligation upstream of the luciferase coding region in pGL3-basic (Promega) to derive pORF9ΔC/EBP-Luc.

Plasmids encoding C/EBPα and C/EBPβ were kindly provided by C. Heckman, Stanford University School of Medicine. These constructs contained the respective cDNAs in plasmid pcDNA3.1 (Invitrogen); to generate a myc-tagged version, the coding region was excised from pC/EBPα as a HindIII-XbaI fragment and cloned into pcDNA3.1-myc (Invitrogen), deriving pC/EBPα-myc. Plasmid constructs pEGFP, p50GFP, p50GFPΔAT-hook, pGL3-50RE, and pGL3-6RE have been previously reported (36).

Cell culture, viruses, and transfections.293T cells were maintained in Dulbecco modified Eagle medium (Life Technologies) supplemented with 10% fetal calf serum. Plasmids used in the transfections were prepared using QIAGEN plasmid kits according to the manufacturer's directions. 293T cells were seeded at 5 × 105 cells per 35-mm-diameter petri dish 24 h prior transfection. Transfections were performed using Lipofectamine (Invitrogen) as described by the manufacturer, using 2 μg of the appropriate DNAs. HVS (strain A11) was propagated in owl monkey kidney (OMK) cells which were maintained in Dulbecco modified Eagle medium (Life Technologies) supplemented with 10% fetal calf serum.

Luciferase reporter assay.Transfected 293T cells (1 × 106) were harvested in 200 μl of passive lysis buffer (Promega). Quantitation of relative light units was determined using the dual-luciferase Stop & Glo reagent using the manufacturer's directions (Promega) and a Berthold luminometer (EG& G Berthold) with a dual-injector system. All assays were performed in triplicate, and each experiment was repeated a minimum of three times.

RT-PCR analysis.RNA was extracted from 293T cells 24 h posttransfection. Cells were lysed using Trizol reagent (Invitrogen). Chloroform (0.2 ml) was then added, and the solution was vortex mixed for 20 s and stored at 20°C for 15 min. Samples were centrifuged for 15 min at 4°C, and the aqueous phase containing nucleic acids was precipitated using 0.5 ml of isopropanol. After the pellet was washed with 70% ethanol, it was resuspended in 50 μl of water and stored at −70°C. First-strand cDNA was reverse transcribed using Superscript II reverse transcriptase (Life Technologies) and an oligo(dT) primer. Reverse transcription-PCR (RT-PCR) was then performed using the following sets of forward and reverse primers (indicated by the f or r letter at the end of the primer designation, respectively): ORF50f, 5′-CTG GAA TAG TCT CTA CAA CAT TAG CA; ORF50r, 5′-CCG CTC GAG GTT TGG ATC TAT GTT GCG ACT CTG; LUCf, 5′-GCG CCA TTC TAT CCG CTG; LUCr, 5′-ATC GAA GGA CTC TGG CAC; hGAPDHf, 5′-CCA CCC ATG GCA AAT TCC ATG GCA; and hGAPDHr, 5′-TCT AGA CGG CAG GTC AGG TCC ACC]). PCR (30 cycles [1 cycle consists of 1 min at 92°C, 1 min at 55°C, and 1 min at 72°C]) was performed with 4 units of Taq polymerase (Promega).

Gel retardation analysis.The following sets of fluorescein-labeled oligonucleotides, set 1 (5′-GGA ATG CAT AAC ATG CAG CAA GAA GCA T and 5′-ATG CTT CTT GCT GCA TGT TAT GCA TTC C) and set 2 (5′-AAG CAT ATA AGA AAA AGA AGA ACA AAG A and 5′-TCT TTG TTC TTC TTT TTC TTA TAT GCT T) (Invitrogen) were annealed and then incubated with nuclear extracts of untransfected cells or cells transfected with pORF50a or p50GFPΔAT-hook prepared by the method of Andrews and Faller (2). The binding reactions were performed in 20 μl of 100 mM KCl, 20 mM HEPES (pH 7.3), 1% glycerol, 0.2 mM EDTA, 5 mM MgCl2, 4 mM dithiothreitol, 0.5 mM phenylmethylsulfonyl fluoride with 1 μg of poly(dI-dC) as a nonspecific competitor. The protein-nucleic acid complexes were separated on a 5% polyacrylamide gel, run in 1% Tris-borate-EDTA buffer, and detected using a Typhoon 9410 variable-mode imager (Amersham).

In vitro transcription/translation.Protein expression from ORF 50, C/EBPα, and C/EBPβ was performed by in vitro transcription/translation using the TNT system (Promega) according to the manufacturer's instructions. Transcription was initiated from the bacteriophage T7 promoter situated upstream of the cloned fragments.

Coimmunoprecipitation assays.293T cells were transfected using 2 μg of the appropriate DNAs. After 24 h, cells were harvested and lysed with lysis buffer (0.3 M NaCl, 1% Triton X-100, 50 mM HEPES buffer, pH 8.0) containing protease inhibitors (leupeptin and phenylmethylsulfonyl fluoride). For each immunoprecipitation, 50 μl of the anti-ORF 50 polyclonal antibody (17) was incubated with protein A-Sepharose beads (Pharmacia Biotech) for 16 h at 4°C. The beads were then pelleted and washed four times in phosphate-buffered saline (PBS). Each cell lysate was then added to the beads and incubated for 16 h at 4°C. The beads were then pelleted and washed four times in lysis buffer and resuspended in Laemmli buffer, and precipitated polypeptides were resolved on a 12% sodium dodecyl sulfate (SDS)-polyacrylamide gel.

Immunoblot analysis.Polypeptides were resolved on 12% SDS-polyacrylamide gels and then soaked for 10 min in transfer buffer (25 mM Tris, 192 mM glycine, 20% methanol [vol:vol]). The proteins were transferred to nitrocellulose membranes by electroblotting for 3 h at 250 mA. After transfer, the membranes were soaked in PBS and blocked by preincubation with 2% (wt/vol) nonfat milk powder for 2 h at 37°C. Membranes were incubated with a 1:2,000 dilution of anti-myc (Sigma) antibody or 1:1,000 dilution of anti-green fluorescent protein (GFP) (Clontech) antibody, washed with PBS, and then incubated for 1 h at 37°C with a 1:1,000 dilution of secondary immunoglobulin conjugated with horseradish peroxidase (Dako) in blocking buffer. After five washes with PBS, the nitrocellulose membranes were developed using an ECL kit (Pierce).

Chromatin immunoprecipitation (ChIP) assays.OMK cells seeded at 5 × 105 cells per 35-mm-diameter petri dish, were infected at a multiplicity of infection of 1. After 24 h, the cells were harvested and ChIP assays performed using the ChIP assay kit (Upstate Biotechnology). Chromatin extraction, cross-linking, sonication, immunoprecipitation, agarose bead elution, and protein removal were carried out according to the manufacturer's protocol. DNA recovered from immunoprecipitates with anti-ORF 50, anti-C/EBPα (Santa Cruz), or anti-ORF 57 monoclonal antibody was used as a template for PCR amplifications. The following specific primers for the HVS ORF 9 promoter were used: 5′-CGG GGT ACC CCC AGA TGT TGA GAA ATC and 5′-CTA GCT AGC GTC AAG ACA GCA ACT CAG. These generated a 291-bp region promoter encompassing 20216 and 20507 bp of the published sequence. Primers ORF73f (5′-CGC GGA TCC ATG GAA GCA GGA CCA AGT ACT CCA) and ORF73r (5′-CCG CTC GAG CCT TCT ATA GGC AGG CTT TTG CT-3), specific for the HVS ORF 73 coding region, were used as a negative control to detect a nonpromoter region. The PCR products were analyzed by electrophoresis on a 1.2% agarose gel.

RESULTS

The ORF 50a protein transactivates the DE ORF 9 promoter and requires the AT-hook DNA binding domain.We have previously shown that the HVS ORF 50a protein can transactivate a range of DE HVS promoters which contain a consensus ORF 50 RE, such as the ORF 6 and ORF 57 promoters (41, 42). However, not all DE genes contain such a consensus RE. We have also shown that the ORF 50a protein can stimulate its own promoter via binding to an AT-rich response element (37). Therefore, we aimed to determine whether the ORF 50a protein can activate DE gene promoters which contain AT-rich elements with their promoter elements. One such promoter is the DE ORF 9 DNA polymerase gene promoter.

To determine whether ORF 50a was capable of stimulating the ORF 9 gene promoter, the putative ORF 9 promoter situated between 20034 and 20507 bp of the HVS genome (1) was PCR amplified and cloned upstream of the luciferase reporter gene to derive pORF9-Luc. 293T cells were then transfected with either 1 μg of pEGFP, p50GFP, or p50GFPΔAT-hook (which contains a deletion of the ORF50 DNA binding domain) (36) in the presence of 1 μg of pORF9-Luc. Cells were harvested 30 h posttransfection, and the protein concentration of each sample was calculated using the DC protein assay kit (Bio-Rad). Moreover, to confirm that the transfection efficiency of these experiments was normalized, immunofluorescence was undertaken to ensure that a comparable number of cells was transfected. In addition, 100 μg of cell extracts was analyzed by Western blotting using a primary monoclonal GFP antibody (Clontech) to demonstrate comparable expression between assays (data not shown). An equal amount of cell extracts was then assayed for luciferase activity by standard methods. Results showed that p50GFP was able to stimulate the ORF 9 promoter to high levels, suggesting that the ORF 50a protein can activate a DE promoter which does not contain a consensus ORF 50 RE (Fig. 1). In contrast, stimulation was drastically reduced, by approximately 95%, using p50GFPΔAT-hook, suggesting that ORF 50a stimulates the ORF 9 gene via direct binding to the promoter.

FIG. 1.
  • Open in new tab
  • Download powerpoint
FIG. 1.

ORF 50 transactivation of the ORF 9 DNA polymerase promoter requires the AT-hook DNA binding domain. 293T cells were cotransfected with pORF9-Luc in the presence of pEGFP, p50GFP, or p50GFPΔAT-hook. Cells were harvested 30 h posttransfection, and cell lysates were assayed for luciferase activity by standard methods. The variations between three replicate assays are indicated. All luciferase data are presented as a percentage of luciferase activity compared to the p50GFP level on the pORF9-Luc promoter construct, with the p50GFP level representing 100% activity. WT, wild type.

Mapping the minimal domain within the ORF 9 promoter that is required for ORF 50a stimulation.To map the ORF 50a responsive regions within the ORF 9 promoter, a series of further promoter deletions were generated (Fig. 2a). The promoter fragments were PCR amplified, and resulting products were cloned upstream of the luciferase reporter gene, deriving pORF9Δ1 to pORF9Δ4.

FIG. 2.
  • Open in new tab
  • Download powerpoint
FIG. 2.

Mapping the minimal domain required for ORF 50 stimulation of the ORF 9 promoter. (a) Schematic representation of the ORF 9 promoter deletion series cloned upstream of the luciferase reporter gene. (b) Effect of the ORF 50 wild-type protein on the stimulation of the ORF 9 promoter deletion series. 293T cells were cotransfected with each ORF 9 promoter-luciferase construct in the presence of pEGFP or p50GFP. Cells were harvested 30 h posttransfection, and cell lysates were assayed for luciferase activity by standard methods. The variations between three replicate assays (each performed twice) are indicated. The increase in stimulation is shown as a value above each bar. (c) Schematic representation of more-refined ORF 9 promoter deletion cloned upstream of the luciferase reporter gene. (d) Effect of the ORF 50a protein on the stimulation of the refined ORF 9 promoter deletion series. 293T cells were cotransfected with pORF9-Luc or pORF9Δ3 to pORF9Δ4 in the presence of pEGFP or p50GFP. Cells were harvested 30 h posttransfection, and cell lysates were assayed for luciferase activity by standard methods. The variations between three replicate assays (each performed twice) are indicated. All luciferase data are presented as a percentage of luciferase activity compared to the p50GFP level on the pORF9-Luc promoter construct, with the p50GFP level representing 100% activity. The increase in stimulation is shown as a value above each bar.

293T cells were transfected with either 1 μg of pEGFP or p50GFP in the presence of 1 μg of pORF9Δ1 to pORF9Δ4. Cells were assayed for luciferase activity, and the data were normalized as described above. p50GFP was capable of stimulating pORF9Δ1 to pORF9Δ3 to levels comparable to that with the full-length promoter; however, stimulation drastically decreased between Δ3 and Δ4. The luciferase activity of pORF9Δ4 is only 15% of the activity of pORF9Δ3 (Fig. 2b). These data suggest that the region required by ORF 50a is within the 100 bp between Δ3 and Δ4, comprising 20286 to 20381 bp of the published sequence (1).

To finely map the ORF 50a RE within the ORF 9 promoter, a further deletion within the mapped region was generated (Fig. 2c). The promoter fragment was PCR amplified and subsequently cloned upstream of the luciferase reporter gene as previously described, deriving pORF9Δ3.5. 293T cells were then transfected with 1 μg of p50GFP in the presence of 1 μg of pORF9Δ3, pORF9Δ3.5, or pORF9Δ4 and assayed for luciferase activity as described previously (Fig. 2d). The results demonstrated that p50GFP was capable of stimulating pORF9Δ3 to levels similar to those previously reported. However, deletion of the region between Δ3 and Δ3.5 resulted in the level of ORF 50a stimulation returning to a level comparable to that with Δ4, approximately 15% of wild-type levels. This, therefore, suggests that the element(s) responsible for ORF 50a stimulation lies between Δ3 and Δ3.5, comprising 20286 to 20335 bp of the published sequence within the ORF 9 promoter.

The ORF 50a protein binds directly to an AT-rich sequence within the ORF 9 promoter.To assess whether the ORF 50a protein binds directly to this region encompassing Δ3 and Δ3.5 of the ORF 9 promoter, gel retardation experiments were performed using two sets of overlapping oligonucleotides spanning the region, termed set I-II (Fig. 3a). The fluorescein-labeled oligonucleotides were incubated with 2.5 μl of nuclear extracts from untransfected or p50GFP- or p50GFPΔAT-hook-transfected 293T cells. The protein-nucleic acid complexes were then separated on a 4% polyacrylamide gel and detected by using a variable-mode imager (Fig. 3b). Results show the formation of a retarded complex with p50GFP-transfected cell extracts when incubated with oligonucleotide set II. No complexes were identified in the presence of 50GFPΔAT-hook, the mutant ORF 50a protein which lacks the DNA binding domain, suggesting that ORF 50a directly binds the ORF 9 promoter between 20307 and 20335 bp of the published sequence. Moreover, to demonstrate that the retarded complex detected is specific for ORF 50a, three further assays were performed. Firstly, the electrophoretic mobility shift assays (EMSAs) were repeated in the presence of 2.5 μl of control reticulocyte lysate or in vitro-translated ORF 50. A retarded complex was detected, demonstrating that ORF 50 directly binds to a RE contained within set II oligonucleotides (Fig. 3bii). Secondly, a supershift assay was undertaken using a monoclonal GFP antibody (Clontech) and a control B23 antibody (Santa Cruz) and the set II oligonucleotides. Two micrograms of GFP or control antibody was added to the binding reaction mixture described above and incubated for 15 min. A clear shift of the p50GFP retarded complex was detected (Fig. 3biii), in contrast to the control antibody (Fig. 3biv), demonstrating specificity for ORF 50a. However, a complete supershift of the ORF 50-GFP antibody complex was not observed; this was probably due to insufficient amounts of antibody being used in the assay. Thirdly, a competition assay was performed using increasing quantities of unlabeled oligonucleotides spanning the ORF 50 RE present within the ORF 6 promoter (42) or an unlabeled random oligonucleotide. The ORF 6 promoter competed out the binding reaction with labeled set II oligonucleotides (Fig. 3bv), in contrast to the random oligonucleotides (Fig. 3bvi), demonstrating that this retarded complex is specific for the ORF50a protein.

  • Open in new tab
  • Download powerpoint
  • Open in new tab
  • Download powerpoint
FIG. 3.

ORF 50 is capable of binding to a 2-bp AT-rich sequence within the ORF 9 promoter. (a) Schematic representation of the oligonucleotide primers spanning the ORF 50 RE within the ORF 9 promoter. (b) (i) EMSAs were performed using fluorescein-labeled oligonucleotides spanning the ORF 9 promoter. Each labeled set of oligonucleotides was incubated with nuclear extracts prepared from untransfected or p50GFP- or p50GFPΔAT-hook-transfected 293T cells. (ii) To demonstrate that ORF 50 interacts directly with set II oligonucleotides, oligonucleotides were incubated with control reticulocyte lysate or in vitro-translated ORF 50 and separated on a polyacrylamide gel. (iii) To demonstrate the specificity of ORF 50 binding, EMSAs were performed by the addition of primary monoclonal GFP antibody (Ab) (rightmost lane). (iv) To show the specificity of the GFP antibody, EMSAs were repeated with a control antibody (Ab). (v) To demonstrate the specificity of the ORF 50 interaction, increased quantities of unlabeled oligonucleotides spanning the ORF 50 RE present within the ORF 6 promoter were used to compete out the binding reaction. (vi) To show that the competition was specific to the ORF 6 oligonucleotides, EMSAs were repeated with a random oligonucleotide.

The ORF 50 response element within the ORF 9 promoter can function as an enhancer element.To further confirm the specificity of ORF 50a binding, the sequences between 20307 and 20335 bp of the published sequence (1) were cloned upstream of a heterologous promoter to determine whether they can confer ORF 50a responsiveness. Set II oligonucleotides were cloned by blunt-end ligation in both possible orientations into pGL3-promoter, which contains the luciferase reporter gene under the control of a minimal SV40 promoter, deriving pGL3-9REI and pGL3-9RE2 (Fig. 4a). Moreover, to compare the specificity of the ORF 9 promoter RE to previously identified REs, similar constructs were produced containing the REs from the ORF 6 and ORF 50 promoters (37). 293T cells were then transfected with 1 μg of pGL3-promoter, pGL3-50RE1, pGL3-9RE2, pGL3-6RE, or pGL3-9RE in the presence of 1 μg of pEGFP or p50GFP and assayed for luciferase activity by standard methods (Fig. 4b). Results demonstrated that p50GFP was able to increase luciferase activity to significant levels from all four RE constructs. This demonstrates that the sequences with the ORF 9 promoter RE are able to confer ORF 50a responsiveness to a heterologous promoter in either orientation and may function as an enhancer sequence. Levels of activity were similar between constructs containing the REs identified within the ORF 50 and ORF 9 promoter, which both contain AT-rich REs. In contrast, activation levels were 70% higher using the construct containing the canonical ORF 6 promoter RE. This variation in activity of ORF 50-responsive genes may have a role as a further regulatory step of the HVS lytic temporal cascade.

FIG. 4.
  • Open in new tab
  • Download powerpoint
FIG. 4.

The 28-bp AT-rich sequence can confer ORF 50 responsiveness to a heterologous promoter. (a) Schematic representation of the insertion of one copy of the 28-bp AT-rich sequence upstream of the enhancerless SV40 promoter in either orientation to generate pGL3-ORF9RE1 and pGL3-ORF9RE2. (b) 293T cells were cotransfected with pGL3-promoter, pGL3-ORF9REI, pGL3-ORF9REII, pGL3-ORF50RE, or pGL3-ORF6RE in the presence of pEGFP or p50GFP. Cells were harvested 30 h posttransfection, and cell lysates were assayed for luciferase activity by standard methods. The variations between three replicate assays are indicated. All luciferase data are presented as a percentage of luciferase activity compared to the p50GFP levels on the pORF9-Luc promoter construct, with the p50GFP level representing 100% activity.

Rta and C/EBPα proteins synergistically activate the ORF 9 promoter.Further analysis of the ORF 9 promoter using the transcription factor element-searching algorithm Matinspector (http://www.genomatix.com ) identified a number of potential transcription factor binding sites. Interestingly, multiple CCAAT enhancer binding protein alpha elements within the promoter and specifically two within the close vicinity of the ORF 50 RE were identified (Fig. 5a). Recent reports have demonstrated that C/EBPα is involved in KSHV ORF 50-mediated transactivation of DE promoters (38, 39). Therefore, the potential involvement of C/EBPα on the ORF 50a stimulation of the ORF 9 promoter was further investigated. To this end, 293T cells were transfected with either 1 μg of pEGFP, p50GFP, or pC/EBPα in the presence of 1 μg of pORF9-Luc. Cells were assayed for luciferase activity, and the data were normalized as described above (Fig. 5b). As a further control, RT-PCR was performed to demonstrate that ORF 50a was expressed in similar quantities in all transfected cells (Fig. 5c). Results showed that no significant increase in luciferase activity was observed in the presence of C/EBPα alone; in contrast, ORF50a activated the ORF 9 promoter to levels similar to those previously shown. However, upon coexpression of ORF 50a and C/EBPα, ORF 9 promoter activity was significantly enhanced to levels of approximately 160% of activity compared to ORF 50a alone (Fig. 5b). Therefore, this suggests that ORF 50a and C/EBPα may have a synergistic effect on the activation of the ORF 9 promoter.

FIG.5.
  • Open in new tab
  • Download powerpoint
FIG.5.

ORF 50 and C/EBPα synergistically activate the ORF 9 promoter. (a) Schematic representation of potential C/EBPα transcription factor binding sites within the ORF 9 promoter. (b) 293T cells were cotransfected with pORF9-Luc in the presence of pEGFP, p50GFP, pC/EBPα, or combinations of these. Cells were harvested 30 h posttransfection, and cell lysates were assayed for luciferase activity by standard methods. The variations between three replicate assays (each performed twice) are indicated. All luciferase data are presented as a percentage of the luciferase activity compared to the p50GFP level on the pORF9-Luc promoter construct, with the p50GFP level representing 100% activity. The increase in stimulation is shown as a value above each bar. (c) RT-PCR analysis to demonstrate that ORF 50 and C/EBPα were expressed in similar quantities in cells cotransfected with pORF9-Luc in the presence of pEGFP, p50GFP, pC/EBPα, or combinations of these. GAPDH, glyceraldehyde-3-phosphate dehydrogenase. (d) RT-PCR analysis of luciferase expression in cells cotransfected with pORF9-Luc in the presence of pEGFP, p50GFP, pC/EBPα, or combinations of these.

To ascertain whether this synergistic activation of the ORF 9 promoter in the presence of ORF 50a and C/EBPα is due to an increase in the levels of luciferase mRNA, RT-PCR analysis was performed. Total RNA was isolated from 293T cells transfected with either 1 μg of pEGFP, p50GFP, or pC/EBPα in the presence of 1 μg of pORF9-Luc, harvested 24 h posttransfection, and used in RT-PCR analysis to determine the quantities of luciferase mRNA (Fig. 5d). Results show that in the presence of ORF 50a, luciferase mRNA levels are increased; this is further enhanced upon expression of ORF50a and C/EBPα. This demonstrates that the transactivation of luciferase reporter gene from the ORF 9 promoter is due to an increase in mRNA levels.

In vitro-translated C/EBPα protein binds directly to the ORF 9 promoter, and binding is required for its transactivation. To determine whether C/EBPα binds directly to the ORF 9 promoter, gel retardation experiments were performed using the two sets of oligonucleotides spanning the region between Δ3 and Δ3.5, namely, set I-II; set I contains a putative C/EBPα binding site identified by Matinspector (Fig. 6a). The fluorescein-labeled oligonucleotides were incubated with 2.5 μl of control reticulocyte lysate or in vitro-translated C/EBPα or C/EBPβ. The protein-nucleic acid complexes were then separated and detected by using a variable-mode imager. Results showed the formation of a retarded complex with in vitro-translated C/EBPα when incubated with oligonucleotide set I (Fig. 6bi). No other complexes were identified, suggesting that C/EBPα directly binds the ORF 9 promoter between 20286 and 20314 bp of the published sequence (1). To confirm that the retarded complex detected is specific for C/EBPα, a competition assay was performed using increasing quantities of unlabeled oligonucleotides spanning a known C/EBPα binding previously identified in the KSHV Rta promoter site or a randomized oligonucleotide (39). The C/EBPα oligonucleotides competed out the binding reaction with labeled set I oligonucleotides (Fig. 6bii), in contrast to the random oligonucleotide, demonstrating that this retarded complex is specific for the C/EBPα protein.

FIG. 6.
  • Open in new tab
  • Download powerpoint
FIG. 6.

C/EBPα binds upstream of the ORF 50 RE within the ORF 9 promoter. (a) Schematic representation of the oligonucleotide primers spanning the potential C/EBPα binding site adjacent to the ORF 50 RE within the ORF 9 promoter. The potential C/EBPα binding site is boxed. (b) (i) EMSAs were performed using fluorescein-labeled set I and II oligonucleotides spanning the potential C/EBPα binding site. Each labeled set of oligonucleotides was incubated with control reticulocyte lysate or in vitro-translated C/EBPα or C/EBPβ and separated on a polyacrylamide gel. (ii) To demonstrate the specificity of the C/EBPα interaction, increased quantities of unlabeled oligonucleotides specific for the C/EBPα binding recognition sequence were used to compete out the binding reaction. EMSAs were repeated with a random oligonucleotide.

To further assess the role of C/EBPα on the activation of the ORF 9 promoter, a mutant promoter was generated that contained a deletion within the putative C/EBPα recognition sequence, termed pORF9ΔC/EBPα-Luc. 293T cells were then transfected with 1 μg of either pORF9-Luc or pORF9ΔC/EBPα-Luc in the presence of 1 μg of pEGFP, p50GFP, or pC/EBPα and assayed for luciferase activity as described previously (Fig. 7). The results demonstrated that p50GFP was capable of stimulating both pORF9-Luc or pORF9ΔC/EBPα-Luc to levels similar to those previously described. However, deletion of the putative C/EBPα region resulted in no enhancement of ORF 9 promoter activity in the presence of C/EBPα, in contrast to the wild-type ORF 9 promoter. This, therefore, suggests that the element responsible for the synergistic response of the ORF 9 promoter in the presence of C/EBPα comprises 20286 to 20314 bp of the published sequence within the ORF 9 promoter.

FIG. 7.
  • Open in new tab
  • Download powerpoint
FIG. 7.

C/EBPα binding is required for the synergistic response on the ORF 9 promoter. 293T cells were cotransfected with pORF9-Luc or p ORF9ΔC/EBPα-Luc in the presence of pEGFP, p50GFP, or pC/EBPα and various combinations of these. Cells were harvested 30 h posttransfection, and cell lysates were assayed for luciferase activity by standard methods. The variations between three replicate assays are indicated. All luciferase data are presented as a percentage of the luciferase activity compared to p50GFP levels on the pORF9-Luc promoter construct, with the p50GFP level representing 100% activity.

ORF 50a and C/EBPα interact with the ORF 9 promoter in vivo.To elucidate whether ORF 50a and C/EBPα interact with the ORF 9 promoter in vivo, ChIP assays were performed. Initially, HVS-infected OMK cells were harvested 24 h postinfection. Then ORF 50 and C/EBPα were immunoprecipitated using an anti-ORF 50 and anti-C/EBPα monoclonal sera, respectively, from the cross-linked protein-DNA complexes. After the removal of the proteins, the recovered DNA was used as a template for the specific amplification of the ORF 9 promoter and a HVS genomic control. Results demonstrate both ORF 50 and C/EBPα proteins were found to be specifically associated with the ORF 9 promoter (Fig. 8). In contrast, the negative controls with anti-ORF 57 antiserum failed to precipitate any ORF 9 promoter. This demonstrates that the ORF 50 and C/EBPα interact with the ORF 9 promoter in vivo.

FIG. 8.
  • Open in new tab
  • Download powerpoint
FIG. 8.

ORF 50 and C/EBPα associate with the ORF 9 promoter in vivo. (i) PCR amplification of the ORF 9 promoter from input, no-antibody controls (No Ab), immunoprecipitates with anti-ORF 57 antisera (ORF 57 Ab), immunoprecipitates with anti-ORF 50 antisera (ORF 50 Ab), or immunoprecipitates with anti-C/EBPα antisera (C/EBPα Ab). (ii) As a negative control, PCR amplification was performed on nonpromoter fragments, specifically the ORF 73 gene from input (lane 1), no-antibody controls (No Ab), immunoprecipitates with anti-ORF 57 antisera (ORF 57 Ab), immunoprecipitates with anti-ORF 50 antisera (ORF 50 Ab), or immunoprecipitates with anti-C/EBPα antisera (C/EBPα Ab).

ORF 50a interacts with C/EBPα.Gel retardation analysis has demonstrated that both ORF 50a and C/EBPα bind the ORF 9 promoter in close proximity. Therefore, to investigate whether ORF 50a and C/EBPα interact directly with each other, coimmunoprecipitation studies were performed. Control untransfected 293T cells were compared with cells transfected with pEGFP, p50GFP, and a myc-tagged expression construct for C/EBPα. After 24 h, the cells were harvested, and cell lysates were utilized in coimmunoprecipitation analysis using an anti-ORF 50 antiserum (17). Polypeptides precipitated from cellular extracts were then analyzed by Western blotting with anti-myc-specific antiserum (Fig. 9). The results demonstrate that C/EBPα coimmunoprecipitates in the presence of both ORF 50a and the mutant ORF 50a construct lacking its DNA binding domain. This result suggests that the HVS ORF 50a can specifically interact with C/EBPα and that the AT-hook DNA binding domain is not required for this interaction.

FIG. 9.
  • Open in new tab
  • Download powerpoint
FIG. 9.

ORF 50 and C/EBPα physically interact. 293T cells were cotransfected with pC/EBPα-myc in the presence of pEGFP, p50GFP, or p50GFPΔAT-hook. (a) Cell lysates were immunoprecipitated using anti-ORF 50 antiserum. Bound proteins were resolved by SDS-polyacrylamide gel electrophoresis, and the presence of C/EBPα-myc was detected using anti-myc antiserum (b). As a control, Western blot analysis was carried out on the supernatants before immunoprecipitation using a anti-GFP antibody to confirm that GFP, 50GFP, and 50GFPΔAT-hook were expressed in all samples. Untrans, untranslated.

DISCUSSION

HVS ORF 50a encodes the major transactivator protein which functions as a molecular switch for the transition from latency to lytic replication by specific transactivation of many downstream HVS promoters (36, 37, 41, 42). Homologues exist in all gammaherpesviruses, and all are thought to function as sequence-specific transactivators (5). HVS ORF 50 activates transcription by binding to specific REs within their promoter region. In at least some promoters, namely, ORF 6 and ORF 57, a consensus RE which shares homology with the EBV Rta RE element, GNCCN9GGNG, has been identified (41, 42). Guanine methylation studies have demonstrated that the CCN9GG motif is essential for EBV Rta binding, suggesting that Rta binds to adjacent major grooves of the DNA (15). A similar mechanism may function for HVS ORF 50a; however, the flanking sequences are significantly different from those of the EBV Rta REs, suggesting that gammaherpesvirus ORF 50 gene products have different sequences required for recognition and fixation of the proteins to their target. Recent analysis has also suggested that the HVS ORF 50a protein can activate HVS promoters with alternative REs, present within the ORF 50 (37) and now the ORF 9 gene promoters. At present, there is limited homology between these alternative REs; however, they are both AT-rich in sequence. We have recently demonstrated that direct binding to either type of ORF 50 RE requires the ORF 50 DNA binding domain which has a high degree of homology to the DNA binding domain encoded by mammalian HMGA chromosomal proteins termed AT-hook proteins (36). Previous studies have shown that mammalian AT-hook DNA binding domains have a dynamic confirmation, which allows them to produce optimal contact with the narrow minor groove regions of AT-rich DNA (18, 25, 26). Therefore, we suggest that due to the absence of direct homology between the two alternative types of REs, the AT-rich DNA sequences are probably being recognized by the ORF 50a AT-hook DNA binding domain.

Interestingly, although all these alternative REs confer ORF 50 responsiveness to a minimal heterologous promoter, the AT-rich RE contained in the ORF 50 and ORF 9 promoters may be less effective than the canonical ORF 6 promoter RE. Perhaps the alternative AT-rich RE found in the ORF 50 and ORF 9 promoters requires additional transcription factor binding sites to be adjacent to the ORF 50 RE for activity to occur, as in the case for the ORF 9 promoter. Although preliminary analysis of the ORF 50 promoter has not identified a similar synergistic response with C/EBPα (unpublished observations), this does not rule out the possibility that other cellular transcription factors enhance ORF 50 promoter activity. At present, it is unknown whether this is due to the direct binding affinity of ORF 50a to these different REs. A comparative study of KSHV ORF 50 binding to four target gene promoters demonstrated that KSHV Rta binds to their alternative REs with various affinities (31). This may result in and partially account for differences in the expression of these Rta target genes. This has led to the hypothesis that these variations in binding affinity could account for multiple genes being expressed at different levels during the lytic replication cycle. Interestingly, the cellular nonhistone DNA architectural protein, HMG-1 may enhance KSHV Rta binding to the REs, particularly the association with the lower affinity REs (32).

Although the ORF 50 binding affinity may have a role in the level of gene expression, indirect mechanisms are also implicated in gammaherpesvirus ORF 50-mediated transactivation. For example, KSHV Rta can be directed to target promoters through interaction with a cellular transcription factor, RBP-Jκ, the primary target of the Notch signaling pathway (19). This interaction targets Rta to the RBP-Jκ recognition sequence, resulting in the replacement of RBP-Jκ's repressive action with KSHV Rta-mediated activation. KSHV Rta has also been shown to interact with the cellular DNA binding protein CCAAT/enhancer binding protein alpha (38, 39). C/EBPα is a member of the leucine zipper family of transcription factors which include c-JUN, c-FOS, ATF, and CREB (24). They have a conserved bZIP domain at the carboxy terminus which can bind to specific DNA binding elements, thereby modulating gene expression (21, 24). C/EBPα promotes differentiation and inhibits cell proliferation, leading to G1 cell cycle arrest by various mechanisms (21, 24). As gammaherpesviruses induce G1 cell cycle arrest during the lytic cycle and C/EBPα is known to inhibit cell proliferation, it is an ideal candidate for recruitment by the herpesviruses during the transition into lytic replication. To support this hypothesis, recent analysis has identified an interaction between C/EBPα and the KSHV Rta and RAP proteins (38, 39). These interactions lead to the activation of both viral and cellular protein during the lytic replication cycle via a presumed indirect “piggyback” targeting mechanism to C/EBPα binding sites on the gene promoters and also the stabilization of the C/EBPα protein. In addition, a direct mechanism may also activate DE KSHV promoters, such as the PAN promoter, whereby ORF 50 and C/EBPα both interact directly with the promoter and the ORF50-C/EBPα protein complex cooperatively transactivates the promoter (39).

Herein, we have also demonstrated a direct interaction between HVS ORF 50a and C/EBPα and shown they act synergistically to activate the ORF 9 promoter. This demonstrates that C/EBPα may well be involved in ORF 50-mediated transactivation of HVS lytic replication. However, the data may not suggest a piggyback mechanism at work. This is supported by a number of points; ORF 50a protein can interact directly with the promoter using gel retardation analysis, but more convincing, the mutant ORF 50a protein 50GFPΔAT-hook (which contains a deletion of the ORF50 DNA binding domain) fails to transactivate the ORF 9 promoter and C/EBPα alone was unable to transactivate the ORF 9 promoter. These results suggest that ORF 50a binding is essential for ORF 9 promoter transactivation. Therefore, it is likely that ORF 50a alone is sufficient to activate the ORF 9 promoter by binding to the AT-rich RE. However, an additional cooperative transactivation mechanism exists; this mechanism is mediated by the ORF50-C/EBPα protein complex binding directly to their adjacent binding sites.

In summary, this investigation has described a likely mechanism for the ORF 50-mediated transactivation of the ORF 9 HVS DNA polymerase in the early stages of the lytic replication cycle. This involves the direct binding of both ORF 50a and C/EBPα to adjacent binding sites within the ORF 9 promoter.

ACKNOWLEDGMENTS

This work was supported in part from grants to A.W. from the Association of International Cancer Research, BBSRC, and Yorkshire Cancer Research.

We thank C. Heckman, Stanford University School of Medicine, for kindly providing cEBPα and cEBPβ expression constructs.

FOOTNOTES

    • Received 18 May 2005.
    • Accepted 1 August 2005.
  • Copyright © 2005 American Society for Microbiology

REFERENCES

  1. 1.↵
    Albrecht, J., J. Nicholas, D. Biller, K. Cameron, B. Biesinger, C. Newman, S. Wittmann, M. Craxton, H. Coleman, and B. Fleckenstein. 1992. Primary structure of the herpesvirus saimiri genome. J. Virol.66:5047-5058.
    OpenUrlAbstract/FREE Full Text
  2. 2.↵
    Andrews, N. C., and D. V. Faller. 1991. A rapid micropreparation technique for extraction of DNA-binding proteins from limiting numbers of mammalian cells. Nucleic Acids Res.19:2499.
    OpenUrlCrossRefPubMedWeb of Science
  3. 3.↵
    Biesinger, B., I. Muller-Fleckenstein, B. Simmer, G. Lang, S. Wittmann, E. Platzer, R. C. Desrosiers, and B. Fleckenstein. 1992. Stable growth transformation of human T lymphocytes by herpesvirus saimiri. Proc. Natl. Acad. Sci. USA89:3116-3119.
    OpenUrlAbstract/FREE Full Text
  4. 4.↵
    Biesinger, B., J. J. Trimble, R. C. Desrosiers, and B. Fleckenstein. 1990. The divergence between two oncogenic herpesvirus saimiri strains in a genomic region related to the transforming phenotype. Virology176:505-514.
    OpenUrlCrossRefPubMedWeb of Science
  5. 5.↵
    Damania, B., J. H. Jeong, B. S. Bowser, S. M. DeWire, M. R. Staudt, and D. P. Dittmer. 2004. Comparison of the Rta/Orf50 transactivator proteins of gamma-2-herpesviruses. J. Virol.78:5491-5499.
    OpenUrlAbstract/FREE Full Text
  6. 6.↵
    Desrosiers, R. C., P. S. Daniel, L. M. Waldron, and N. L. Letvin. 1986. Nononcogenic deletion mutants of herpesvirus saimiri are defective for in vitro immortalization. J. Virol.57:701-705.
    OpenUrlAbstract/FREE Full Text
  7. 7.↵
    Dourmishev, L. A., A. L. Dourmishev, D. Palmeri, R. A. Schwartz, and D. M. Lukac. 2003. Molecular genetics of Kaposi's sarcoma-associated herpesvirus (human herpesvirus 8) epidemiology and pathogenesis. Microbiol. Mol. Biol. Rev.67:175-212.
    OpenUrlAbstract/FREE Full Text
  8. 8.↵
    Fickenscher, H., and B. Fleckenstein. 2001. Herpesvirus saimiri. Philos. Trans. R. Soc.356:545-567.
    OpenUrl
  9. 9.↵
    Fleckenstein, B., and R. C. Desrosiers. 1982. Herpesvirus saimiri and herpesvirus ateles, p. 253-332. In B. Roizman (ed.), The herpesviruses, vol. 1. Plenum Press, New York, N.Y.
    OpenUrl
  10. 10.↵
    Goodwin, D. J., M. S. Walters, P. G. Smith, M. Thurau, H. Fickenscher, and A. Whitehouse. 2001. Herpesvirus saimiri open reading frame 50 (Rta) protein reactivates the lytic replication cycle in a persistently infected A549 cell line. J. Virol.75:4008-4013.
    OpenUrlAbstract/FREE Full Text
  11. 11.↵
    Gradoville, L., J. Gerlach, E. Grogan, D. Shedd, S. Nikiforow, C. Metroka, and G. Miller. 2000. Kaposi's sarcoma-associated herpesvirus open reading frame 50/Rta protein activates the entire viral lytic cycle in the HH-B2 primary effusion lymphoma cell line. J. Virol.74:6207-6212.
    OpenUrlAbstract/FREE Full Text
  12. 12.↵
    Grassmann, R., and B. Fleckenstein. 1989. Selectable recombinant herpesvirus saimiri is capable of persisting in a human T-cell line. J. Virol.63:1818-1821.
    OpenUrlAbstract/FREE Full Text
  13. 13.↵
    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.
    OpenUrlAbstract/FREE Full Text
  14. 14.
    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.
    OpenUrlCrossRefPubMedWeb of Science
  15. 15.↵
    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.
    OpenUrlCrossRefPubMedWeb of Science
  16. 16.↵
    Hall, K. T., M. S. Giles, D. J. Goodwin, M. A. Calderwood, I. M. Carr, A. J. Stevenson, A. F. Markham, and A. Whitehouse. 2000. Analysis of gene expression in a human cell line stably transduced with herpesvirus saimiri. J. Virol.74:7331-7337.
    OpenUrlAbstract/FREE Full Text
  17. 17.↵
    Hall, K. T., A. J. Stevenson, D. J. Goodwin, P. C. Gibson, A. F. Markham, and A. Whitehouse. 1999. The activation domain of herpesvirus saimiri R protein interacts with the TATA-binding protein. J. Virol.73:9756-9763.
    OpenUrlAbstract/FREE Full Text
  18. 18.↵
    Himes, S. R., R. Reeves, J. Attema, M. Nissen, Y. Li, and M. F. Shannon. 2000. The role of high-mobility group I(Y) proteins in expression of IL-2 and T cell proliferation. J. Immunol.164:3157-3168.
    OpenUrlAbstract/FREE Full Text
  19. 19.↵
    Liang, Y., J. Chang, S. J. Lynch, D. M. Lukac, and D. Ganem. 2002. The lytic switch protein of KSHV activates gene expression via functional interaction with RBP-Jκ (CSL), the target of the Notch signaling pathway. Genes Dev.16:1977-1989.
    OpenUrlAbstract/FREE Full Text
  20. 20.↵
    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. Virology252:304-312.
    OpenUrlCrossRefPubMedWeb of Science
  21. 21.↵
    McKnight, S. L. 2001. McBindall—a better name for CCAAT/enhancer binding proteins? Cell107:259-261.
    OpenUrlCrossRefPubMedWeb of Science
  22. 22.↵
    Neipel, F., J. Albrecht, and B. Fleckenstein. 1997. Cell-homologous genes in the Kaposi's sarcoma-associated rhadinovirus human herpesvirus 8: determinants of its pathogenicity? J. Virol.71:4187-4192.
    OpenUrlFREE Full Text
  23. 23.↵
    Nicholas, J., L. S. Coles, C. Newman, and R. W. Honess. 1991. Regulation of the herpesvirus saimiri (HVS) delayed-early 110-kilodalton promoter by HVS immediate-early gene products and a homolog of the Epstein-Barr virus R trans activator. J. Virol.65:2457-2466.
    OpenUrlAbstract/FREE Full Text
  24. 24.↵
    Ramji, D. P., and P. Foka. 2002. CCAAT/enhancer-binding proteins: structure, function and regulation. Biochem. J.365:561-575.
    OpenUrlCrossRefPubMedWeb of Science
  25. 25.↵
    Reeves, R., and L. Beckerbauer. 2001. HMGI/Y proteins: flexible regulators of transcription and chromatin structure. Biochim. Biophys. Acta1519:13-29.
    OpenUrlCrossRefPubMedWeb of Science
  26. 26.↵
    Reeves, R., and M. S. Nissen. 1990. The A.T-DNA-binding domain of mammalian high mobility group I chromosomal proteins. A novel peptide motif for recognizing DNA structure. J. Biol. Chem.265:8573-8582.
    OpenUrlAbstract/FREE Full Text
  27. 27.↵
    Russo, J. J., R. A. Bohenzky, M.-C. Chien, J. Chen, M. Yan, D. Maddalena, J. P. Parry, D. Peruzzi, I. S. Edelman, Y. Chang, and P. S. Moore. 1996. Nucleotide sequence of the Kaposi sarcoma-associated herpesvirus (HHV8). Proc. Natl. Acad. Sci. USA93:14862-14867.
    OpenUrlAbstract/FREE Full Text
  28. 28.↵
    Simmer, B., M. Alt, I. Buckreus, S. Berthold, B. Fleckenstein, E. Platzer, and R. Grassmann. 1991. Persistence of selectable herpesvirus saimiri in various human haematopoietic and epithelial cell lines. J. Gen. Virol.72:1953-1958.
    OpenUrlCrossRefPubMed
  29. 29.
    Smith, P. G., S. A. Burchill, D. Brooke, P. L. Coletta, and A. Whitehouse. 2005. Efficient infection and persistence of a herpesvirus saimiri-based gene delivery vector into human tumour xenografts and multicellular spheroid cultures. Cancer Gene Ther.12:248-256.
    OpenUrlCrossRefPubMed
  30. 30.↵
    Smith, P. G., P. L. Coletta, A. F. Markham, and A. Whitehouse. 2001. In vivo episomal maintenance of a herpesvirus saimiri-based gene delivery vector. Gene Ther.8:1762-1769.
    OpenUrlCrossRefPubMed
  31. 31.↵
    Song, M. J., H. Deng, and R. Sun. 2003. Comparative study of regulation of RTA-responsive genes in Kaposi's sarcoma-associated herpesvirus/human herpesvirus 8. J. Virol.77:9451-9462.
    OpenUrlAbstract/FREE Full Text
  32. 32.↵
    Song, M. J., S. Hwang, W. Wong, J. Round, D. Martinez-Guzman, Y. Turpaz, J. Liang, B. Wong, R. C. Johnson, M. Carey, and R. Sun. 2004. The DNA architectural protein HMGB1 facilitates RTA-mediated viral gene expression in gamma-2 herpesviruses. J. Virol.78:12940-12950.
    OpenUrlAbstract/FREE Full Text
  33. 33.↵
    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. USA95:10866-10871.
    OpenUrlAbstract/FREE Full Text
  34. 34.↵
    Szomolanyi, E., P. Medveczky, and C. Mulder. 1987. In vitro immortalization of marmoset cells with three subgroups of herpesvirus saimiri. J. Virol.61:3485-3490.
    OpenUrlAbstract/FREE Full Text
  35. 35.↵
    Virgin, H., IV, P. Latreille, P. Wamsley, K. Hallsworth, K. Weck, A. Dal Canto, and S. Speck. 1997. Complete sequence and genomic analysis of murine gammaherpesvirus 68. J. Virol.71:5894-5904.
    OpenUrlAbstract/FREE Full Text
  36. 36.↵
    Walters, M. S., K. T. Hall, and A. Whitehouse. 2004. The herpesvirus saimiri open reading frame (ORF) 50 (Rta) protein contains an AT hook required for binding to the ORF 50 response element in delayed-early promoters. J. Virol.78:4936-4942.
    OpenUrlAbstract/FREE Full Text
  37. 37.↵
    Walters, M. S., K. T. Hall, and A. Whitehouse. 2005. The herpesvirus saimiri Rta protein autostimulates via binding to a non-consensus response element. J. Gen. Virol.86:581-587.
    OpenUrlCrossRefPubMed
  38. 38.↵
    Wang, S. E., F. Y. Wu, M. Fujimuro, J. Zong, S. D. Hayward, and G. S. Hayward. 2003. Role of CCAAT/enhancer-binding protein alpha (C/EBPα) in activation of the Kaposi's sarcoma-associated herpesvirus (KSHV) lytic-cycle replication-associated protein (RAP) promoter in cooperation with the KSHV replication and transcription activator (RTA) and RAP. J. Virol.77:600-623.
    OpenUrlAbstract/FREE Full Text
  39. 39.↵
    Wang, S. E., F. Y. Wu, Y. Yu, and G. S. Hayward. 2003. CCAAT/enhancer-binding protein-alpha is induced during the early stages of Kaposi's sarcoma-associated herpesvirus (KSHV) lytic cycle reactivation and together with the KSHV replication and transcription activator (RTA) cooperatively stimulates the viral RTA, MTA, and PAN promoters. J. Virol.77:9590-9612.
    OpenUrlAbstract/FREE Full Text
  40. 40.↵
    West, J. T., and C. Wood. 2003. The role of Kaposi's sarcoma-associated herpesvirus/human herpesvirus-8 regulator of transcription activation (RTA) in control of gene expression. Oncogene22:5150-5163.
    OpenUrlCrossRefPubMed
  41. 41.↵
    Whitehouse, A., M. Cooper, K. T. Hall, and D. M. Meredith. 1998. The open reading frame (ORF) 50a gene product regulates ORF 57 gene expression in herpesvirus saimiri. J. Virol.72:1967-1973.
    OpenUrlAbstract/FREE Full Text
  42. 42.↵
    Whitehouse, A., A. Stevenson, M. Cooper, and D. Meredith. 1997. Identification of a cis-acting element within the herpesvirus saimiri ORF 6 promoter that is responsive to the HVS.R transactivator. J. Gen. Virol.78:1411-1417.
    OpenUrlCrossRefPubMed
  43. 43.↵
    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.
    OpenUrlAbstract/FREE Full Text
PreviousNext
Back to top
Download PDF
Citation Tools
The Herpesvirus Saimiri Replication and Transcription Activator Acts Synergistically with CCAAT Enhancer Binding Protein Alpha To Activate the DNA Polymerase Promoter
Louise Wakenshaw, Matthew S. Walters, Adrian Whitehouse
Journal of Virology Oct 2005, 79 (21) 13548-13560; DOI: 10.1128/JVI.79.21.13548-13560.2005

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Print

Alerts
Sign In to Email Alerts with your Email Address
Email

Thank you for sharing this Journal of Virology article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
The Herpesvirus Saimiri Replication and Transcription Activator Acts Synergistically with CCAAT Enhancer Binding Protein Alpha To Activate the DNA Polymerase Promoter
(Your Name) has forwarded a page to you from Journal of Virology
(Your Name) thought you would be interested in this article in Journal of Virology.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Share
The Herpesvirus Saimiri Replication and Transcription Activator Acts Synergistically with CCAAT Enhancer Binding Protein Alpha To Activate the DNA Polymerase Promoter
Louise Wakenshaw, Matthew S. Walters, Adrian Whitehouse
Journal of Virology Oct 2005, 79 (21) 13548-13560; DOI: 10.1128/JVI.79.21.13548-13560.2005
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Top
  • Article
    • ABSTRACT
    • MATERIALS AND METHODS
    • RESULTS
    • DISCUSSION
    • ACKNOWLEDGMENTS
    • FOOTNOTES
    • REFERENCES
  • Figures & Data
  • Info & Metrics
  • PDF

KEYWORDS

CCAAT-Enhancer-Binding Protein-alpha
DNA-Directed DNA Polymerase
Herpesvirus 2, Saimiriine
Immediate-Early Proteins
Trans-Activators
transcriptional activation
Viral Proteins

Related Articles

Cited By...

About

  • About JVI
  • Editor in Chief
  • Editorial Board
  • Policies
  • For Reviewers
  • For the Media
  • For Librarians
  • For Advertisers
  • Alerts
  • RSS
  • FAQ
  • Permissions
  • Journal Announcements

Authors

  • ASM Author Center
  • Submit a Manuscript
  • Article Types
  • Ethics
  • Contact Us

Follow #Jvirology

@ASMicrobiology

       

 

JVI in collaboration with

American Society for Virology

ASM Journals

ASM journals are the most prominent publications in the field, delivering up-to-date and authoritative coverage of both basic and clinical microbiology.

About ASM | Contact Us | Press Room

 

ASM is a member of

Scientific Society Publisher Alliance

 

American Society for Microbiology
1752 N St. NW
Washington, DC 20036
Phone: (202) 737-3600

Copyright © 2021 American Society for Microbiology | Privacy Policy | Website feedback

Print ISSN: 0022-538X; Online ISSN: 1098-5514