Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter...

18
Mutations of amino acids in the DNA-recognition domain of EpsteinBarr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments Richard Park a , Lee Heston b , Duane Shedd b , Henri-Jacques Delecluse d , George Miller a,b,c, a Department of Molecular Biophysics and Biochemistry, Yale University School of Medicine, New Haven, CT 06520, USA b Department of Pediatrics, Yale University School of Medicine, New Haven, CT 06520, USA c Department of Epidemiology and Public Health, Yale University School of Medicine, New Haven, CT 06520, USA d Department of Tumor Virology, German Cancer Research Center, Im Neuenheimer Feld 242, Heidelberg, Germany abstract article info Article history: Received 2 June 2008 Returned to author for revision 31 July 2008 Accepted 8 September 2008 Available online 19 October 2008 Keywords: EpsteinBarr virus ZEBRA EA-D Nuclear localization Immunouorescence Viral replication compartments ZEBRA, a transcription factor and DNA replication protein encoded by the EpsteinBarr virus (EBV) BZLF1 gene, plays indispensable roles in the EBV lytic cycle. We recently described the phenotypes of 46 single amino acid substitutions introduced into the DNA-recognition region of ZEBRA [Heston, L., El-Guindy, A., Countryman, J., Dela Cruz, C., Delecluse, H.J., and Miller, G. 2006]. The 27 DNA-binding-procient mutants exhibited distinct defects in their ability to activate expression of the kinetic classes of viral genes. Four phenotypic variants could be discerned: wild-type, defective at activating Rta, defective at activating early genes, and defective at activating late genes. Here we analyze the distribution of ZEBRA within the nucleus and the localization of EA-D (the viral DNA polymerase processivity factor), an indicator of the development of replication compartments, in representatives of each phenotypic group. Plasmids encoding wild-type (WT) and mutant ZEBRA were transfected into 293 cells containing EBV-bacmids. WT ZEBRA protein was diffusely and smoothly distributed throughout the nucleus, sparing nucleoli, and partially recruited to globular replication compartments. EA-D induced by WT ZEBRA was present diffusely in some cells and concentrated in globular replication compartments in other cells. The distribution of ZEBRA and EA-D proteins was identical to WT following transfection of K188R, a mutant with a conservative change. The distribution of S186A mutant ZEBRA protein, defective for activation of Rta and EA-D, was identical to WT, except that the mutant ZEBRA was never found in globular compartments. Co-expression of Rta with S186A mutant rescued diffuse EA-D but not globular replication compartments. The most striking observation was that several mutant ZEBRA proteins defective in activating EA-D (R179A, K181A and A185V) and defective in activating lytic viral DNA replication and late genes (Y180E and K188A) were localized to numerous punctate foci. The speckled appearance of R179A and Y180E was more regular and clearly dened in EBV-positive than in EBV-negative 293 cells. The Y180E late-mutant induced EA-D, but prevented EA-D from localizing to globular replication compartments. These results show that individual amino acids within the basic domain inuence localization of the ZEBRA protein and its capacity to induce EA-D to become located in mature viral replication compartments. Furthermore, these mutant ZEBRA proteins delineate several stages in the processes of nuclear re-organization which accompany lytic EBV replication. © 2008 Elsevier Inc. All rights reserved. Introduction EpsteinBarr virus (EBV) infection of the host begins in oral and pharyngeal epithelia where the virus replicates in a lytic manner. The virus subsequently spreads to B lymphocytes where it persists in a latent state for the lifetime of the host. (Rickinson and Kieff, 1996). During latency, EBV gene expression is restricted to a subset of viral genes and no new infectious virus is produced. The production of infectious virus and propagation of the virus from cell to cell and host to host is dependent upon the reactivation of a subpopulation of latent EBV genomes into the lytic phase of infection. In the lytic phase, a sequential cascade of viral gene expression results in a 1001000 fold amplication of the EBV genome, packaging of unit-length viral genomes into capsids, tegumentation, envelopment, the release of infectious virions, and, eventually, host cell lysis. The switch from the latent to lytic phases is triggered by a major viral regulatory protein, ZEBRA, encoded by the BZLF1 gene (Countryman and Miller, 1985). ZEBRA is sufcient to disrupt EBV latency and to initiate the lytic program in all cell types, and is also required for continued progression of the lytic program through viral late gene expression. Virology 382 (2008) 145162 Corresponding author. Department of Pediatrics, Yale University School of Medicine, Room LSOG 420, New Haven, CT 06520, USA. Fax: +1 203 785 6961. E-mail address: [email protected] (G. Miller). 0042-6822/$ see front matter © 2008 Elsevier Inc. All rights reserved. doi:10.1016/j.virol.2008.09.009 Contents lists available at ScienceDirect Virology journal homepage: www.elsevier.com/locate/yviro

Transcript of Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter...

Page 1: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

Virology 382 (2008) 145–162

Contents lists available at ScienceDirect

Virology

j ourna l homepage: www.e lsev ie r.com/ locate /yv i ro

Mutations of amino acids in the DNA-recognition domain of Epstein–Barrvirus ZEBRA protein alter its sub-nuclear localization and affect formation ofreplication compartments

Richard Park a, Lee Heston b, Duane Shedd b, Henri-Jacques Delecluse d, George Miller a,b,c,⁎a Department of Molecular Biophysics and Biochemistry, Yale University School of Medicine, New Haven, CT 06520, USAb Department of Pediatrics, Yale University School of Medicine, New Haven, CT 06520, USAc Department of Epidemiology and Public Health, Yale University School of Medicine, New Haven, CT 06520, USAd Department of Tumor Virology, German Cancer Research Center, Im Neuenheimer Feld 242, Heidelberg, Germany

⁎ Corresponding author. Department of PediatricMedicine, Room LSOG 420, New Haven, CT 06520, USA.

E-mail address: [email protected] (G. Miller).

0042-6822/$ – see front matter © 2008 Elsevier Inc. Aldoi:10.1016/j.virol.2008.09.009

a b s t r a c t

a r t i c l e i n f o

Article history:

ZEBRA, a transcription fact Received 2 June 2008Returned to author for revision 31 July 2008Accepted 8 September 2008Available online 19 October 2008

Keywords:Epstein–Barr virusZEBRAEA-DNuclear localizationImmunofluorescenceViral replication compartments

or and DNA replication protein encoded by the Epstein–Barr virus (EBV) BZLF1gene, plays indispensable roles in the EBV lytic cycle. We recently described the phenotypes of 46 singleamino acid substitutions introduced into the DNA-recognition region of ZEBRA [Heston, L., El-Guindy, A.,Countryman, J., Dela Cruz, C., Delecluse, H.J., and Miller, G. 2006]. The 27 DNA-binding-proficient mutantsexhibited distinct defects in their ability to activate expression of the kinetic classes of viral genes. Fourphenotypic variants could be discerned: wild-type, defective at activating Rta, defective at activating earlygenes, and defective at activating late genes. Here we analyze the distribution of ZEBRA within the nucleusand the localization of EA-D (the viral DNA polymerase processivity factor), an indicator of the developmentof replication compartments, in representatives of each phenotypic group. Plasmids encoding wild-type(WT) and mutant ZEBRA were transfected into 293 cells containing EBV-bacmids. WT ZEBRA protein wasdiffusely and smoothly distributed throughout the nucleus, sparing nucleoli, and partially recruited toglobular replication compartments. EA-D induced by WT ZEBRA was present diffusely in some cells andconcentrated in globular replication compartments in other cells. The distribution of ZEBRA and EA-Dproteins was identical to WT following transfection of K188R, a mutant with a conservative change. Thedistribution of S186A mutant ZEBRA protein, defective for activation of Rta and EA-D, was identical to WT,except that the mutant ZEBRA was never found in globular compartments. Co-expression of Rta with S186Amutant rescued diffuse EA-D but not globular replication compartments. The most striking observation wasthat several mutant ZEBRA proteins defective in activating EA-D (R179A, K181A and A185V) and defective inactivating lytic viral DNA replication and late genes (Y180E and K188A) were localized to numerous punctatefoci. The speckled appearance of R179A and Y180E was more regular and clearly defined in EBV-positive thanin EBV-negative 293 cells. The Y180E late-mutant induced EA-D, but prevented EA-D from localizing toglobular replication compartments. These results show that individual amino acids within the basic domaininfluence localization of the ZEBRA protein and its capacity to induce EA-D to become located in mature viralreplication compartments. Furthermore, these mutant ZEBRA proteins delineate several stages in theprocesses of nuclear re-organization which accompany lytic EBV replication.

© 2008 Elsevier Inc. All rights reserved.

Introduction

Epstein–Barr virus (EBV) infection of the host begins in oral andpharyngeal epithelia where the virus replicates in a lytic manner. Thevirus subsequently spreads to B lymphocytes where it persists in alatent state for the lifetime of the host. (Rickinson and Kieff, 1996).During latency, EBV gene expression is restricted to a subset of viralgenes and no new infectious virus is produced. The production of

s, Yale University School ofFax: +1 203 785 6961.

l rights reserved.

infectious virus and propagation of the virus from cell to cell and hostto host is dependent upon the reactivation of a subpopulation of latentEBV genomes into the lytic phase of infection. In the lytic phase, asequential cascade of viral gene expression results in a 100–1000 foldamplification of the EBV genome, packaging of unit-length viralgenomes into capsids, tegumentation, envelopment, the release ofinfectious virions, and, eventually, host cell lysis. The switch from thelatent to lytic phases is triggered by a major viral regulatory protein,ZEBRA, encoded by the BZLF1 gene (Countryman and Miller, 1985).ZEBRA is sufficient to disrupt EBV latency and to initiate the lyticprogram in all cell types, and is also required for continuedprogression of the lytic program through viral late gene expression.

Page 2: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

146 R. Park et al. / Virology 382 (2008) 145–162

ZEBRA is a 245 amino acid phosphoprotein belonging to the basiczipper (bZIP) family of proteins. Like other bZIP proteins, ZEBRAcontains a transactivation domain (aa 1 to 166), a basic DNA-recognition region (aa 178 to 194), and a coiled-coil dimerizationdomain (aa195 to 225) (Farrell et al. 1989; Chang and Liu, 2000; Chiand Carey, 1993; Flemington and Speck, 1990; Flemington et al., 1992;Taylor et al., 1991). Unlike cellular bZIP proteins, ZEBRA contains a C-terminal “tail” which may stabilize the homodimer (Petosa et al.,

Fig. 1. Intra-nuclear distribution of ZEBRA, EA-D, Rta and nucleolar proteins during lytic actransfected with wild-type ZEBRA and double stained for ZEBRA (iii, vi) and EA-D (i, ii, iv, v); Z(xiii–xvi). Digital images of single channel emissions and simultaneous EA-D-transmitted ligreduced or increased ZEBRA staining, respectively.

2006). ZEBRA performsmany functions necessary for lytic replication:(i) as a transcription factor, ZEBRA binds to heptamer motifs, termedZEBRA response elements (ZRE's), located within the promoters oflytic viral genes and activates gene expression. Some genes, such asBRLF1, encoding another viral transcription factor, Rta, are activatedindependently by ZEBRA. Other genes, such as BMRF1, encoding theDNA polymerase processivity factor, are activated by ZEBRA in synergywith Rta (Liu and Speck, 2003; Feederle et al., 2000; Lieberman et al.,

tivation of 293 cells containing EBV-bacmids. BZKO (i–iii) and 2089 cells (iv–xvi) wereEBRA (vii, ix) and nucleolin (viii, ix); or ZEBRA (x, xii) and fibrillarin (xi, xii); Rta and EA-Dht images (ii, v, xiv) are shown. Blue and yellow arrows indicate nuclear sub-regions of

Page 3: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

147R. Park et al. / Virology 382 (2008) 145–162

1990; Quinlivan et al., 1993). (ii) As an essential DNA replicationprotein, ZEBRA binds directly to the origin of lytic DNA replication(oriLyt) which contains seven ZRE's. Binding of ZEBRA to oriLyt isregulated by phosphorylation of residue S173 (El-Guindy et al., 2007).Binding of ZEBRA to oriLyt is a prerequisite for viral DNA replication,and is believed to recruit and stabilize components of the core viralreplication complex at oriLyt (Fixman et al., 1995; Hammerschmidtand Sugden, 1988; Schepers et al., 1993, 1996). A direct physicalinteraction between ZEBRA and several viral replication proteinsincluding the viral helicase, primase, primase-associated factor,polymerase and polymerase processivity factor has been demon-strated (Baumann et al., 1999; Gao et al., 1998; Zhang et al., 1996; Liaoet al., 2001). (iii) ZEBRA also induces profound alterations in host cells.For example, ZEBRA arrests the cell cycle at G0/G1 in EBV-negativecells (Cayrol and Flemington, 1996). ZEBRA disrupts PML bodies,causing ND10-associated proteins to disperse during lytic replication(Adamson and Kenney, 2001; Bell et al., 2000). ZEBRA modulates theactivities of several cellular transcription factors including C/EBP, NF-KB p65, and c-Myb (Morrison and Kenney, 2004; Morrison et al.,2004; Kenney et al., 1992; Wu et al., 2003).

Nuclear compartmentalization of lytic viral processes is acharacteristic common to all herpesviruses, including herpes simplexviruses (HSV), cytomegalovirus (CMV) and EBV. Lytic EBV DNAsynthesis was observed by FISH to occur in discrete intra-nuclearglobular sub-domains (Gradoville et al., 2002), similar to the so-calledreplication compartments studied intensively in HSV infection(Lukonis and Weller, 1996; Liptak et al., 1996). However, the role ofZEBRA in the formation of these replication compartments hasreceived little investigation. In detergent-extracted B95-8 cellscontaining an inducible BZLF1 gene, ZEBRA protein localized diffuselythroughout the nucleus in a fine-granular staining pattern during the

Fig. 2. Three distribution patterns of intra-nuclear EA-D in 293 cells containing an EBV-bacmwere selected at random and scored for their EA-D phenotype.

early stages of lytic induction. At later stages of infection, after theonset of viral DNA replication, ZEBRA redistributed from a diffusepattern into distinct co-localization with globular EBV replicationcompartments (Daikoku et al., 2005). In a separate study, ZEBRA waspresent diffusely throughout the nucleus in D98/HR-1 cells transfectedwith a BZLF1 expression vector; however ZEBRA did not redistribute toglobular replication compartments (Bell et al., 2000). These differingobservations may be attributed to differences in cell type, detergentextraction of soluble proteins, or overexpression of ZEBRA.

Previously, we created a comprehensive series of 46 single aminoacid substitutions throughout the basic DNA-recognition region ofZEBRA and demonstrated that specific amino acids within this regionplay critical and distinct roles in the capacity of ZEBRA to bind DNA, toactivate lytic gene expression, and to promote lytic viral DNAreplication (Heston et al., 2006). Nineteen of the 46 point mutantswhich did not bind DNA containing a high affinity ZRE, ZIIIBcompletely lost the ability to activate expression of all viral genesassayed. Twenty-seven ZEBRA mutants retained the capacity to bindZIIIB DNA, although with differences in affinity. These DNA-binding-competent ZEBRAmutants were classified into four functional groups:(i) eight mutants were similar to wild-type and showed no defect inthe ability to activate transcription of the Rta gene BRLF1, the earlygene BMRF1, or the late genes BFRF3 and BLRF2. (ii) Nine mutantswere defective in activating Rta expression and all other downstreamgenes. (iii) Three mutants were competent to activate Rta expression,yet were defective in activation of the BMRF1 product, EA-D. (iv)Seven ZEBRA mutants were not defective in activating Rta or BMRF1expression, but were unable to promote lytic viral DNA synthesis andexpression of the BFRF3 and BLRF2 late genes.

The alterations of activity resulting from point mutants in the DNA-recognition region of ZEBRA raised two questions. First, do these

id. 2089 cells were transfected with wild-type ZEBRA, fixed, then stained for EA-D. Cells

Page 4: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

148 R. Park et al. / Virology 382 (2008) 145–162

mutations cause specific changes in the intra-nuclear localization ofZEBRA? Second, do the mutations interfere with the formation ofreplication compartments? To address these questions, and to furtherthe understanding of the spatial organization and nuclear compart-mentalization of EBV lytic processes, we performed a series of indirectimmunofluorescence analyses (IFA) to observe and compare thedistributions of wild-type ZEBRA with mutant forms of ZEBRA. Wealso studied the effects of thesemutationsupon thedistributionof EA-D,a marker for formation of replication compartments. We report herethat point mutations within the basic DNA-binding domain of ZEBRAcause specific changes in ZEBRA localization, and affect the formation ofviral replication compartments.

Fig. 3. Effect of amino acid substitutions within the basic region on localization of ZEBRA andcells containing EBV-bacmids. BZKO (i–iii, vii–ix, xiii–xv, xix–xxi) and 2089 (iv–vi, x–xii, xvi–xstained for ZEBRA and EA-D (i–xxiv) or Rta and EA-D (xxv–xxxvi). (i–vi) S186A Rta-mutanZEBRA; (xix–xxiv, xxxi–xxxvi) (K188R) WT-like mutant ZEBRA. Yellow arrow indicates a nu

Results

Localization of ZEBRA, Rta and EA-D proteins in lytically induced 293cells containing wild-type or BZKO mutant EBV-bacmids

Two previous studies of viral protein localization during the EBVlytic phase using indirect immunofluorescence to analyze anti-IgG-induced Akata cells (Takagi et al., 1991) or B95-8 cells containing atetracycline-inducible BZLF1 gene (Daikoku et al., 2005) described theaccumulation of ZEBRA, EA-D and Rta proteins in distinct intra-nuclear globular foci. Since these globular structures, usuallynumbering 2 to 5 per cell, co-localized with lytically induced

Rta, activation of BMRF1 expression, and formation of replication compartments in 293viii, xxii–xxxvi) cells were transfectedwithmutant ZEBRA constructs, fixed, then doublet ZEBRA; (vii–xii) R179A early-mutant ZEBRA; (xiii–xviii, xxv–xxx) Y180E late-mutantclear sub-region of increased ZEBRA staining.

Page 5: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

Fig. 3 (continued).

149R. Park et al. / Virology 382 (2008) 145–162

accumulations of EBV DNA as seen by FISH and with sites of viral DNAsynthesis as seen by BrdU staining, they have been termed replicationcompartments. In contrast, studies involving ZEBRA-transfected D98/HR1 cells described diffuse intra-nuclear staining of ZEBRA with noevidence of accumulation at replication compartments. (Bell et al.,2000; Adamson and Kenney, 2001).

To study the effect of mutations in the basic domain of ZEBRA uponthe localization of ZEBRA, and recruitment of ZEBRA and other lyticEBV proteins to replication compartments, we used a humanembryonic kidney cell line, 293, harboring EBV-bacmids. The bacmidscontained an EBV genome in which the BZLF1 gene was inactivated(“BZKO”) or the complete EBV genome (“2089”) (Delecluse et al.,1998). The EBV lytic cycle can be induced in these cells by transfectionof ZEBRA. Point mutants of amino acids in the basic domain of ZEBRAhave unique phenotypes in these cells (Heston et al., 2006). To verify

the utility of this EBV-bacmid system, we first observed thedistributions of ZEBRA, Rta, and EA-D proteins during the lytic phaseinduced by transfection of full-length wild-type BZLF1 gene fused to aconstitutive CMV promoter (Fig. 1). We found that the distribution ofthese three proteins in BZKO cells was indistinguishable from theirdistribution in 2089 cells (Fig. 1 and data not shown).

Using the 293 cell/EBV-bacmid system, all ZEBRA-positive cellsshowed a fine-granular, relatively smooth textured ZEBRA stainingthat was diffuse throughout the nucleus (Fig. 1: panels iii, vi). Inmost ZEBRA-positive cells, small sub-regions of decreased or absentZEBRA staining were interspersed within the diffuse intra-nuclearstaining (Fig. 1: panels iii, vi; blue arrows). ZEBRA was nevercompletely recruited to replication compartments in 293 cells withEBV-bacmids. However, in most cells which showed mature globularreplication compartment formation, as visualized by EA-D staining

Page 6: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

Fig. 3 (continued).

150 R. Park et al. / Virology 382 (2008) 145–162

(Fig. 1: panel i), the diffuse ZEBRA staining was slightly butunmistakably increased in areas coincident with the globularreplication compartments (Fig. 1: panel iii; yellow arrow). Thus,ZEBRA expressed after transfection was diffusely present in thenucleus and partially recruited to viral replication compartments. Ina minority of cells with formation of globular replication compart-ments, however, ZEBRA was not increased at sites containing theglobules with EA-D. (Fig. 1: panel vi.).

The intra-nuclear sub-regions of reduced or absent ZEBRAimmunostaining in ZEBRA-positive cells were roughly circular inshape and resembled the size and outline of nucleoli. To determine ifthese regions correspond to nucleoli, 2089 cells transfected with wild-type ZEBRA were immunostained with antibodies specific for ZEBRAand for the nucleolar proteins, nucleolin and fibrillarin. Doublestaining with anti-ZEBRA and anti-nucleolin antibodies (Fig. 1: panels

vii-ix) showed that nucleolin localized almost exclusively to theregions spared of ZEBRA staining. Double staining for ZEBRA andfibrillarin (Fig. 1: panels x–xii) also showed fibrillarin was found inregions where ZEBRA staining was absent. Wild-type ZEBRA thereforelocalizes diffusely throughout the intra-nuclear space, except that it isexcluded from nucleoli.

In B95-8 cells, Rta has been reported to distribute diffuselythroughout the nucleus at early stages of replication, and to localizeprogressively in globular replication compartments, togetherwithEA-D,at later stages of replication (Daikoku et al., 2005). In 2089 cells, studied42 h after transfection of wild-type ZEBRA, Rta exhibited a pattern oflocalization identical to that of EA-D. Dual staining of EA-D and Rtashowed co-localization of both proteins in discrete globules (Fig. 1:panels xiii–xvi); in other cells (not shown) both Rta and EA-D werediffusely distributed or present in non-uniform furrows.

Page 7: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

Fig. 4. Phenotype of ZEBRA point mutants in 293 cells containing a wild-type EBV-bacmid. Shown are panels from immunoblots of 2089 cells which had been transfectedwith empty vector (CMV) or vectors containing wild-type ZEBRA (gZ) or ZEBRA pointmutants. The immunoblots were probed with monospecific antibodies to Rta, EA-D,ZEBRA (Z), BFRF3 (FR3), and BLRF2 (LR2). The class of each mutant is defined as E: early,L: late, R: Rta, WT: wild-type.

151R. Park et al. / Virology 382 (2008) 145–162

Patterns of EA-D expression

EA-D, the EBV DNA polymerase processivity factor, has been shownto co-localize precisely with globular replication compartments, andtherefore is a useful marker for sites of EBV DNA replication duringlytic phase (Daikoku et al., 2006;Wang et al. 2005; Liao et al., 2001). Incells doubly stained for EA-D and ZEBRA, all cells which were positivefor EA-D were also positive for ZEBRA; however, EA-D was notdetected in a significant number of BZKO cells which were positive forZEBRA (see Discussion). There were 3 patterns of distribution of EA-Dobserved 42 h after transfection (Fig. 2). In approximately 28% of thosecells which expressed EA-D, the protein was found in discrete intra-nuclear globular sub-domains, resembling the globular replicationcompartments described in earlier reports (Fig. 1: panels i, iv, xiii; Fig.2: panels iii, vi). The intensely immunoreactive EA-D-positive globuleswere often surrounded by a very lightly stained diffuse intra-nuclearEA-D. The texture of EA-D staining, both within globular replicationcompartments and in the surrounding intra-nuclear space wassmooth. In approximately 50% of cells positive for EA-D there was adiffuse, smooth, fine-grained intra-nuclear texture of the proteinwithout distinct globule formation (Fig. 2: panels i, iv). In approxi-mately 22% of cells positive for EA-D the protein was neither diffusenor discretely globular, but was non-uniform and furrowed (Fig. 2:panels ii, v). The different distributions of EA-D staining are likely torepresent different stages of the viral life cycle (see Discussion).

The experiments of Fig. 1 and Fig. 2 demonstrated that localizationof ZEBRA and its capacity to induce replication compartments couldbe studied in 293 cells containing EBV-bacmids, a system particularlysuitable for the study of ZEBRA mutants. Furthermore, several stagesof EA-D distribution could be observed after transfection of ZEBRA,thus allowing a comparison of EA-D distribution following expressionof WT and mutant ZEBRA proteins.

Specific ZEBRA mutations affect localization of ZEBRA, expression andlocalization of EA-D, and formation of replication compartments

Representative ZEBRA mutants from each of four functionalcategories were transfected into BZKO or 2089 cells, containing EBVbacmids, and examined for the localization and morphology of ZEBRAand EA-D proteins (Fig. 3). The mutants which were studied in thisexperiment were: 1) S186A, defective at induction of Rta expression;2) R179A, active at inducing Rta, but defective at inducing EA-D; 3)Y180E, active at inducing EA-D, but defective at activating lytic viralDNA synthesis and late protein expression; and 4) K188R, a mutantindistinguishable from wild-type (Table 1). For each mutant thepatterns of ZEBRA and EA-D localization in 2089 cells with wild-type

Table 1Summary of the Phenotypes of ZEBRA mutants used in this study

Mutant Class Binds ZIIB Expression ofindicated protein

ZEBRAdistribution

Replicationcompartments

Rta EA-D FR3/LR2

R179A E + + − − Speckles −Y180E L + + + − Speckles −K181A R W W − − Speckles −A185V R W W − − Speckles −S186A R + − − − Diffuse −S186E DB − − − − Diffuse −K188A L + + + − Speckles −K188R WT + + + + Diffuse +

E=early; fails to activate EA-D.L=late; fails to activate DNA replication or late genes.R=Rta; absent or weak induction of Rta.DB=DNA-binding defect.WT=wild–type.W=weak binding or activation.

EBV genomes were identical to that seen in BZKO cells with BZLF1-inactivated genomes.

S186A is a well-studied ZEBRA mutant which is defective atdisruption of latency due to an inability to activate expression of Rta(Francis et al., 1997). Co-expression of Rta rescues the ability to activateearly gene expression (Adamson and Kenney,1998; Francis et al.,1999).The intra-nuclear distribution of S186A mutant ZEBRA in BZKO and2089 cells was similar to wild-type ZEBRA (Fig. 3: panels iii, vi). Thetransfected S186A ZEBRA protein was confined to the nucleus in afinely-grained, relatively smooth, diffuse pattern. Sub-regions ofreduced or no mutant S186A ZEBRA staining were also observed.However, unlikewild-type ZEBRA, the S186Amutant proteinwas neverobserved in globular areas of increased ZEBRA staining. Therefore,S186A does not lead to formation of replication compartments. Cellsexpressing Z(S186A) never contained EA-D (Fig. 3: panels i, ii, iv, v), afinding consistent with previous analyses by immunoblotting (Franciset al., 1997; Adamson and Kenney, 1998; Heston et al., 2006).

In BZKO and 2089 cells the appearance of the ZEBRA early mutant,R179A, was dramatically different than wild-type and S186A mutantZEBRA. Instead of the diffuse distribution of wild-type and S186AZEBRA, the R179A mutant localized to numerous small discretepunctate foci within the nuclei (Fig. 3: panels ix, xii) of many cells.These discrete foci were uniform in size and were, in general, evenlyspaced throughout the nucleus. Sub-regions with little or no ZEBRAstaining, consistent with nucleoli and similar to those observed withWT and S186A, were often seen. No EA-D protein was seen in cellsexpressing the R179A mutant (Fig. 3: panels vii, viii, x, xi), inaccordance with previous immunoblot studies (Heston et al., 2006;Fig. 4). On a single cell level, transfection with wild-type ZEBRA andthe ZEBRA mutants resulted in a broad range of expression levels ofZEBRA protein in individual cells. Some cells expressed relatively highlevels of ZEBRA protein and others expressed lower levels. Yetregardless of the level of wild-type ZEBRA protein expressed in anygiven cell, the discrete foci seen with R179A was not seen in cellstransfected with wild-type ZEBRA. This result shows that the speckledmutant phenotype is not caused by reduced or increased levels ofprotein. Thus, the arginine at position 179, in addition to beingrequired for transcriptional activation of some early and late genes, isalso essential for proper localization of ZEBRA. Mutation of R179 toalanine dramatically alters the distribution of the protein within thenucleus.

Page 8: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

152 R. Park et al. / Virology 382 (2008) 145–162

When transfected into BZKO and 2089 cells the late ZEBRA mutantY180E also localized to numerous discrete punctate foci (Fig. 3: panelsxv, xviii) in a significant proportion of ZEBRA-positive cells, producinga ZEBRA distribution remarkably similar to the pattern observed with

Fig. 5. Comparison of intra-nuclear distributions ofwild-type andmutant ZEBRAproteins in 29cells (iii, iv, vii, viii, xi, xii, xv, xvi, xix, xx) were transfected with expression plasmids for wild-t(xvii–xx). Cells were fixed and stained 42 h after transfection. 2089 cells were double stained findicates a discrete focus of mutant Y180E ZEBRA in a 293 cell.

mutant R179A. The localization of EA-D following lytic cycle activationby Y180E differed from that induced by wild-type ZEBRA protein.Consistent with previous immunoblot experiments, Y180E ZEBRAinduced high levels of EA-D expression (Heston et al., 2006; Fig 4.).

3 cells with andwithout EBV-bacmids. 2089 cells (i, ii, v, vi, ix, x, xiii, xiv, xvii, xviii) and 293ype ZEBRA (i–iv), R179A (v–viii), Y180E (ix–xii), S186A (xiii–xvi), or K188R mutant ZEBRAor ZEBRA and EA-D, and 293 cells were double stained for ZEBRA and lamin B. Red arrow

Page 9: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

153R. Park et al. / Virology 382 (2008) 145–162

However, unlike EA-D staining induced by wild-type ZEBRA, EA-Dinduced by Y180E was distributed diffusely throughout the nucleus innearly all cells positive for EA-D (Fig. 3: panels xiii, xiv, xvi, xvii). In2089 cells transfected with Y180E, a small percentage, approximately3% (4 of 142 cells selected at random), of cells positive for EA-Dshowedmature viral replication compartments, compared to approxi-mately 28% (40 of 145 cells) of 2089 cells transfected with wild-typeZEBRA. Mutation of the tyrosine at position 180 to glutamatetherefore, allows EA-D expression but prevents the formation ofreplication compartments. Since the Y180Emutant is defective at lyticviral DNA replication, these findings provide further evidence thatglobules containing EA-D are sites of accumulation of lyticallyreplicated EBV DNA.

The WT mutant, K188R, activates Rta expression, early geneexpression, and late gene expression at levels similar to wild-typeZEBRA (Heston et al., 2006; Fig. 4). Immunofluorescence analysis ofBZKO and 2089 cells transfected with the K188R mutant showeddistributions of ZEBRA and EA-D proteinwhichwere indistinguishablefrom cells transfected with wild-type ZEBRA. ZEBRA staining inK188R-transfected cells was diffuse throughout the nucleus with arelatively smooth texture. Small spared sub-regions where ZEBRAstaining was reduced or absent were interspersed within the diffuseintra-nuclear staining (Fig. 3: panels xxi, xxiv;). As seen in wild-typeZEBRA-transfected cells, the diffuse K188R ZEBRA staining was oftenslightly, yet unmistakably, increased in areas co-localizing withglobular replication compartments. Thus, like wild-type ZEBRA,K188R was also partially recruited to viral replication compartments(Fig. 3: panel xxi; yellow arrow). EA-D staining in K188R-transfectedcells was also indistinguishable from wild-type ZEBRA induced cells.In a significant proportion of EA-D positive cells, EA-D was present indiscrete globular replication compartments (Fig. 3: panels xix, xx, xxii,xxiii).

In 2089 cells transfected with wild-type ZEBRA, Rta localized tomature globular replication compartments very similarly to EA-D (Fig.1: panels xiii, xv, xvi). To determine the effects of the Y180E and K188RZEBRA mutants on Rta distribution, 2089 cells transfected with Y180EZEBRA or K188R ZEBRAwere observed for Rta expression at 42 h post-transfection. Transfection with Y180E ZEBRA caused Rta to belocalized either diffusely and uniformly (Fig. 3: panels xxv–xxvii) ordiffusely but non-uniformly (Fig. 3: panels xxviii–xxx), as was seenwith EA-D. A small percentage (2.8%: 4 of 142 cells selected at random)of cells positive for Rta expression did show localization to globularcompartments; however, this was markedly reduced from the 28% ofcells showing Rta in globular compartments in cells transfected withwild-type ZEBRA. The few cells transfected with Y180E which showedRta in globules were also the only ones to show EA-D in globularcompartments. The Y180E mutation therefore allows Rta expression;however, Rta is not found in globular replication compartments, as theresult of Y180E's failure to stimulate lytic DNA replication. Thedistribution of Rta in 2089 cells transfected with K188R ZEBRAmutantwas the same as the distribution of Rta in cells transfected with wild-type ZEBRA (Fig. 3: panels xxxi–xxxvi).

Phenotype of each mutant as assessed by immunoblotting

The immunofluorescence analyses illustrated in Fig. 3 showed thatthe intra-nuclear distribution of each ZEBRA mutant in 2089 cells,containing a wild-type BZLF1 gene within the EBV-bacmid, wasidentical to that seen in BZKO cells containing a bacmid in which theBZLF1 gene was insertionally inactivated. The identity of the two EBV-bacmids in the two cell lines was confirmed by Southern blotting witha probe of BamHI Z. The BZKO cells contained the diagnostic 3 kbBamH1 fragment representing the insertionally inactivated BZLF1gene and 2089 cells contained a wild-type BamH1 Z fragment (datanot shown). To verify that the mutant phenotypes observed in 2089cells were consistent with previous characterization in BZKO cells,

each ZEBRAmutant was assayed by immunoblotting for its capacity toactivate expression of downstream EBV lytic proteins, Rta, EA-D,BFRF3, and BLRF2, in 2089 cells. Each ZEBRA mutant induced thepattern of viral proteins expected for its kinetic class (Fig. 4). Probingwith antibody to ZEBRA showed that equivalent levels of proteinwerebeing expressed by the wild-type and mutant ZEBRA genes (Fig. 4).The maintenance of the mutant ZEBRA phenotype in 2089 cellsstrongly suggests that the transfected mutant ZEBRA proteins did notauto-stimulate the endogenous wild-type BZLF1 gene in these cells(see Discussion).

Intra-nuclear distribution of ZEBRA basic domain mutants in 293 cellslacking an EBV-bacmid

To observe the intracellular distribution of the ZEBRA mutants incells lacking an EBV genome, andmore specifically to test whether theformation of punctate ZEBRA foci observed with the R179A and Y180Emutations required additional viral factors, each ZEBRA mutant wastransfected into 293 cells devoid of an EBV-bacmid. In 293 cells wild-type ZEBRA was observed to be in a diffuse fine-grain intra-nucleardistribution (Fig. 5: panel iv) that was similar to the distribution ofwild-type ZEBRA seen in 2089 cells (Fig. 5, panel ii). However, in 293cells ZEBRA was not present in globular replication compartments asseen in some 2089 cells. Transfection of S186A mutant into 293 cellsalso produced a diffuse and smoothly textured intra-nuclear distribu-tion (Fig. 5, panel xvi) which appeared to be identical to thedistribution seen in 2089 (Fig. 5, panel xiv) or BZKO cells, and wasindistinguishable from wild-type ZEBRA. Transfection of K188Rmutant into 293 cells, similarly to wild-type ZEBRA, produced adiffuse fine-grain intra-nuclear distribution (Fig. 5: panel xx) whichwas very similar to the distribution seen in 2089 cells (Fig. 5, panelxviii). Thus, the intra-nuclear distribution of wild-type, S186A, andK188R mutant ZEBRA proteins was indistinguishable in EBV-positiveand EBV-negative 293 cells.

The distribution of R179A and Y180Emutant ZEBRA proteins in 293cells differed from the smooth diffuse distribution of wild-type ZEBRAand S186A in the same cells, and also differed from the discretelyspeckled localization of these two mutants in cells containing EBV-bacmids. In the nuclei of 293 cells, R179A and Y180E mutant ZEBRAproteins were distributed non-uniformly. The localization patternsranged from partially diffuse, coarsely textured staining (Fig. 5, panelviii) to a non-diffuse and uneven, lumpy distribution (Fig. 5, panel xii),in which a few discrete punctate foci were seen (Fig. 5, panel xii: redarrow). The non-uniform and uneven distribution of R179A and Y180Eproteins observed in 293 cells could be clearly distinguished from thepattern of distinctly-defined and evenly spaced discrete punctate fociseen in 2089 and BZKO cells containing EBV-bacmids (Fig. 5, panels viand x). The differences in distribution of R179A or Y180E mutantZEBRA proteins in 293 cell backgrounds with and without an EBVgenome suggests two possibilities which are not mutually exclusive:(1) the mutation in the basic domain itself may affect intra-nuclearlocalization of ZEBRA protein, even in EBV-negative cells; or (2)cellular and viral factors may partially contribute to the mutantspeckled phenotype seen in cells containing EBV-bacmids. Theinability of either R179A or Y180E mutant to produce in 293 cellsthe numerous clearly distinct evenly distributed punctate foci whichwere seen in 2089 or BZKO cells suggests that additional viral factorsconferred by lytic EBV infection are likely to be necessary to producethe speckled phenotype (see Discussion).

Effect of viral DNA replication inhibitors, phosphonoacetic acid (PAA)and acyclovir (ACV), on localization of ZEBRA and EA-D in cellstransfected with wild-type or R179A mutant ZEBRA

The next experiments attempted to determine whether thespeckled distribution of ZEBRA, observed with a mutant, such as

Page 10: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

154 R. Park et al. / Virology 382 (2008) 145–162

Page 11: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

155R. Park et al. / Virology 382 (2008) 145–162

R179A, that does not induce viral replication, could be reproducedwith wild-type ZEBRA when lytic viral DNA replication was inhibitedby anti-viral drugs. During HSV-1 lytic infection the speckled pre-replication compartment stage is transient (Quinlan et al., 1984), butmay be visualized by blocking progression of the lytic phase with viralDNA replication mutants or by treatment with viral DNA replicationinhibitors such as phosphonoacetic acid (PAA) or acyclovir (ACV)(Lukonis and Weller, 1996). To observe the effect of inhibiting viralDNA synthesis on the localization of ZEBRA and EA-D proteins, both ofwhich are presumptive components of an EBV pre-replicationcomplex, 2089 cells were transfected with WT or R179A mutantZEBRA and treated with PAA or ACV (Fig. 6). In cells transfected withWT ZEBRA and treated with PAA or ACV, ZEBRA staining remaineddiffuse throughout the nucleus. ZEBRA was never seen in a speckleddistribution in the presence of the inhibitors of viral DNA replication.Moreover, globular regions of increased ZEBRA staining were neverdetected in cells treated with PAA or ACV.

The localization of EA-D in cells transfected with wild-type ZEBRAand treated with PAA or ACV differed from the distribution of EA-D inuntreated cells. Whenever EA-D was seen in PAA/ACV-treated cells itwas diffusely present in the nucleus and never recruited to globularstructures (Fig. 6: panels iv, vii). Thus PAA or ACV blocked theappearance of mature replication compartments which require lyticviral DNA replication for their formation. As described in Fig. 2, 50% of2089 cells transfected withWT ZEBRA, but not treated with inhibitorsof viral replication, contained diffuse intra-nuclear EA-D which wasnot found in globular replication compartments. However, the diffuseEA-D staining in PAA/ACV-treated cells differed from the smoothlytextured and relatively homogeneous diffuse staining seen inuntreated cells. PAA/ACV treatment caused EA-D staining to be non-uniform and to consist of numerous clumped aggregations or foci (Fig.6, iv and vii). These foci of EA-D clumps produced in the presence ofPAA and ACV were uneven in size, intensity, and spacing and did notresemble the uniformly sized and evenly spaced punctate ZEBRA fociseen with some ZEBRA mutants, such as R179A and Y180E.

PAA and ACV treatment did not alter the formation of discretepunctate ZEBRA foci by the R179A ZEBRA mutant. The speckledphenotype of R179A mutant ZEBRA in PAA/ACV-treated cells (Fig. 6:panels xv, xviii) was indistinguishable from that observed in untreatedcells (Fig. 6: panel xii). These results show that viral DNA replication isnot required for formation of the speckled ZEBRA foci by the R179Amutant. Moreover, since R179A does not activate BMRF1, the resultsshow that discrete punctate foci of ZEBRA are generated indepen-dently of the presence of the viral polymerase processivity factor.

Intra-nuclear distribution of S186A mutant ZEBRA is not affected byexogenous expression of Rta

The diffuse distribution of ZEBRA S186A mutant in 2089 cellsdiffers from the punctate distribution of ZEBRA induced by the R179Aand Y180E mutants. One possible reason for the different distributionpattern of S186A, as compared with the R179A and Y180E mutants,may lie in the inability of S186A to activate expression of Rta protein.Therefore we further tested whether the diffuse distribution of theS186A ZEBRA mutant was altered when Rta was exogenously co-expressed with S186A mutant ZEBRA. Previous experiments haveshown that exogenous expression of Rta rescues BMRF1 transcrip-tional activation by the S186A ZEBRA mutant (Francis et al., 1999;Adamson and Kenney, 1998; Heston et al., 2006). 2089 cells co-transfected with S186A and Rta-expression plasmids were stained forZEBRA and EA-D, as a marker for Rta activity. Cells co-transfected withS186A and Rta showed bright EA-D staining, indicating that Rta was

Fig. 6. Effect of phosphonoacetic acid (PAA) and acyclovir (ACV) treatment on the localcompartments. 2089 cells were transfected with wild-type ZEBRA (i–ix) or R179A mutant ZEthen fixed and double stained for ZEBRA and EA-D.

active (Fig. 7: panels iv, v). However in all cells positive for EA-D, theprotein was uniformly and diffusely distributed throughout the intra-nuclear space, and not in globular replication compartments.Exogenous expression of Rta did not affect the distribution of S186Amutant ZEBRA which remained diffusely and smoothly distributedthroughout the intra-nuclear space (Fig. 7, panel vi). These resultsshowed that the diffuse distribution of Z(S186A) is an intrinsicproperty of the mutant protein and is not correlated with its failure toactivate Rta or EA-D expression.

To test whether the diffuse distribution of the S186A mutantZEBRAwas common to other ZEBRA mutants deficient at activation ofRta and EA-D, we studied the distribution of two additional mutants,K181A and A185V. Both mutants bind ZIIIB DNA relatively weakly;K181A binds 37% and A185V 10% as actively as wild-type (Heston et al.,2006). K181A activates low levels of Rta and A185V activates onlytrace levels of the protein. Transfection of K181A (Fig. 7, panel ix) orA185V (Fig. 7, panel xii) into 2089 cells resulted in a discrete punctatedistribution of ZEBRA similar to that seen with the R179A and Y180Emutants. The localization of S186A mutant ZEBRA therefore differedfrom other Rta-deficient ZEBRA mutants.

We also compared the distribution of the S186A mutant with thedistribution of another mutant at the same position, S186E. Bothmutations completely abolish ZEBRA's capacity to disrupt latency;however, the S186A mutant retains its capacity to bind ZIIIB DNA,whereas S186E does not bind ZIIIB DNA (Heston et al., 2006). Thedistribution of S186E ZEBRA appeared similar to S186A (Fig. 7, panelxv). S186E was smoothly and diffusely distributed throughout theintra-nuclear space; no discrete puncta were seen in any cell; and aswas seen with S186A, sub-regions of reduced S186E staining werepresent. These results make it unlikely that the diffuse distribution ofS186A is directly related to its capacity to bind DNA.

Intra-nuclear distribution of wild-type and R179A mutant ZEBRA inD98/HR-1 cells

The speckled distribution of ZEBRA mutants, such as R179A andY180E, in 293 cells containing EBV-bacmids is a new phenomenon.Therefore we wished to investigate whether speckle formationspecific to certain ZEBRA mutants could be observed in other cellbackgrounds. To determine whether WT or R179A mutant ZEBRAwould form discrete speckled punctate foci in other cell lines infectedwith EBV, we examined ZEBRA and EA-D localization in D98/HR-1cells transfected with WT ZEBRA or R179A mutant ZEBRA (Fig. 8). Thedistribution of WT ZEBRA and EA-D in D98/HR-1 cells was similar tothat observed in 2089 and BZKO cells (Fig. 1). ZEBRA staining wasfinely grained, relatively smooth in texture, and distributed diffuselythroughout the nucleus (Fig. 8: panels iii, vi). ZEBRA was notcompletely recruited to replication compartments in ZEBRA-trans-fected D98/HR-1 cells. However, as was seen in BZKO and 2089 cells, inmany D98/HR-1 cells which showed mature globular replicationcompartments, as visualized by EA-D staining (Fig. 8: panel i), thediffuse ZEBRA staining was slightly but unmistakably increased inareas coincident with the globular replication compartments (Fig. 8:panel iii; yellow arrow). Thus, ZEBRA was partially recruited to viralreplication compartments. However, in some D98/HR-1 cells contain-ing globular replication compartments, as in the 293 cells with EBV-bacmids, ZEBRA was not increased at the site of EA-D globules (datanot shown).

Three patterns of EA-D staining were seen in D98/HR-1 cellsanalyzed 42 h after transfectionwithWT ZEBRA. These resembled thepatterns in BZKO and 2089 cells (Fig. 2). In some cells intra-nuclearEA-D was diffusely distributed (Fig. 8: panels iv, v); in other cells EA-D

ization of wild-type and R179A mutant ZEBRA and on the formation of replicationBRA (x–xviii), treated with 500 μM PAA (iv–vi, xiii–xv) or 50 μM ACV (vii–ix, xvi–xviii),

Page 12: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

Fig. 7. Intra-nuclear distribution of ZEBRAmutantswhich are defective at activation of Rta andEA-D. 2089 cellswere transfectedwith S186Amutant ZEBRA (i–iii), S186Aplus Rta (iv–vi),K181A mutant ZEBRA (vii–ix), A185V mutant ZEBRA (x–xii), or S186E mutant ZEBRA (xiii–xv). Cells were fixed and double stained for ZEBRA and EA-D.

156 R. Park et al. / Virology 382 (2008) 145–162

was found in discrete globular replication compartments (Fig. 8:panels i, ii); and other cells contained unevenly distributed EA-D (notshown).

Fig. 8. Nuclear localization, activation of EA-D, and replication compartment formation by wwith wild-type ZEBRA (i–vi) or R179A mutant ZEBRA (vii–xviii). Cells were fixed and doublZEBRA staining.

D98/HR-1 cells transfected with R179A ZEBRA showed a spectrumof nuclear morphological characteristics. One cellular subpopulationstained positively for ZEBRA and were negative for EA-D. In many of

ild-type and R179A mutant ZEBRA in D98/HR-1 cells. D98/HR-1 cells were transfectede stained for ZEBRA and EA-D. Yellow arrows indicate nuclear sub-regions of increased

Page 13: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

157R. Park et al. / Virology 382 (2008) 145–162

Page 14: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

158 R. Park et al. / Virology 382 (2008) 145–162

these cells, ZEBRA was present in numerous small discrete punctatefoci in the intra-nuclear space (Fig. 8: panel ix). In this subpopulationof D98/HR-1 cells the phenotype of R179A mutant ZEBRA resembledthe phenotype seen in BZKO and 2089 cells (see Fig. 3, panels ix, xii).Another subpopulation of D98/HR-1 cells transfected with R179AZEBRA stained positively for EA-D protein and for ZEBRA (Fig. 8:panels x, xiii, xvi). This population was never observed in BZKO or2089 cells. Some of the EA-D-positive D98/HR-1 cells observed aftertransfection of R179A displayed a wild-type phenotype: EA-D waslocalized to globular replication compartments (Fig. 8: panels x, xi);ZEBRA was distributed diffusely throughout the intra-nuclear space(Fig. 8: panel xii) and regions of increased ZEBRA staining co-localizedwith replication compartments (Fig. 8: panel xii, yellow arrow). Inanother pattern, also similar to a pattern observed with WT ZEBRA in293 cells with the EBV-bacmids, both EA-D and ZEBRA weredistributed diffusely and uniformly throughout the nucleus (Fig. 8:panels xiii–xv; compare Fig. 1). Because the R179A mutant itself isincapable of activating BMRF1 expression in cells containing EBV-bacmids, the presence of BMRF1-positive cells following transfectionof R179A ZEBRA into D98/HR-1 cells suggests that the R179A mutantauto-stimulates the endogenous wild-type BZLF1 gene in this cellbackground. The two patterns of EA-D distribution, globular anddiffuse, can be attributed to auto-activation of the endogenous BZLF1gene (see Discussion).

A novel pattern of EA-D and ZEBRA distribution, never observed inthe 293 cell/EBV-bacmid system, was also seen in some EA-D-positiveD98/HR-1 cells after transfection of the R179A mutant. In these cellsEA-D was localized to many heavily staining globular replicationcompartments amidst a more lightly stained diffuse smoothlytextured surrounding intra-nuclear space (Fig. 8: panel xvi). In thesame cells, ZEBRA was also heavily concentrated in replicationcompartments, amidst a lighter stained surrounding intra-nuclearspace (Fig. 8: panel xviii). However, in replication compartments andin the lightly stained surrounding space ZEBRA had a definite speckledpattern. Cells of this type seemed to exhibit a mixture of both thediscrete punctate R179A mutant phenotype and the wild-type ZEBRAphenotype.

Discussion

As the major EBV lytic regulatory protein, ZEBRA functions withinthe nucleus as a transcriptional activator and as a lytic DNA replicationorigin-binding protein. In our previous study, a series of pointmutations introduced into the basic DNA-recognition domain ofZEBRA were characterized with respect to the resulting defects intranscriptional activation and DNA replication (Heston et al., 2006). Inthis study we further investigated the effects of some of thesemutations on the intra-nuclear distribution of ZEBRA protein, and onthe formation of replication compartments. We found that specificamino acid changes within the basic domain are associated withspecific changes in ZEBRA localization, and also influence theformation of mature replication compartments. The mutants can begrouped into four classes based on these morphological changes(Table 1). Several amino acid substitutions altered the localization ofZEBRA from its typical diffuse distribution into a speckled phenotypecontaining numerous, small, discrete, evenly spaced and evenly sizedfoci. The speckled phenotype of one group of mutants (i.e. K181A,A185V, and R179A) was accompanied by a defect in ZEBRA's capacityto activate Rta or EA-D; whereas other mutants with a speckleddistribution (i.e. Y180E and K188A) induced expression of EA-Dwhich,however, was not recruited to globular replication compartments. Athird distinct class of ZEBRA mutations at position S186 was diffuselydistributed in the nucleus in a pattern reminiscent of wild-typeprotein; however, S186A and S186E did not induce Rta or downstreamviral genes, and did not form replication compartments. A fourth classof mutation (i.e. K188R), found in our earlier study to have no defects

in transcriptional activation or DNA replication (Heston et al., 2006),had a normal diffuse intra-nuclear distribution of ZEBRA andpromoted globular replication compartment formation. Thus, thisgroup of ZEBRA mutants not only arrest the EBV lytic cycle at distinctstages, they are distributed in distinct patterns in the nucleus,suggesting that they induce distinct phases of nuclear re-organization.

Cell background considerations in analysis of the intra-nucleardistribution of ZEBRA mutants and formation ofreplication compartments

We found that the 293 cell/EBV-bacmid system is particularly wellsuited for analysis of the intracellular distribution of mutant ZEBRAproteins. Since the BZLF1 gene is disrupted in the BZKO cell line, thereis no endogenous ZEBRA expressed and all the ZEBRA protein is madefrom the transfected plasmid. However, some BZKO cells whichexpressed wild-type ZEBRA did not express EA-D. The failure of ZEBRAexpression to be accompanied by signs of EBV lytic activation is likelyto be due to loss of the EBV-bacmid from some BZKO cells. The EBV-bacmid appears to be more stably maintained in 2089 cells.Remarkably, all the mutant ZEBRA phenotypes which were observedin BZKO cells were preserved in 2089 cells, which was shown tocontain an intact BZLF1 gene (Figs. 3 and 4, and data not shown). It islikely that none of the ZEBRA mutants directly auto-stimulated theendogenous BZLF1 gene in 2089 cells; however it is also possible thatoverexpression of the transfected mutant ZEBRA proteins may havehad a dominant-negative effect or have masked wild-type ZEBRAprotein expressed from the endogenous BZLF1 gene. Since many ofthese mutants also activated Rta expression, Rta did not cross-activateBZLF1 in 2089 cells, as it does in some cell backgrounds.

The intra-nuclear distribution of ZEBRA mutants was also studiedin 293 cells that did not contain EBV-bacmids. These experiments (Fig.5) showed that nuclear localization of ZEBRA and all the mutants wasindependent of the presence of EBV gene products. The appearance ofwild-type, S186A, and K188R mutants was identical in 293 cells withandwithout EBV. However, the discrete speckled appearance of R179Aand Y180E was only observed in EBV-positive 293 cells. Thesemutants, however, were distributed unevenly in the nuclei of 293cells. This experiment suggests that in some instances the mutationsthemselves may affect intra-nuclear location, independent of anyeffects the mutations may have on the EBV life cycle.

Parallels between HSV-1 and EBV in formation of the viral DNAreplication complex

Parallels have been demonstrated between EBV and HSV-1 in thecomposition of and the process of formation of the core viral replicationcomplex. As a DNA replication protein, ZEBRA binds to oriLyt andrecruits other viral replication proteins to oriLyt. The assembly of thiscore replication complex at oriLyt allows initiation of viral DNAsynthesis. Formation of the replication complex relies on manydifferent contact points between ZEBRA and the individual corereplication proteins. The EBV helicase (BBLF4), primase (BSLF1), andprimase-associated factor (BBLF2/3) directly bind each other to form atripartite complex that interactswith the viral polymerase (BALF5) (Gaoet al., 1998; Yokoyama et al., 1999; Fujii et al., 2000). The primase sub-complex (BSLF1–BBLF2/3) interactswith theN-terminus of ZEBRA (Gaoet al., 1998); the EBV helicase (BBLF4) interacts with an adjacent butseparate region of ZEBRA (Liao et al., 2001). EA-D (the polymeraseprocessivity factor) interacts with both BALF5 and ZEBRA (Kiehl andDorsky,1991; Li et al.,1987; Zhang et al.,1996), and the single-strandedDNA-binding protein (BALF2) interacts with the EA-D–BALF5 complex(Zeng et al., 1997). Similar interactions among the homologous corereplication proteins of HSV-1 have been demonstrated (Marsden et al.,1997; McLean et al., 1994; Hamatake et al., 1997). The origin-bindingfunction of ZEBRA is assumed by the HSV-1 UL9 proteinwhich contacts

Page 15: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

159R. Park et al. / Virology 382 (2008) 145–162

the primase-associated factor, UL8 (McLean et al., 1994). All of the EBV-encoded replication proteins would be expected to be present in theglobular replication compartments. In our experiments, thus far, wefind ZEBRA (BZLF1), EA-D (BMRF1) and Rta (BRLF1) in replicationcompartments (Figs.1, 3).We do not yet understand the role of EBV-Rtain replication.

Effects of inhibiting HSV-1 and EBV DNA synthesis on the formation ofviral pre-replication sites and mature replication compartments

When HSV-1 lytic replication is blocked, either by treatment withviral DNA synthesis inhibitors such as PAA or ACV, or by infection withviral DNA replication-defective mutants, mature replication compart-ments donot form. Instead, the core viral replicationproteins localize tonumerous discrete punctate foci that assume a speckled intra-nucleardistribution (Liptak et al., 1996). These punctate foci, which are often inclose association with ND10 or PML bodies, have been termed pre-replicative sites, and represent the precursors to mature globular viralreplication compartments. Time lapse microscopy shows that thenumerous HSV-1 pre-replication foci actively coalesce and accrue toform larger mature globular replication compartments (Taylor et al.,2003). Transfection of four HSV-1 replication proteins (the helicase,primase, primase-associated factor, and polymerase processivityfactor) is sufficient to form pre-replicative sites even in uninfectedcells (Uprichard and Knipe, 1997); however, in virally infected cells theorigin-binding protein is also required (Lukonis et al., 1997).

In our studies, in parallel with those in HSV-1, the formation ofmature replication compartments was blocked in 293/EBV bacmidcells transfected with wild-type ZEBRA and treated with inhibitors ofviral DNA replication (Fig. 6). However, PAA or ACV treatment of 2089cells induced into the lytic cycle by WT ZEBRA did not reproduce, inour experiments, the speckled appearance of HSV-1 pre-replicationsites or the speckled distribution of some ZEBRA mutants (Fig. 6,panels vi, ix). Instead, a non-uniform clumped intra-nuclear distribu-tion of EA-D was seen in 2089 cells that expressed ZEBRA and weretreated with PAA or ACV (Fig. 6, panels iv, vii). The non-uniform andclumped distribution of EA-D in cells treated with inhibitors of lyticreplication suggests that discrete pre-replication foci were presentand were in the process of coalescing to form mature replicationcompartments. The absence of punctate ZEBRA foci may indicate thatZEBRA's association with the replication complex, or with oriLyt, istransient. ZEBRA, which might act as a tether for the replicationcomplex by virtue of its high affinity for oriLyt DNA, would beexpected to dissociate rapidly from the fully assembled replicationcomplex, perhaps immediately following recruitment of the viral DNApolymerase, in order to allow movement of the complex along thetemplate DNA. The inhibitory effects of PAA and ACV would exertthemselves after the replication complex is fully assembled; thereafterZEBRA would dissociate from the DNA and would be diffuselydistributed. The speckle-forming mutants of ZEBRA, however, mayblock progression of assembly of the core replication complex at astage before the dissociation of ZEBRA from the DNA. In support of thishypothesis, the mutant R179A remains in a speckled distribution in2089 cells treated with PAA or ACV (Fig. 6, panels xv, xviii). Thisfinding suggests that the R179A speckles are formed prior to theevents which are inhibited by PAA and ACV. The events sensitive tothese inhibitors involve the viral DNA polymerase. Therefore it may beassumed that Z(R179A) does not activate expression or interactproperly with BALF5.

The speckled distribution of ZEBRA mutants may representpre-replication foci

The composition of the speckled foci formed by mutant ZEBRAproteins, and the mechanism underlying their formation is not yetknown. All the mutations with a speckled distribution halt the

progression of the lytic cascade at early stages of viral gene expression,before lytic viral DNA synthesis (Heston et al. 2006). Therefore, itseems possible that the speckled distribution may represent atransient intermediary stage of ZEBRA localization, corresponding toa pre-replication phase for EBV as has been observed for HSV-1 (deBruyn Kops and Knipe, 1988; Quinlan et al., 1984). Several lines ofevidence support the hypothesis that the speckled distribution ofsome ZEBRA mutants represents EBV pre-replication foci. Thespeckled distribution of ZEBRA mutant Y180E was accompanied bytwo other deficiencies which point to a defect during lytic DNAreplication in the assembly or function of the viral core replicationcomplex: (i) 2089 cells transfected with Y180E were incapable ofdevelopingmature globular replication compartments, despite havingnormal expression levels of the polymerase processivity factor, EA-D(Fig. 3, panels xiii, xvi). (ii) The Y180E mutant lacks the ability tostimulate lytic viral DNA synthesis (Heston et al., 2006.). These twodefects were also characteristic of another late mutant, K188A, whichformed speckles similar to Y180E (data not shown).

A third observation which supports the hypothesis that specklesformed by mutant ZEBRA proteins represent pre-replication sites, asopposed to aberrant recruitment of mutant ZEBRA to pre-existingcellular structures, is that the characteristically distinct speckleddistributions of the R179A and Y180E mutations seen in 2089 andBZKO cells were not seen in 293 cells lacking EBV (Fig 5, panels VIII,XII). This observation would indicate that the speckled distribution ofmutant ZEBRA is not an intrinsic effect of the mutation alone, butrather requires the presence of additional viral factors or virallyinduced changes to the host cell. It should also be noted, however, thatthe roughly textured distribution pattern of the R179A and Y180Emutants in 293 cells differed from the smooth, even distribution seenwith wild-type ZEBRA in the same cell type. Thus the speckledphenotype may reflect both intrinsic qualities of the mutant ZEBRAprotein and accompanying alterations in cellular or viral structures.

One difficulty with the hypothesis that the speckled ZEBRAdistribution represents pre-replicative sites is that some ZEBRAmutants which appeared in speckles i.e. R179A, K181A and A185Varrest the EBV lytic cycle at very early stages, after Rta but before EA-Dexpression (Fig. 3; Table 1). If, in amanner similar to HSV-1, four or fiveEBV replication proteins must be expressed in order to form speckledpre-replication sites, these early mutants would be expected to induceexpression of these replicative proteins. If they are found not to induceexpression of these replication proteins, then the formation ofspeckled EBV pre-replication sites may occur by a mechanism thatdiffers from HSV-1. Alternatively, the mutant ZEBRA speckles mayrepresent structures other than pre-replication foci. The pre-replica-tion speckles observed in HSV-1 infection have been defined by ICP8,the single-stranded DNA-binding protein homologue of EBV BALF2,whereas ZEBRA is a functional homologue of HSV-1 UL9, the origin-binding protein.

Hypotheses explaining the diffuse distribution of mutant Z(S186A) andZ(S186E)

The smooth distribution of the S186A and S186E ZEBRAmutants in2089 cells contrasts with the speckled distribution seen with severalother ZEBRA mutants (Fig. 5). This observation may contribute tounderstanding the mechanism by which speckled foci are generated.The difference between the sub-nuclear localization of S186 mutantsand speckle-forming ZEBRA mutants may not be attributable todifferences in binding to DNA. Although the S186Amutant is defectivein binding to ZREs which contain a methylated CpG (Bhende et al.,2005), this mutant retains the ability to bind many other ZRE's as wellas, or better than, wild-type or speckle-formingmutants (Francis et al.,1997; Heston et al., 2006; Adamson and Kenney, 1998). Furthermore,S186E does not bind DNA, yet is also diffusely distributed throughoutthe nucleus. S186A and S186E both differ, however, from other DNA-

Page 16: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

160 R. Park et al. / Virology 382 (2008) 145–162

binding-competent ZEBRA mutants in their inability to activate viralgene expression. This finding suggested that some viral geneexpression is required for formation of the speckled distribution ofZEBRA. Since a major difference between S186A and other DNA-binding ZEBRA mutants is an inability to induce expression of Rta, weexplored the effect of co-transfection of Rta (Fig. 7). Co-transfection ofRta with S186A rescued expression of EA-D; however addition of Rtadid not alter the diffuse distribution of S186A or EA-D (Fig. 7, panels iv,v, vi). Thus expression of Rta and EA-D alone are not sufficient forproduction of the speckled ZEBRA phenotype. It is likely that otherviral or cellular proteins not induced by the combination of S186A andRta are required for production of the speckled phenotype.

It is possible that S186A, by itself or in combination with Rta andEA-D, does not activate expression of one or more members of thetripartite helicase–primase complex necessary for generation of HSV-1 or EBV pre-replication sites (Uprichard and Knipe, 1997). OtherZEBRA mutants, such as R179A and Y180E, may induce expression ofthese proteins to levels sufficient for speckle formation. Experimentsanalyzing the levels of expression of each of the viral replicationproteins induced by these ZEBRA mutants may provide an answer tothis question.

Historical insights into the intra-nuclear re-distribution of the EBV DNApolymerase processivity factor

In classical experiments the Henles described EA-D as “earlyantigen-diffuse” in Raji cells superinfected with the P3J-HR-1strain ofEBV (Henle et al., 1971). We now know that the EBV DNA in Raji cellsdoes not undergo lytic replication of EBV DNA due to a deletion inBALF2, the EBV gene encoding the single-stranded DNA-bindingprotein (Zhang et al., 1988; Decaussin et al., 1995). Therefore, Raji cellsinduced into the lytic cycle express early but not late viral proteins.Moreover, the P3J-HR-1 strain of EBV which induced EA-D in Raji cellscontains “het DNA” from which BZLF1, the ZEBRA gene, is constitu-tively expressed as a result of rearrangement of the viral DNA (Milleret al., 1985; Rooney et al., 1988). Therefore, the ZEBRA protein isresponsible for activating EBV lytic gene expression after super-infection of Raji cells. Furthermore, when introduced into Raji cells onstable oriP plasmids, ZEBRA activates early but not late EBV geneexpression (Gradoville et al., 1990).

We observed EA-D in a diffuse distribution in the nuclei of cellsunder the following experimental conditions: 1) in some BZKO, 2089or D98/HR-1 EA-D positive cells transfected with WT ZEBRA (Figs. 2,8); 2) in all EA-D-positive cells expressing the ZEBRA mutant Y180Ewhich activates early but not late EBV genes; 3) in cells treated withPAA or ACV and transfected withWT ZEBRA (Fig. 6); and 4) in cells co-expressing the S186A mutant and Rta (Fig. 7). This combinationactivates early genes but does not activate late genes (Adamson andKenney, 1998). Thus, as surmised by the Henles, when EA-D isdiffusely distributed throughout the nucleus, it is a marker for theearly stages of viral gene expression, preceding viral replication.

A dramatically different distribution of EA-D in globular replicationcompartments was observed following transfection of WT ZEBRA orthe K188R WT mutant into BZKO, 2089 or D98/HR-1 cells (Figs.1,2,3,5,8). Globular EA-D was also observed in D98/HR-1 cells trans-fected with the R179A early mutant, probably as the result of auto-stimulation of the endogenousWT ZEBRA (Fig. 8). Globular replicationcompartments were not observed under conditions known to blocklate gene expression, including transfection of the late-mutant Y180E,treatment with inhibitors of lytic replication, or co-expression ofS186A and Rta. Thus, the presence of EA-D in globular replicationcompartments is a marker for late phases of the viral life cycle, duringor after lytic DNA replication. At late stages the distribution of EA-D isneither “early” nor “diffuse”. The findings suggest that two differentfunctions for the BMRF1 gene product may be accompanied by twodifferent intra-nuclear distributions. At early times, when it is distri-

buted diffusely in the nucleus, the protein may act as a transcriptionfactor, as previously suggested (Zhang et al., 1996). At late times, whenpresent in globular compartments, it may serve its role as the DNApolymerase processivity factor.

The cellular components of the replication globules are poorlycharacterized. Recently several cellular DNA repair proteins werefound to be recruited to EBV replication compartments (Kudoh et al.,2005; Daikoku et al., 2006); however, the role of DNA repairmechanisms in lytic viral replication is poorly understood. Signalswhich cause re-location of EA-D to these compartments, themechanisms, the motor and the scaffolds for this dramatic processof nuclear re-organization remain to be elucidated. ZEBRA mutantswhich facilitate or inhibit the development of globular replicationcompartments will be effective tools in dissecting the mechanismsinvolved in development of these critical intra-nuclear structures andin elucidating the complex process of EBV lytic DNA replication.

Materials and methods

Cell lines

293 is a human embryonic kidney cell line immortalized by theearly region of adenovirus (Van der Eb et al., 1977). 2089 is a 293cell line stably transfected with a bacmid containing the full-lengthEBV genome and a hygromycin B-resistance gene. BZKO is a 293 cellline containing an EBV-bacmid in which the BZLF1 gene has beeninactivated by insertion of a kanamycin resistance cassette (Feederleet al., 2000). D98/HR1 is a cell line generated by fusion of a humanepithelial cell line, D98, and an EBV-positive Burkitt's lymphoma cellline, P3HR1 (Glaser and O'Neill, 1972). 293 cells were maintained inRPMI 1640 complete media, supplemented with 10% fetal bovineserum, 50 U/mL penicillin–streptomycin, and 1 μg/mL amphotericinB. 2089 cells and BZKO cells were maintained in RPMI 1640complete media containing 100 μg/mL hygromycin B (Calbiochem).D98/HR1 cells were maintained in Dulbecco's modified minimalessential complete media supplemented with MAGGT (0.6 μMmethotrexate, 50 μM adenosine, 50 μM guanosine, 0.1 mM glycine,16 μM thymidine).

Antibodies

In immunofluorescence experiments, ZEBRA was detected with apolyclonal antibody (S1605-1) prepared from rabbits immunized withpurified full-length ZEBRA. ZEBRA protein was expressed in E. coli froma pET22b vector containing the BZLF1 gene and purified over a nickel-agarose column. EA-D was detected using the mouse monoclonalantibody R3.1 (Pearson et al., 1983). In immunoblot experiments,ZEBRA was detected using a rabbit polyclonal antibody describedpreviously (El-Guindy and Miller, 2004). Rta and BLRF2 protein weredetected using rabbit polyclonal antisera described previously (Hestonet al., 2006). BFRF3 was detected with a rabbit polyclonal antibody(S1931-1) raised against full-length BFRF3 protein expressed andpurified from E. coli. Lamin B was detected using a commerciallyavailable goat polyclonal antibody (Santa Cruz; sc-6216). Antibodies tonucleolin (ab13541) and fibrillarin (ab18380) were purchased fromAbcam. The following secondary antibodies, used in immunofluores-cence experiments, were purchased from Jackson ImmunoResearchLaboratories, Inc.: FITC-conjugated sheep anti-mouse IgG (#515-095-062), Texas Red-conjugated donkey anti-rabbit IgG (#711-075-152),FITC-conjugated donkey anti-goat IgG (#705-095-147).

Indirect immunofluorescence

Cells were grown on glass coverslips and transfected with plasmidDNA using DMRIE-C reagent (Invitrogen). After 8 h the transfectionreagent was replaced with growth media. In certain experiments,

Page 17: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

161R. Park et al. / Virology 382 (2008) 145–162

growthmedia contained 500 μMphosphonoacetic acid (PAA) or 50 μMacyclovir (ACV). Forty one hours after transfection, a time previouslydetermined to be adequate for detection of lytic viral DNA replication,cells were fixed in methanol for 30 min. at −20 °C, washed with PBS,and incubated in blocking solution (10% human serum in PBS) for 1 hat room temperature. Cells were stained with primary antibodydiluted in blocking solution for 1 h at room temperature in humidifiedchambers. Cells were washed with PBS, then incubated withsecondary antibody diluted 1:200 in blocking solution for 1 h atroom temperature in opaque humidified chambers. Cells werewashedwith PBS, briefly rinsed in distilled H2O to remove salts, thenmountedon glass slides using Vectashield mounting media (Vector Labora-tories). An Olympus confocal laser scanningmicroscopewith Fluoviewsoftware was used to obtain digital images of fluorescence andtransmitted light. Excited fluorophore emissions were collected eitherindividually, or simultaneously under detection settings whichreduced cross-talk between the green and red channels to non-detectable levels.

Acknowledgments

This study was supported by NIH grants CA12055 and CA16038.We thank S. Rivkees and C.Wendler for use of the confocal microscopyfacility at the Yale Child Health Research Center. We thank Ayman El-Guindy and Jill Countryman for many helpful discussions.

References

Adamson, A.L., Kenney, S.C., 1998. Rescue of the Epstein–Barr virus BZLF1 mutant, Z(S186A), early gene activation defect by the BRLF1 gene product. Virol. 251 (1),187–197.

Adamson, A.L., Kenney, S., 2001. Epstein–Barr virus immediate-early protein BZLF1 isSUMO-1 modified and disrupts promyelocytic leukemia bodies. J. Virol. 75,2388–2399.

Baumann,M., Feederle, R., Kremmer, E., Hammerschmidt,W.,1999. Cellular transcriptionfactors recruit viral replication proteins to activate the Epstein–Barr virus origin oflytic DNA replication, oriLyt. EMBO J. 18 (21), 6095–6105.

Bell, P., Lieberman, P.M., Maul, G.G., 2000. Lytic but not latent replication of Epstein–Barrvirus is associated with PML and induces sequential release of nuclear domain 10proteins. J. Virol. 74, 11800–11810.

Bhende, P.M., Seaman, W.T., Delecluse, H.J., Kenney, S.C., 2005. BZLF1 activation of themethylated form of the BRLF1 immediate-early promoter is regulated by BZLF1residue 186. J. Virol. 79, 7338–7348.

Cayrol, C., Flemington, E.K., 1996. The Epstein–Barr virus bZIP transcription factor Ztacauses G0/G1 cell cycle arrest through induction of cyclin-dependent kinaseinhibitors. EMBO J. 15, 2748–2759.

Chang, L.K., Liu, S.T., 2000. Activation of the BRLF1 promoter and lytic cycle of Epstein–Barrvirus by histone acetylation. Nucleic Acids Res. 28, 3918–3925.

Chi, T., Carey, M., 1993. The ZEBRA activation domain: modular organization andmechanism of action. Mol. Cell. Biol. 13, 7045–7055.

Countryman, J., Miller, G., 1985. Activation of expression of latent Epstein–Barrherpesvirus after gene transfer with a small cloned subfragment of heterogeneousviral DNA. Proc. Natl. Acad. Sci. U. S. A. 82, 4085–4089.

Daikoku, T., Kudoh, A., Fujita, M., Sugaya, Y., Isomura, H., Shirata, N., Tsurumi, T., 2005.Architecture of replication compartments formed during Epstein–Barr virus lyticreplication. J. Virol. 79, 3409–3418.

Daikoku, T., Kudoh, A., Sugaya, Y., Iwahori, S., Shirata, N., Isomura, H., Tsurumi, T., 2006.Postreplicative mismatch repair factors are recruited to Epstein–Barr virusreplication compartments. J. Biol. Chem. 281, 11422–11430.

de Bruyn Kops, A., Knipe, D.M., 1988. Formation of DNA replication structures in herpesvirus-infected cells requires a viral DNA binding protein. Cell 55, 857–868.

Decaussin, G., Leclerc, V., Ooka, T., 1995. The lytic cycle of Epstein–Barr virus in thenonproducer Raji line can be rescued by the expression of a 135-kilodalton proteinencoded by the BALF2 open reading frame. J. Virol. 69 (11), 7309–7314.

Delecluse, H.J., Hilsendegen, T., Pich, D., Zeidler, R., Hammerschmidt, W., 1998.Propagation and recovery of intact, infectious Epstein–Barr virus from prokaryoticto human cells. Proc. Natl. Acad. Sci. U. S. A. 95, 8245–8250.

El-Guindy, A.S., Miller, G., 2004. Phosphorylation of Epstein–Barr virus ZEBRA protein atits casein kinase 2 sites mediates its ability to repress activation of a viral lytic cyclelate gene by Rta. J. Virol. 78 (14), 7634–7644.

El-Guindy, A., Heston, L., Delecluse, H.J., Miller, G., 2007. Phosphoacceptor site S173 inthe regulatory domain of Epstein–Barr Virus ZEBRA protein is required for lyticDNA replication but not for activation of viral early genes. J. Virol. 81 (7),3303–3316.

Farrell, P.J., Rowe, D.T., Rooney, C.M., Kouzarides, T., 1989. Epstein–Barr virus BZLF1trans-activator specifically binds to a consensus AP-1 site and is related to c-fos.EMBO J. 8, 127–132.

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

Fixman, E.D., Hayward, G.S., Hayward, S.D., 1995. Replication of Epstein–Barr virusoriLyt: lack of a dedicated virally encoded origin-binding protein and dependenceon Zta in cotransfection assays. J. Virol. 69 (5), 2998–3006.

Flemington, E., Speck, S.H., 1990. Evidence for coiled-coil dimer formation by anEpstein–Barr virus transactivator that lacks a heptad repeat of leucine residues.Proc. Natl. Acad. Sci. U. S. A. 87, 9459–9463.

Flemington, E.K., Borras, A.M., Lytle, J.P., Speck, S.H., 1992. Characterization of theEpstein–Barr virus BZLF1 protein transactivation domain. J. Virol. 66, 922–999.

Francis, A.L., Gradoville, L., Miller, G., 1997. Alteration of a single serine in the basicdomain of the Epstein–Barr virus ZEBRA protein separates its functions oftranscriptional activation and disruption of latency. J. Virol. 71, 3054–3061.

Francis, A., Ragoczy, T., Gradoville, L., Miller, G., 1999. Amino acid substitutions revealdistinct functions of serine 186 of the ZEBRA protein in activation of lytic cyclegenes and synergy with the EBV Rta transactivator. J. Virol. 73, 4543–4551.

Fujii, K., Yokoyama, N., Kiyono, T., Kuzushima, K., Homma, M., Nishiyama, Y., Fujita, M.,Tsurumi, T., 2000. The Epstein–Barr virus Pol catalytic subunit physically interactswith the BBLF4–BSLF1–BBLF2/3 complex. J. Virol. 74, 2550–2557.

Gao, Z., Krithivas, A., Finan, J.E., Semmes, O.J., Zhou, S., Wang, Y., Hayward, S.D., 1998. TheEpstein–Barr virus lytic transactivator Zta interacts with the helicase–primasereplication proteins. J. Virol. 72, 8559–8567.

Glaser, R., O, 'Neill, F.J., 1972. Hybridization of Burkitt lymphoblastoid cells. Science 176(40), 1245–1247.

Gradoville, L., Grogan, E., Taylor, N., Miller, G., 1990. Differences in the extent of activationof Epstein–Barr virus replicative gene expression among four nonproducer cell linesstably transformed by oriP/BZLF1 plasmids. Virology 178 (2), 345–354.

Gradoville, L., Kwa, D., El-Guindy, A., Miller, G., 2002. Protein kinase C-independentactivation of the Epstein–Barr virus lytic cycle. J. Virol. 76 (11), 5612–5626.

Hamatake, R.K., Bifano, M., Hurlburt, W.W., Tenney, D.J., 1997. A functional interactionof ICP8, the herpes simplex virus single-stranded DNA-binding protein, and thehelicase–primase complex that is dependent on the presence of the UL8 subunit.J. Gen. Virol. 78, 857–865.

Hammerschmidt, W., Sugden, B., 1988. Identification and characterization of oriLyt, alytic origin of DNA replication of Epstein–Barr virus. Cell 55 (3), 427–433.

Henle, G., Henle, W., Klein, G., 1971. Demonstration of two distinct components in theearly antigen complex of Epstein–Barr virus-infected cells. Int. J. Cancer 8 (2),272–282.

Heston, L., El-Guindy, A., Countryman, J., Dela Cruz, C., Delecluse, H.J., Miller, G., 2006.Amino acids in the basic domain of Epstein–Barr Virus ZEBRA protein play distinctroles in DNA binding, activation of early lytic gene expression and promotion ofviral DNA replication. J. Virol. 80, 9115–9133.

Kenney, S.C., Holley-Guthrie, E., Quinlivan, E.B., Gutsch, D., Zhang, Q., Bender, T., Giot, J.F.,Sergeant, A.,1992. The cellular oncogene c-myb can interact synergistically with theEpstein–Barr virus BZLF1 transactivator in lymphoid cells. Mol. Cell. Biol. 12 (1),136–146.

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

Kudoh, A., Fujita, M., Zhang, L., Shirata, N., Daikoku, T., Sugaya, Y., Isomura, H.,Nishiyama, Y., Tsurumi, T., 2005. Epstein–Barr virus lytic replication elicits ATMcheckpoint signal transductionwhile providing an S-phase-like cellular environment.J. Biol. Chem. 280 (9), 8156–8163.

Li, J.S., Zhou, B.S., Dutschman, G.E., Grill, S.P., Tan, R.S., Cheng, Y.C., 1987. Association ofEpstein–Barr virus early antigen diffuse component and virus-specified DNApolymerase activity. J. Virol. 61 (9), 2947–2949.

Liao, G., Wu, F.Y., Hayward, S.D., 2001. Interaction with the Epstein–Barr virus helicasetargets Zta to DNA replication compartments. J. Virol. 75 (18), 8792–8802.

Lieberman, P.M., Hardwick, J.M., Sample, J., Hayward, G.S., Hayward, S.D., 1990. The Ztatransactivator involved in induction of lytic cycle gene expression in Epstein–Barrvirus-infected lymphocytes binds to both AP-1 and ZRE sites in target promoter andenhancer regions. J. Virol. 64 (3), 1143–1155.

Liptak, L.M., Uprichard, S.L., Knipe, D.M., 1996. Functional order of assemby of herpessimplex virus DNA replication proteins into prereplicative site structures. J. Virol. 70(3), 1759–1767.

Liu, P., Speck, S.H., 2003. Synergistic autoactivation of the Epstein–Barr virusimmediate-early BRLF1 promoter by Rta and Zta. Virol. 310 (2), 199–206.

Lukonis, C.J., Weller, S.K., 1996. Characterization of nuclear structures in cells infectedwith herpes simplex virus type 1 in the absence of viral DNA replication. J. Virol. 70,1751–1758.

Lukonis, C.J., Burkham, J., Weller, S.K., 1997. Herpes simplex virus type 1 prereplicativesites are a heterogeneous population: only a subset are likely to be precursors toreplication compartments. J. Virol. 71 (6), 4771–4781.

Marsden, H.S., McLean, G.W., Barnard, E.C., Francis, G.J., MacEachran, K., Murphy, M.,McVey, G., Cross, A., Abbotts, A.P., Stow, N.D., 1997. The catalytic subunit of the DNApolymerase of herpes simplex virus type 1 interacts specificallywith the C terminus ofthe UL8 component of the viral helicase–primase complex. J. Virol. 71, 6390–6397.

McLean, G.W., Abbotts, A.P., Parry, M.E., Marsden, H.S., Stow, N.D., 1994. The herpessimplex virus type 1 origin-binding protein interacts specifically with the viral UL8protein. J. Gen. Virol. 75, 2699–2706.

Miller, G., Heston, L., Countryman, J., 1985. P3HR-1 Epstein–Barr virus withheterogeneous DNA is an independent replicon maintained by cell-to-cell spread.J. Virol. 54 (1), 45–52.

Morrison, T.E., Kenney, S.C., 2004. BZLF1, an Epstein–Barr virus immediate-earlyprotein, induces p65 nuclear translocation while inhibiting p65 transcriptionalfunction. Virol. 328 (2), 219–232.

Page 18: Mutations of amino acids in the DNA-recognition domain of Epstein–Barr virus ZEBRA protein alter its sub-nuclear localization and affect formation of replication compartments

162 R. Park et al. / Virology 382 (2008) 145–162

Morrison, T.E., Mauser, A., Klingelhutz, A., Kenney, S.C., 2004. Epstein–Barr virusimmediate-earlyproteinBZLF1 inhibits tumornecrosis factoralpha-induced signalingand apoptosis by downregulating tumor necrosis factor receptor 1. J. Virol. 78 (1),544–549.

Pearson, G.R., Vroman, B., Chase, B., Sculley, T., Hummel, M., Kieff, E., 1983. Identificationof polypeptide components of the Epstein–Barr virus early antigen complex withmonoclonal antibodies. J. Virol. 47 (1), 193–201.

Petosa, C., Morand, P., Baudin, F., Moulin, M., Artero, J.B., Muller, C.W., 2006. Structuralbasis of lytic cycle activation by the Epstein–Barr virus ZEBRA protein. Mol. Cell 21(4), 565–572.

Quinlan, M.P., Chen, L.B., Knipe, D.M., 1984. The intranuclear location of a herpessimplex virus DNA–binding protein is determined by the status of viral DNAreplication. Cell 36, 857–868.

Quinlivan, E.B., Holley-Guthrie, E.A., Norris, M., Gutsch, D., Bachenheimer, S.L., Kenney,S.C., 1993. Direct BRLF1 binding is required for cooperative BZLF1/BRLF1 activationof the Epstein–Barr virus early promoter, BMRF1. Nucleic Acids Res. 21 (14),1999–2007.

Rickinson, A.B., Kieff, E., 1996. Epstein–Barr virus, In: Fields, B.N., Knipe, D.M., Howley, P.M. (Eds.), 3rd Ed. Fields Virology, Vol. 2. Lippincott-Raven, Philadelphia, PA,pp. 2397–2446.

Rooney, C., Taylor, N., Countryman, J., Jenson, H., Kolman, J., Miller, G., 1988. Genomerearrangements activate the Epstein–Barr virus gene whose product disruptslatency. Proc. Natl. Acad. Sci. U. S. A. 85 (24), 9801–9805.

Schepers, A., Pich, D., Hammerschmidt, W., 1993. A transcription factor with homologyto the AP-1 family links RNA transcription and DNA replication in the lytic cycle ofEpstein–Barr virus. EMBO J. 12 (10), 3921–3929.

Schepers, A., Pich, D., Hammerschmidt, W., 1996. Activation of oriLyt, the lytic origin ofDNA replication of Epstein–Barr virus, by BZLF1. Virol. 220 (2), 367–376.

Takagi, S., Takada, K., Sairenji, T., 1991. Formation of intranuclear replicationcompartments of Epstein–Barr virus with redistribution of BZLF1 and BMRF1gene products. Virol. 185, 309–315.

Taylor, N., Flemington, E., Kolman, J.L., Baumann, R.P., Speck, S.H., Miller, G., 1991. ZEBRAand a Fos-GCN4 chimeric protein differ in their DNA-binding specificities for sites inthe Epstein–Barr virus BZLF1 promoter. J. Virol. 65, 4033–4041.

Taylor, T.J., McNamee, E.E., Day, C., Knipe, D.M., 2003. Herpes simplex virus replicationcompartments can formbycoalescence of smaller compartments. Virol. 309, 232–247.

Uprichard, S.L., Knipe, D.M., 1997. Assembly of herpes simplex virus replication proteinsat two distinct intranuclear sites. Virol. 229, 113–125.

Van der Eb, A.J., Mulder, C., Graham, F.L., Houweling, A., 1977. Transformation withspecific fragments of adenovirus DNAs. I. Isolation of specific fragments withtransforming activity of adenovirus 2 and 5 DNA. Virol. 2 (3–4), 115–132.

Wang, J.T., Yang, P.W., Lee, C.P., Han, C.H., Tsai, C.H., Chen, M.R., 2005. Detection ofEpstein–Barr virus BGLF4 protein kinase in virus replication compartments andvirus particles. J. Gen. Virol. 86, 3215–3225.

Wu, F.Y., Chen, H., Wang, S.E., ApRhys, C.M., Liao, G., Fujimuro, M., Farrell, C.J., Huang, J.,Hayward, S.D., Hayward, G.S., 2003. CCAAT/enhancer binding protein alphainteracts with ZTA and mediates ZTA-induced p21(CIP-1) accumulation and G(1)cell cycle arrest during the Epstein–Barr virus lytic cycle. J. Virol. 77 (2), 1481–1500.

Yokoyama, N., Fujii, K., Hirata, M., Tamai, K., Kiyono, T., Kuzushima, K., Nishiyama, Y.,Fujita, M., Tsurumi, T., 1999. Assembly of the Epstein–Barr virus BBLF4, BSLF1 andBBLF2/3 proteins and their interactive properties. J. Gen. Virol. 80, 2879–2887.

Zeng, Y., Middeldorp, J., Madjar, J.J., Ooka, T., 1997. Amajor DNA binding protein encodedby BALF2 open reading frame of Epstein–Barr virus (EBV) forms a complex withother EBV DNA-binding proteins: DNAase, EA-D, and DNA polymerase. Virol. 239(2), 285–295.

Zhang, C.X., Decaussin, G., Daillie, J., Ooka, T.,1988. Altered expression of two Epstein–Barrvirus early genes localized in BamHI-A in nonproducer Raji cells. J. Virol. 62 (6),1862–1869.

Zhang, Q., Hong, Y., Dorsky, D., Holley-Guthrie, E., Zalani, S., Elshiekh, N.A., Kiehl, A., Le, T.,Kenney, S., 1996. Functional and physical interactions between the Epstein–Barrvirus (EBV) proteins BZLF1 and BMRF1: effects on EBV transcription and lyticreplication. J. Virol. 70 (8), 5131–5142.