elucidation of functional domains of chandipura virus nucleocapsid protein involved in...

10
Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation Arindam Mondal a,b , Raja Bhattacharya a,b,1 , Tridib Ganguly c , Subhradip Mukhopadhyay a,b,2 , Atanu Basu d , Soumen Basak a,b,3 , Dhrubajyoti Chattopadhyay a,b, a Department of Biochemistry, University of Calcutta, Kolkata, India b Department of Biotechnology, University of Calcutta, Kolkata, India c Department of Biological Sciences, Indian Institute of Science Education and Research, Kolkata, India d National Institute of Virology (ICMR), Pune, India abstract article info Article history: Received 30 April 2010 Returned to author for revision 4 June 2010 Accepted 20 July 2010 Available online 21 August 2010 Keywords: Chandipura virus (CHPV) Nucleocapsid protein Encapsidation Oligomerization Leader RNA Na-deoxycholate (DOC) Chandipura virus, a member of the vesiculovirus genera, has been recently recognized as an emerging human pathogen. Previously, we have shown that Chandipura virus Nucleocapsid protein N is capable of binding to both specic viral leader RNA as well as non-viral RNA sequences, albeit in distinct monomeric and oligomeric states, respectively. Here, we distinguish the regions of N involved in oligomerization and RNA binding using a panel of deletion mutants. We demonstrate that deletion in the N-terminal arm completely abrogates self-association of N protein. Monomer N specically recognizes viral leader RNA using its C-terminal 102 residues, while oligomerization generates an additional RNA binding surface involving the N-terminal 320 amino acids of N overlapping with a protease resistant core that is capable of forming nucleocapsid like structure and also binding heterogeneous RNA sequences. Finally, we propose a model to explain the mechanism of genome encapsidation of this important human pathogen. © 2010 Elsevier Inc. All rights reserved. Introduction Chandipura virus (CHPV), a member of the vesiculovirus genera and rhabdoviridae family, was shown to be associated with recent outbreaks of encephalitis in parts of India (Basak et al., 2007; Tandale et al., 2008; Chadha et al., 2005). CHPV possesses a non-segmented negative sense RNA genome that is encapsidated by nucleocapsid protein (N) into a helical structure. The viral RNA dependent RNA polymerase (RdRp), comprised of large protein (L) and its cofactor phosphoprotein (P), remains associated with the encapsidated genome. The genomic RNA in its encapsidated form is not only protected from cellular ribonucleases but also acts as an obligatory template for viral RNA synthesis. The CHPV genome comprises of a 49 nt leader gene (l), followed by ve transcriptional units coding for viral polypeptides and a short non-transcribed trailer sequence (t). Upon infection, viral RdRp transcribes the genome from the 3end to sequentially synthesize leader RNA (l-RNA) and 5 discrete viral mRNAs. During replication, however, the same polymerase reads through the gene junctions and produces a positive sense comple- ment of the genome that then acts as a template for the synthesis of progeny genome (Basak et al., 2007; Banerjee et al., 1987). The strategy employed by the virus in switching from discrete mRNAs to full-length genome synthesis remains elusive. Nucleocapsid assembly occurs concomitant to rhabdoviral genome replication (Gubbay et al., 2001). It was suggested that N-mediated encapsidation of nascent viral RNA is important for suppressing the transcriptiontermination signal during replication (Blumberg et al., 1981). The encapsidation of viral genome relies on critical NN as well as NRNA contacts. For VSV and CHPV, the initiation signal for encapsidation resides in the rst half of the positive strand leader RNA (Bhattacharya et al., 2006; Blumberg et al., 1983; Moyer et al., 1991; Pattnaik et al., 1995). A general comparison of Mononegavirales N protein sequences revealed a short conserved central region that was implicated in NN interaction (Masters and Banerjee, 1987; Myers et al., 1997). Rhabdoviral nucleo- capsid proteins lack a consensus RNA binding motif within its primary sequence. Crystal structure of both VSV (Green et al., 2006) and Rabies virus (Albertini et al., 2006)NRNA complex indicated that this central region is constituted of two structural lobes, which form a hydrophobic cavity with positively charged residues that interact with the RNA phosphate backbone. A similar strategy is used by respiratory syncytial virus to accommodate RNA in its NRNA rings (Tawar et al., 2009). Further, structural comparison of the nucleoproteins from three negative strand RNA viruses Borna disease virus, VSV and Inuenza A virus Virology 407 (2010) 3342 Corresponding author. Department of Biochemistry, University of Calcutta, 35 Ballygunge Circular Road, Kolkata 700019, India. Fax: +91 33 24614983. E-mail address: [email protected] (D. Chattopadhyay). 1 Albert Einstein College of Medicine, USA. 2 Center for Vascular and Inammatory Diseases, University of Maryland School of Medicine, Baltimore, MD, USA. 3 National Institute of Immunology, New Delhi, India. 0042-6822/$ see front matter © 2010 Elsevier Inc. All rights reserved. doi:10.1016/j.virol.2010.07.032 Contents lists available at ScienceDirect Virology journal homepage: www.elsevier.com/locate/yviro

Upload: arindam-mondal

Post on 13-Sep-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

Virology 407 (2010) 33–42

Contents lists available at ScienceDirect

Virology

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

Elucidation of functional domains of Chandipura virus Nucleocapsid protein involvedin oligomerization and RNA binding: Implication in viral genome encapsidation

Arindam Mondal a,b, Raja Bhattacharya a,b,1, Tridib Ganguly c, Subhradip Mukhopadhyay a,b,2, Atanu Basu d,Soumen Basak a,b,3, Dhrubajyoti Chattopadhyay a,b,⁎a Department of Biochemistry, University of Calcutta, Kolkata, Indiab Department of Biotechnology, University of Calcutta, Kolkata, Indiac Department of Biological Sciences, Indian Institute of Science Education and Research, Kolkata, Indiad National Institute of Virology (ICMR), Pune, India

⁎ Corresponding author. Department of BiochemistBallygunge Circular Road, Kolkata 700019, India. Fax: +

E-mail address: [email protected] (D. Chattopadhyay).1 Albert Einstein College of Medicine, USA.2 Center for Vascular and Inflammatory Diseases, Un

Medicine, Baltimore, MD, USA.3 National Institute of Immunology, New Delhi, India.

0042-6822/$ – see front matter © 2010 Elsevier Inc. Adoi:10.1016/j.virol.2010.07.032

a b s t r a c t

a r t i c l e i n f o

Article history:Received 30 April 2010Returned to author for revision 4 June 2010Accepted 20 July 2010Available online 21 August 2010

Keywords:Chandipura virus (CHPV)Nucleocapsid proteinEncapsidationOligomerizationLeader RNANa-deoxycholate (DOC)

Chandipura virus, a member of the vesiculovirus genera, has been recently recognized as an emerginghuman pathogen. Previously, we have shown that Chandipura virus Nucleocapsid protein N is capable ofbinding to both specific viral leader RNA as well as non-viral RNA sequences, albeit in distinct monomericand oligomeric states, respectively. Here, we distinguish the regions of N involved in oligomerization andRNA binding using a panel of deletion mutants. We demonstrate that deletion in the N-terminal armcompletely abrogates self-association of N protein. Monomer N specifically recognizes viral leader RNA usingits C-terminal 102 residues, while oligomerization generates an additional RNA binding surface involving theN-terminal 320 amino acids of N overlapping with a protease resistant core that is capable of formingnucleocapsid like structure and also binding heterogeneous RNA sequences. Finally, we propose a model toexplain the mechanism of genome encapsidation of this important human pathogen.

ry, University of Calcutta, 3591 33 24614983.

iversity of Maryland School of

ll rights reserved.

© 2010 Elsevier Inc. All rights reserved.

Introduction

Chandipura virus (CHPV), a member of the vesiculovirus generaand rhabdoviridae family, was shown to be associated with recentoutbreaks of encephalitis in parts of India (Basak et al., 2007; Tandaleet al., 2008; Chadha et al., 2005). CHPV possesses a non-segmentednegative sense RNA genome that is encapsidated by nucleocapsidprotein (N) into a helical structure. The viral RNA dependent RNApolymerase (RdRp), comprised of large protein (L) and its cofactorphosphoprotein (P), remains associated with the encapsidatedgenome. The genomic RNA in its encapsidated form is not onlyprotected from cellular ribonucleases but also acts as an obligatorytemplate for viral RNA synthesis. The CHPV genome comprises of a 49nt leader gene (l), followed by five transcriptional units coding forviral polypeptides and a short non-transcribed trailer sequence (t).Upon infection, viral RdRp transcribes the genome from the 3′ end tosequentially synthesize leader RNA (l-RNA) and 5 discrete viralmRNAs. During replication, however, the same polymerase reads

through the gene junctions and produces a positive sense comple-ment of the genome that then acts as a template for the synthesis ofprogeny genome (Basak et al., 2007; Banerjee et al., 1987). Thestrategy employed by the virus in switching from discrete mRNAs tofull-length genome synthesis remains elusive.

Nucleocapsid assembly occurs concomitant to rhabdoviral genomereplication (Gubbay et al., 2001). It was suggested that N-mediatedencapsidation of nascent viral RNA is important for suppressing thetranscription–termination signal during replication (Blumberg et al.,1981). The encapsidation of viral genome relies on critical N–N as well asN–RNA contacts. For VSV andCHPV, the initiation signal for encapsidationresides in the first half of the positive strand leader RNA (Bhattacharyaet al., 2006; Blumberg et al., 1983;Moyer et al., 1991; Pattnaik et al., 1995).A general comparison ofMononegaviralesN protein sequences revealed ashort conserved central region that was implicated in N–N interaction(Masters and Banerjee, 1987; Myers et al., 1997). Rhabdoviral nucleo-capsid proteins lack a consensus RNA binding motif within its primarysequence. Crystal structure of both VSV (Green et al., 2006) and Rabiesvirus (Albertini et al., 2006) N–RNA complex indicated that this centralregion is constituted of two structural lobes, which form a hydrophobiccavity with positively charged residues that interact with the RNAphosphate backbone. A similar strategy is used by respiratory syncytialvirus to accommodate RNA in its N–RNA rings (Tawar et al., 2009).Further, structural comparison of the nucleoproteins from three negativestrand RNA viruses – Borna disease virus, VSV and Influenza A virus –

Page 2: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

34 A. Mondal et al. / Virology 407 (2010) 33–42

showed that despite lowhomology in the amino acid sequence, the shapeof the cavity remains invariant, indicating the involvement of similarstructuralmotifs in the nucleocapsid assembly (Luo et al., 2007a,b). Tawaret al. (2009) furtherpointedout conservedpackingof fourα-helices in theC-terminal domain of variousmononegalovirusesNproteins (Tawar et al.,2009), suggesting a possible functional importance of this domain.Mutational analysishasalso identified threegroupsof intermolecularN–Ncontacts spanning four neighboring N molecules (Zhang et al., 2008)involved in stabilizing the encapsidation complex. Although these studieshave identified the molecular interactions that stabilize the nucleocapsidcomplex in vesiculoviruses, however, the dynamic interactions involvedin the initiation phase of encapsidation that mediate specificity towardsviral RNA sequences remain largely unknown. Further, the domains of Nprotein of Chandipura virus involved in N–N or N–RNA interactions havenot been yet characterized.

The N protein, if expressed alone, forms aggregates (Majumdar et al.,2004) and binds non-viral RNA (Bhattacharya et al., 2006; Green et al.,2000; 2006; Iseni et al., 1998; Yang et al., 1998). Previously, we haveshown (Bhattacharya et al., 2006) that the CHPV N protein exists in twodifferent forms, a higher oligomeric form and a monomeric form. In itsoligomeric form, N demonstrates broad RNA binding specificity andinteracts with both viral and non-viral RNAs. However, monomer N onlyrecognizes the specific sequence element present in viral leader RNA.Indeed, P was shown to prevent the oligomerization of CHPV N andmaintains it in monomeric form thereby imparting RNA bindingspecificity on N (Majumdar et al., 2004; Bhattacharya et al., 2006).Based on our observations, we proposed a model where monomer Nspecifically recognizes viral leader RNA in the nucleation step (initiation).In the elongation phase, however, N oligomerization generates broadRNA binding specificity allowing progressive encapsidation of diversesequences present on the CHPV genomic RNA. Our hypothesiscorroborates with the model proposed by Lamb and Kolakofsky toexplain rhabdoviral genome encapsidation (Lamb and Kolakofsky, 2001).However the mechanisms that allow CHPV N to bind RNA with differentspecificity in the nucleation and elongation phases remain unclear.

Here, we delineate the regions of CHPV N responsible for self-assembly and those involved in RNA binding. Through our domainanalysis, we demonstrate that a protease resistant N-terminal core ofthe CHPV N protein possesses all the structural elements required forthe formation of ring shaped nucleocapsid like structures. Synergisticparticipation of two distinct regions within this protease resistantcore appears to be indispensable for oligomerization of N. The samecore region, upon N protein oligomerization, also generates a RNAbinding interface that can bind to diverse RNA sequences. Interest-ingly, however, a C-terminal region active in its monomeric formenables N to specifically bind viral leader RNA sequences. Ourobservations allow us to propose a refined model that includesmechanistic explanations for the observed differences in the RNAbinding specificities of N during discrete stages of encapsidation.

Results

Chandipura virus N protein adopts a conformation similar to that ofother rhabdoviral nucleoproteins

Pairwise comparison of the nucleoprotein sequences of members ofthe rhabdoviridae family identifies N as the most conserved structuralprotein (Marriott, 2005). VSV and CHPV N proteins contain 422 aminoacids each and have ~50% identity (63%with conservative replacements)(Masters and Banerjee, 1987). VSV and Rabies virus nucleoproteins sharea bi-lobed structure with a RNA molecule tightly sequestered betweenthe lobes in an oligomeric assembly (Albertini et al., 2006; Green et al.,2006).

Previously, we have described expression and purification of CHPVN protein from E. coli in a P protein independent system (Bhattacharyaet al., 2006; Majumder et al., 2001). The purified N protein was shown

to be free of bacterial RNA (Majumdar et al., 2004). To verify whetherthis purified form of CHPV N is conformationally similar to otherrhabdoviral N proteins, we examined the protease sensitivity of CHPVN. The recombinant N protein was incubated with chymotrypsin forthe indicated durations (Fig. 1A) and was resolved using SDS-PAGE.Protease digestion for only 2 min produced ~30 kDa and ~15 kDapolypeptides at the expense of the 45 kDa full-length protein(Fig. 1A). Extended digestion, however, resulted in further decay ofthese initial fragments. Formation of two distinct fragments indicatesthe presence of a single highly exposed chymotrypsin site on Nprotein. N protein harbors 46 chymotrypsin sites in its primarystructure as predicted by the Expasy peptide cutter. To investigate theexact location of the exposed site, we repeated the limited proteolysisexperiment using an N-terminal histidine tagged N (His-N). His-Ndisplayed a similar chymotrypsin digestion pattern compared tountagged N on SDS-PAGE (Fig. 1B). Subsequent immunoblotting usingan anti-His antibody detected an N-terminal 30 kDa fragment(Fig. 1C) indicating that the exposed chymotrypsin site residesbetween residues 270 and 280 in the native N protein. Our resultcorroborates with similar proteolytic cleavage patterns of VSV andRabies virus N proteins (Banerjee et al., 1987; Sarkar et al., 2010; Iseniet al., 1998) and was utilized to design various deletion mutants ofCHPV N, as discussed later.

Next, we performed transmission electron microscopy afternegative staining of the soluble N protein preparation. RecombinantCHPV N protein consistently revealed a ring like morphology withboth decameric (diameter 12 nm±0.5 nm) and undecameric subunitcompositions (Fig. 1C) similar to that observed in VSV N (Chen et al.,2004; Green et al., 2000). These rings were thought to act as buildingblocks for the characteristic helical structures of viral RNPs (Iseni et al.,1998; Thomas et al., 1985). A similar proteolytic cleavage pattern, ringlike assembly and strong sequence homology underscores thestructural similarity of the CHPV N protein with nucleoproteins ofother related Rhabdoviruses.

Construction of a panel of deletion mutants of CHPV N protein

Crystal structures of the VSV and Rabies N proteins (Green et al.,2006; Albertini et al., 2006) reveal that N has two lobes containingmainly α helices, which come together to form a cavity thataccommodates RNA. The N-terminal lobe contains seven α helicesalong with four β strands. The C-terminal lobe begins at residueSer220 and contains eight α helices. An N-terminal arm (residues 1–22) containing two anti parallel β strands and a C-terminal extendedloop (residues 340–375) were also found to have important roles inoligomerization and RNA binding (Green et al., 2006; Zhang et al.,2008). Crystal structure of respiratory syncytial virus N protein alsopointed out to the similar bi-lobed morphology but comparativelysmaller C-terminal domain (107 residues) (Tawar et al., 2009).

Considering the conformational similarity (Fig. 1) as well assequence homology between VSV and CHPV N proteins, we havegenerated a panel of deletion mutants to dissect the role of differentregions of CHPV N in oligomerization and RNA binding. We haveconstructed N(1–320) to check the functional relatedness of thechymotrypsin-resistant N-terminal core with wild type N protein.Also the C-terminal lobe (residues 221–422 in VSV N, Green et al.,2006) was deleted to generate N(1–220) and almost the entire N-terminal lobe was deleted to generate N(180–422). Deletion of the N-terminal arm (residues 1–22) of VSV N was shown to hamper its RNAbinding ability (Zhang et al., 2008). Similarly, we have deleted N-terminal 47 amino acids to generate the mutant N(48–422). A shortpolypeptide composed of the first N-terminal 47 amino acids was alsogenerated. Comparison of different rhabdovirus N protein sequencesrevealed that the central region of the protein (particularly residues190–270) is highly conserved. To test the functional importance of

Page 3: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

Fig. 1. CHP virus N Protein adopts a conformation similar to that of other rhabdoviral nucleoproteins. Partial proteolysis of untagged wild type (A) or N-terminal histidine tagged Nprotein (B) was carried out with Chymotrypsin for different time periods as indicated. The resulting digested products were resolved through SDS-PAGE and stained with CoomassieBlue. (C) Chymotrypsin digested products of His-N (identical to B) were immunoblotted against anti-His antibody. (D) Representative transmission electron micrographs of full-length N protein negative stained with uranyl acetate at a magnification of 135,000 k. Size bar is built into the micrograph. Characteristic ring shaped structure is highlighted in theinset.

35A. Mondal et al. / Virology 407 (2010) 33–42

this region, we have generated N(1/180–265/422) harboring aninternal deletion of 85 amino acids spanning residues 180–265.

These deletion mutants (Fig. 2A) were expressed in E. coli, purified(Fig. 2B, Materials and methods section) and confirmed by westernblot (Fig. 2C). Spectrophotometric measurement of absorbance at 260and 280 nm indicated the absence of contaminating bacterial RNAs inthese truncated protein preparations. Further, we tested phenol-chloroform extracts obtained from these protein preparations inUrea-PAGE to confirm the absence of nucleic acids. Circular dichroismwas performed to determine the structural integrity of the individualdeletion mutants of N. We have measured the α helical content ofdifferent deletion mutants of CHPV N protein to compare with thetheoretical α helical content of the respective deletion mutantsestimated by using X-ray crystal structure of the highly homologousVSV N (Green et al., 2006). Table 1 shows that the percentage of αhelical content of the deletion mutants, calculated by a CDdeconvolution software, corroborates well with the respectivetheoretical values. These data highlight that the secondary structureof different domains of N protein are largely retained in the deletionmutants.

Two discrete regions of N protein control the homotypic interaction

To understand how the oligomeric state of the N protein dictatesits RNA binding specificity, we attempted to dissect RNA binding fromoligomerization. To map the region of N protein involved in formingand stabilizing the oligomeric form, we utilized a His-tag pull downassay. CHPV N was shown to undergo reversible conversion from

oligomer to monomer upon treatment with an anionic detergentSodium deoxycholate (DOC), where it reverts back to its oligomericstate upon removal of the detergent (Bhattacharya et al., 2006). Here,we utilized a modified version of His-tag pull down assay to test theability of the truncated N proteins to interact with the full-length N. N-terminal histidine tagged full-length N (His-N) and each of thetruncated proteins weremixed together in 1:1molar ratio and treatedwith DOC. The mixture was then dialyzed to remove DOC andincubated with Ni–NTA resin (Materials and Methods section). Afterwash, His-N was eluted from the resin with 250 mM imidazole,thereby also co-eluting proteins that remain bound to His-N.Untagged full-length N, used as a positive control, eluted exclusivelywith His-N indicating a strong N–N interaction (Fig. 3A, lane 2). Thecarboxy-terminal deletion mutants, N(1–320) and N(1–220) alsointeracted with His-N in our assay (Figs. 3B and C) whereas the aminoterminal deletion mutant N(180–422) and N(48–422) eluted entirelyin the flow-through and wash fractions (Figs. 3E and F). These resultssuggest a dispensable role of the C-terminal half of the N protein inoligomerization, while first 47 amino acids appear to be important.However, we noticed that a significant fraction of N(1–220) elutes inthe flow-through implying a weak interaction in this mutant ascompared to N(1–320). We also analyzed the N(1–47) mutant tofurther confirm an important role of the N-terminal 47 amino acids inthe homotypic interaction (Fig. 3D). Interestingly, N(1/180–265/422)harbors an internal deletion, also failed to interact with His-N(Fig. 3G). In summary, our analyses suggest that an N-terminal regioncontaining first 47 amino acids and a central region encompassingresidues 180 to 265 are critical for the self-assembly of N.

Page 4: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

Fig. 2. Different deletion mutants of N protein. (A) Schematic representation of thedeletion mutants of N. The black boxes represent the regions that are present in therespective deletion mutants. (B) A SDS-PAGE depicting the purity of the preparations ofvarious deletion mutants of N, namely N(1–47) (lane 3), N(1–220) (lane 4), N(1–320)(lane 5), N(1/180–265/422) (lane 6), N(180–422) (lane 7) and N(48–422) (lane 8)after ion-exchange chromatography. Both untagged (lane 1) and His-tagged (lane 2)full-length N Proteins were also included as controls. (C) To confirm the identity ofvarious N deletion mutants, the gel in (B) was subjected to western blot analysis usinganti-N antibody.

Table 2Oligomerization status of the deletion mutants of N.

Truncated N proteins Estimated molecular weight (kDa) Oligomerizationstatus

N(1–320) 302 OctamerN(1–220) 108 TetramerN(1–47) 16 TrimerN(48–422) 42 MonomerN(180–422) 18.5 MonomerN(1/180–265/422) 67 Dimer

36 A. Mondal et al. / Virology 407 (2010) 33–42

Both amino terminal and central regions are necessary for the assemblyof N oligomers

Next, we examined the oligomeric status of the deletion mutantsby gel filtration and sucrose density gradient centrifugation. To thisend, full-length or deletion mutants of N were subjected to gelfiltration chromatography (Fig. 4A). The molecular weights of thesemutant proteins, in their native forms, were estimated from theelution volumes of the respective peak fractions using proteinmolecular weight standards (Table 2, Materials andmethods section).As described earlier, the oligomeric assembly of full-length N(Bhattacharya et al., 2006) eluted between ferritin (440 kDa) andcatalase (232 kDa). The N-terminal deletion mutants N(48–422) and

Table 1Percentage of α-helix content of wild type and truncated N proteins determinedexperimentally and calculated from crystal structure of VSV N.

Proteins % α-Helix

Experimental Theoretical

Nwt 36.30% 38.86%N(1–320) 35.20% 37.80%N(1–220) 37.30% 35%N(1/180–265/422) 37.50% 33.40%N(48–422) 42.80% 42%N(180–422) 38.40% 44%

N(180–422) eluted with an estimated molecular weight of 41.6 kDaand 20 kDa, respectively, suggesting that these mutants indeed existas monomers. In contrast, N(1–320) eluted as a distinct oligomericspecies indicating that the C-terminal region is dispensable foroligomeric N assembly. Interestingly, N(1–220), which interactsweakly with His-N (Fig. 3), eluted as a lower order tetrameric species.These results suggest that a central region of N spanning 220–320residues may play a role in the ordered self-assembly. Indeed, N(1/180–265/422) co-eluted with BSA (67 kDa) with an estimatedoligomeric state of a dimer. This dimer may have resulted fromweak N–N interaction that was not detected in our His-tag pull downassay. Nevertheless, the inability of this mutant to form higheroligomers confirmed that the central region, in addition to the N-terminal part, is critical for the proper assembly of monomer N intooligomeric complexes.

Additionally, we performed sucrose density gradient centrifuga-tion. Here, full-length N protein or different deletion mutants wereloaded onto a 10–60% pre-formed sucrose gradient and the fractionswere analyzed in SDS-PAGE (Materials and methods section). Asdescribed earlier (Bhattacharya et al., 2006), full-length N wasdetected in fractions 10–12 from the bottom of the gradientcorresponding to an 11.5 S particle (Fig. 4B). However, deoxycholatetreatment disrupts the oligomers, producing the monomeric form,which sediments at the bottom of the gradient as a 3 S particle (datanot shown, (Bhattacharya et al., 2006)). The N-terminal deletionmutants N(48–422) and N(180–422) as well as the central regiondeletionmutant N(1–180/265–422) were detected in fractions 18–20

Fig. 3.His-tag pull down assay identifies the involvement of two discreet regions withinthe N-terminal core domain of N protein in homotypic interaction. Full-length His-Nwasmixed either with untagged full-length N (A) or with the various N protein deletionmutants as indicated (B through G) and was treated with DOC. After removal of DOC,the mixture was loaded onto the Ni–NTA column and washed subsequently with eitherbinding buffer, or with buffer containing 25 mM and 50 mM imidazole. The Ni–NTAbound proteins were finally eluted with 250 mM imidazole (lane 2). All fractionsincluding the flow-through (lane 1) and washings (data not shown) were collected andresolved through 12% SDS-PAGE.

Page 5: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

Fig. 4. Presence of N-terminal 320 amino acids is sufficient for self-association and the formation of ring shaped structures characteristics of wild type N. (A) Full-length and differentdeletion mutants of N proteins were analyzed by gel filtration chromatography on a Superdex200 (10/300 GL) column at 4 °C. Elution profile was expressed as the percent relativeabsorbance at 280 nm. (B) Wild type and different deletion mutants of N protein were subjected to sucrose density gradient centrifugation through 10–60% pre-formed sucrosegradients. Fractions were collected from the bottom of the gradient and resolved through 12% SDS-PAGE. (C) Representative transmission electronmicrographs of N(1-320) negativestained with uranyl acetate at a magnification of 135,000 k(B). Size bars are built into the micrograph. Characteristic ring shaped structure is highlighted in the inset and aggregatedstructure is indicated by arrow.

37A. Mondal et al. / Virology 407 (2010) 33–42

as a 2.9 S particle. Therefore, our sedimentation data indicates anecessary role of the N-terminal (1–47aa) and central (180–265aa)regions of the N protein in its oligomerization. Interestingly, N(1/180–265/422) sedimented as a monomer in contrast to a dimericelution profile in the gel filtration analysis. This difference could beattributed to the instability of weak N(1/180–265/422) dimers in thedensity gradient centrifugation. The C-terminal deletion mutant N(1–320), however, sedimented as a higher oligomer similar to full-lengthN. In sum, CHPV N protein synergistically utilizes both the N-terminaland the central regions to form higher oligomers and the C-terminaldomain is not required for its self-assembly. Unlike the His-tag pulldown assay, gel filtration chromatography and sucrose densitygradient centrifugation provided additional stoichiometric and sta-bility information on the oligomeric status of various truncated Nproteins. These information relate to the distinct involvement ofdifferent regions in stabilizing the oligomeric assembly of N.

Next, we used transmission electron microscopy to examine thestructural features of the oligomers formed by N(1–320), whichretains both the central and the N-terminal N–N interaction domains.Electron micrographs of purified N(1–320) oligomers at a magnifica-tion of 135,000 (Fig. 4C) revealed a similar ring shaped morphologyreminiscent of full-length N with an analogous outer diameterof ~12 nm. We also observed some non-specific clusters formed bystacking of two or more rings (Fig. 4C; white arrow) in this mutant, aresult that is consistent with a proposed negative role of the C-terminal region in the self-assembly of the nucleoproteins of relatedviruses (Elton et al., 1999; Kho et al., 2003). Also a recent report on

VSV N protein showed that the disruption of interactions among theneighboring C-terminal regions results in the formation of ring likestructures larger than decamers (Ge et al., 2010). Previously, it wasshown that trypsin treatment generates a 17 kDa fragment from theC-terminus of Rabies virus N protein without impacting the overallnucleocapsid assembly (Iseni et al., 1998). Consistent with the priorreport, our analysis suggests that the N-terminal 320 residues retainall the necessary elements for mediating N–N interaction, oligomericassembly and ring shaped nucleocapsid formation.

N Protein employs two distinct modes for RNA binding

Wehave previously shown (Bhattacharya et al., 2006) that oligomericN binds to both viral and non-viral RNA to form a high molecular weightcomplex (Complex I). Deoxycholate treatment disrupts N oligomers intomonomers, which then specifically recognize viral leader RNA to form a1:1 N:l-RNA complex (Complex II). Since the oligomeric status of Ncontrols its RNA binding specificity, we sought to examine various Nprotein mutants in an in-vitro RNA binding assay. Increasing concentra-tions of full-length CHPV N or its deletion mutants, either untreated orpretreated with deoxycholate, were incubated with radiolabeled leaderRNA. Resulting RNA–protein complexes were analyzed by EMSA.Consistent with our prior report (Bhattacharya et al., 2006), deoxycholatepretreatment of full-lengthN resulted in the formation of fastermigratinglower order Complex II (Fig. 5A, lanes 5–7) at the expense of Complex I(Fig. 5A, lanes 2–4). Pretreatment with 1% deoxycholate maintained N inits monomeric form, even at a relatively high protein concentration

Page 6: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

38 A. Mondal et al. / Virology 407 (2010) 33–42

(4 μM), as evident from almost exclusive Complex II formation under thiscondition (lane 7). N(1–320) effectively formed Complex I with leaderRNA (Fig. 5B, lanes 2–4). In contrast to full-length N, however,pretreatment with deoxycholate completely impaired the ability of N(1–320) to bind RNA in its monomeric form, even at high proteinconcentrations. These results indicate that C-terminal region spanningthe last 100 residues of N protein is indispensable for RNA binding in itsmonomeric form (treated with deoxycholate). In its oligomeric state, Nutilizes an additional region spanning residues 1–320, which is alsocritical for oligomerization, to form Complex I with RNA. We tested ourhypothesis using another C-terminal deletionmutant N(1–220) (Fig. 5C),which further confirmed the necessity of the C-terminal region for RNAbinding by monomer N.

Similar experiments with the N-terminal deletion mutants N(180–422) and N(48–422) resulted in the formation of Complex II exclusivelyboth in presence and absence of DOC (Figs. 5D and E), suggesting thatoligomerization of N is essential for the formation of Complex I. Deletionof 22 amino acids from the N-terminal arm of VSV N resulted in total lossof RNA from the oligomeric N protein (Zhang et al., 2008). In strikingcontrast, our analysis using N(48–422) indicated that deletion of 47amino acids from the N-terminus does not impair the RNA binding abilityof CHPV N but disrupts its oligomeric assembly (Fig. 5E). N(1/180–265/422), which exists as a mixture of monomers and weakly associateddimers, generated bothmonomer: l-RNA (Complex II) aswell as dimer: l-RNA complexes (Complex I′, Fig. 5F, lane 2). However, N(1/180–265/422) formed mostly dimeric complexes with leader RNA at higherprotein concentrations(Fig. 5F, lane 4). Pretreatment with deoxycholate

Fig. 5.N protein distinct complexes with viral RNA in its different oligomeric states. In-vitro sin the presence of increasing amount of full-length N protein (A), N(1–320) (B), N(1–220) (Cconcentrations of the respective proteins pretreated with 1% DOC for 15 min at 37 °C (Lane

disrupted these dimers into monomers, which then readily bound l-RNAforming monomeric Complex II (Fig. 5F, lanes 5–7). Apart from the factthat oligomerization defectiveN(1/180–265/422) is incapable of forminga higher order Complex I, our results also imply that the conservedresidues from180 to 265 are dispensable for RNAbinding bymonomerN.The ability of weakly associated N(1/180–265/422) dimers to formComplex I′ also signify that Complex I could be generated fromComplex IIthrough protein oligomerization. Altogether our analyses suggest that Nprotein employs two different modes for RNA binding. In its monomericform, the C-terminal region (320–422aa) is indispensable for RNArecognition while in its oligomeric form N relies on an alternate RNAbinding surface generated through protein oligomerization, distinct fromthe C-terminal region.

Distinct domains of N protein impart different RNA binding specificities

Monomeric N–P complex has an inherent ability to specificallyencapsidate only its cognate genome RNA (Bhattacharya et al., 2006;Chen et al., 2007). Considering that N protein utilizes two separateregions for RNA binding; we asked if these two regions also differ intheir RNA binding specificities. To this end, full-length or truncated Nproteins either pretreated or untreated with deoxycholate, wereincubated with an excess of unlabeled non-viral RNAs. Subsequently,radiolabelled leader RNA was added to the reaction prior to resolvingit through native PAGE (Materials and methods section). In ourcompetition assay, we used 2–20 fold molar excess of unlabeled non-viral RNA as compared to labeled leader RNA. Consistent with our

ynthesized radio labeled leader RNAwas incubated in binding buffer at 37 °C for 15 min), N(180–422) (D) N(48–422) (E) and N(1–180/265–422) (F) (Lanes 2–4) or with same5–7). The resulting RNA–protein complexes were resolved through 4% native PAGE.

Page 7: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

39A. Mondal et al. / Virology 407 (2010) 33–42

previous report (Bhattacharya et al., 2006), non-viral RNA efficientlycompeted out viral leader RNA binding to oligomeric (Complex I,Fig. 6A, lanes 2–6) but not monomeric (Complex II, Fig. 6A, lanes 7–11) N protein. Our analysis revealed that leader RNA binding by N(1–320) could be also competed out by non-viral RNA (Fig. 6B, lanes 2–6).Indeed, complete abrogation of Complex I formation with a 5-foldmolar excess of non-viral RNA suggests that the N(1–320) oligomer,similar to N oligomer, binds RNA with a broad sequence specificity.This result was further verified using another C-terminal deletionmutant N(1–220) (Fig. 6C).

In contrast, the oligomerization defective mutants, N(180–422)and N(48–422), which bind RNA as monomers, were able tospecifically recognize viral leader RNA even in the presence of 20fold molar excess of unlabeled non-viral RNA (Figs. 6D and E). Ourdata confirms that the monomeric form of N protein has an inherentability to recognize specific sequences on leader RNA. It also indicatesthat the C-terminal 100 residues of N, which are indispensable forRNA recognition as monomer, also impart RNA binding specificity.Finally we asked whether the two complexes (I′ and II) formed by N(1/180–265/422) with l-RNA behave differently with respect to theRNA binding specificity. Here, we used a concentration of N(1/180–265/422) protein (~5 μg) that allows for the formation of both thecomplexes even in the absence of deoxycholate. Complex I′ thatrepresents the dimer–RNA complex was readily competed out by 5fold molar excess of unlabeled non-specific RNA (Fig. 6F, lane 4). Incontrast, Complex II that engages protein monomer, was stable even

Fig. 6. RNA binding specificity of N protein is determined by its C-terminal region. Full-lengN(1–180/265–422) (F) were incubated in binding buffer without (lane 2) or along with 2-fspecific RNA, derived from pGEM-3z, at 37 °C for 15 min. Subsequently the reaction mixturefinally resolved through 4% native PAGE. Effect of pretreatment with 1% DOC of full-length N p

in the presence of 20 fold molar excess of unlabeled non-viral RNA(Fig. 6F, lanes 2–6). Our result indicates that N(1/180–265/422)monomers preserve the region necessary for sequence specificrecognition of viral leader RNA. Moreover, a careful inspection ofthe competition profiles suggested that N(1/180–265/422) showsmore stringent specificity as compared to N-terminal deletions, N(180–422) or N(48–422). It seems that the mutant N(1/180–265/422) resembles wild type N with respect to its specificity towardsleader RNA in monomeric form. In sum, our analysis suggests that theregions of N protein involved in oligomerization are dispensable forspecific RNA recognition. Instead, N protein utilizes a region spanningits C-terminal 100 residues to specifically bind the viral leadersequence.

Discussion

Encapsidation is not only important for producing transcriptioncompetent genome particles that are resistant to cellular RNases, butalso for the transcription–replication switch (Blumberg et al., 1981).Yet themechanism that allows N protein to specifically recognize viralsequences present at the genome termini during initiation ofencapsidation and to progressively enwrap diverse RNA sequencesinto nucleocapsid structure in the elongation phase remains unclear.Previously, we have reported distinct RNA binding specificities ofmonomer and oligomer N (Bhattacharya et al., 2006). In this study, we

th N (A, lanes 2–6), N(1–320) (B), N(1–220) (C), N(180–422) (D) N(48–422) (E) orold (lane 3), 5-fold (lane 4), 10-fold (lane 5) and 20-fold (lane 6) molar excess of non-was chased with radiolabeled l-RNA for an additional 10 min. The reaction mixture wasrotein on its RNA binding specificity has also been examined as a control (A lanes 7–11).

Page 8: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

40 A. Mondal et al. / Virology 407 (2010) 33–42

have analyzed a panel of deletionmutants of N protein to decipher themechanism of differential RNA binding specificities of N protein.

Similar to VSV and Rabies virus N proteins (Banerjee et al., 1987; Iseniet al., 1998), CHPVN protein harbors an N-terminal core of 30 kDa, that isresistant to chymotrypsin digestion (Fig. 1). By using His-tag pull downassay, gel filtration analysis and sucrose density gradient centrifugationassay, we determined that an N-terminal (1–47) and a central (180–264)region within the chymotrypsin-resistant core are indispensable for theoligomeric assembly of N (Figs. 3 and 4). It is noteworthy that N(1/180–265/422), harboring a deletion in the central domain necessary for N–Ninteraction, formed weak dimers in our gel filtration assay. This dimerwas, however, disrupted into its monomeric form under the centrifugalforce in our density gradient assay (compare Figs. 4A and B). Our datacorroborates well with previous analyses in related VSV, Sendai, measlesand human parainfluenza type 2 viruses, which also reveal an essentialrole of the N-terminal two-thirds of the nucleoprotein in N–N interaction(Buchholz et al., 1993; Myers et al., 1997; Myers et al., 1999; Liston et al.,1997; Nishio et al., 1999). In contrast, the C-terminal 102 amino acidswere not only dispensable for N–N interaction but also for the oligomericassembly of CHPV N into ring like structures (Fig. 4C). Indeed,nucleocapsid morphology was intact upon proteolytic removal of a17 kDa fragment from the C-terminus of the Rabies virus nucleoprotein(Iseni et al., 1998). Based on the essential requirement of both the regionsin oligomerization, our mutational analyses suggest that the N-terminaland the central region of CHPVNparticipate synergistically tomediateN–N interaction, oligomerization and ring like structure formation.

Monomer N, generated by disrupting the N oligomers withdeoxycholate, was shown to bind viral RNA (Fig. 5A and (Bhattacharyaet al., 2006)). Consistent with this analysis, oligomerization defectivemutants, N(48–422), N(180–422), and N(1/180–265/422), efficientlyinteracted with viral leader RNA (Figs. 5D–F) suggesting that an intrinsicRNA binding ability is hardwired in monomeric CHPV N subunits. N(1/180–265/422), which retains six conserved amino acids implicated inmediating RNA contacts within VSV nucleocapsid (Green et al., 2006),produced bothmonomer:l-RNA and dimer:l-RNA complexes. Importanceof a C-terminal residue (R408) of VSV N in RNA binding and subsequentencapsidation (Rainsford et al., 2010) was highlighted. Interestingly,CPHV N(1–320), which lacks the C-terminal region, efficiently boundRNA in its oligomeric form. However, DOC treatment that generatesmonomeric species, completely abrogated its RNA binding ability(Fig. 5B). Therefore, our results suggest that RNA binding by monomerN is controlled through a C-terminal region comprising of 102 aminoacids. However, the ability of N(1–320) to bind RNA as an oligomerindicates that N protein through oligomerization generates an alternateRNA binding surface. Interestingly, these 320 residues retain both the N-terminal arm (1–22aa) and the central region, previously shown to becritical for proper conformational attainment of the RNA binding cavity inVSV N (Green and Luo, 2009; Zhang et al., 2008). Indeed, site-specificmutation (R143K) in this region of VSVN protein impaired encapsidationdue to a possible defect in the elongation phase (Rainsford et al., 2010).

In contrast to the oligomeric form, monomer N acquires specificityfor the viral leader sequence (Bhattacharya et al., 2006). Our analysisreveals that oligomerization defective mutants of N recognize viralRNA specifically to form a 1:1 protein:RNA complex even in thepresence of molar excess of non-viral RNA (Fig. 6). On the contrary, N(1–320) that binds RNA only as oligomer, showed broad RNA bindingspecificity validating the essential role of the C-terminal 102 residuesin specific recognition of viral sequences. Additionally, dimer:l-RNAcomplexes formed by N(1/180–265/422) were efficiently competedout by non-viral RNAs. Higher affinity of N(1/180–265/422) to leaderRNA compared to the N-terminal deletion mutants (Fig. 6) implies anadditional role of the N-terminal 47 amino acids in furtherstabilization of the monomer N:l-RNA complex.

Complex formation with P prevents the non-specific RNA bindingby N (Masters and Banerjee, 1988; Yang et al., 1998). It was shownthat measles virus N protein utilizes overlapping regions responsible

for self-assembly and also P binding (Bankamp et al., 1996). VSV Pprotein was shown to utilize different N and C-terminal domains torecognize monomer N and oligomer N–RNA complex (Chen et al.,2007; Green and Luo, 2009). Although VSV N possesses a C-terminaldomain that mediated interaction between N oligomer and P protein(Green and Luo, 2009), the domain of monomer N responsible for N0–

P complex formation is yet to be elucidated.Based on these results and our current analyses, we propose a

refined model to explain discrete steps of encapsidation. We suggestthat complex formationwith P disrupts N–N assembly and preserves asubset of N in its monomeric form. Monomer N utilizes its C-terminalregion to bind viral RNA specifically. This specific recognition leads tothe formation of the nucleation complex. Subsequent N–N interactionthat involves the N-terminal arm of one N monomer and the centralregion of another adjacent N-moiety generates an RNA bindingsurface, which is capable of accommodating diverse RNA sequences inthe elongation phase of encapsidation. RNA is tightly sequestered inthis cavity through extensive intermolecular interactions. Theinherent flexibility of the N-terminal lobe (Zhang et al., 2008) in theRNA bound oligomeric N seems to be essential for processivity duringencapsidation. Also, contributions from the C-terminal lobe in theformation of a stable nucleocapsid structure cannot be ignored.Oligomerization states were shown to control DNA binding specifi-cities of several host transcription factors (Takemori et al., 2009) andDNA repair enzymes (Popuri et al., 2008). The ability of the integraseprotein of the Human Immunodeficiency virus to recognize viral DNAends was also shown to depend on the oligomeric state of the protein(Lesbats et al., 2008). Recent X-ray crystallographic analyses haveprovided important insights into the molecular interactions thatstabilize the nucleocapsid structure of VSV and other related viruses.Our biochemical analyses exploring the dynamic encapsidationprocess provide additional information on the involvement of discreteregions of the CHPVN protein, in its different oligomeric forms, duringdistinct stages of encapsidation.

In conclusion Chandipura virus differently utilizes two distinctoligomeric states of a single protein during its encapsidation process.Further biophysical and biochemical characterizations of specific RNAbinding by the C-terminal region of N protein may guide designing oftherapeutic molecules that exclusively target viral encapsidationwithout any detrimental side effects on host metabolism.

Materials and methods

Oligonucleotides used in this studywere purchased from Promega,USA. The column chromatography materials were from AmershamBiosciences. Radioactive biomolecules were from BRIT, India. All otherchemicals and biochemicals were of analytical grade.

Cloning of full-length N gene and its deletion mutants

Untagged version of full-length N gene was obtained from pET3a-NC (Majumder et al., 2001) by restriction digestion and cloned intopET33b vector to obtain N-terminal histidine tagged version of N.Similarly, different restriction digested fragments obtained frompET3a-NC were cloned into pET3a to obtain various deletion mutantconstructs expressing N(1–320aa), N(1–220aa) and N(1–47aa).Deleted versions of N gene, N(48–422aa) and N(180–422aa)respectively, were amplified using pET-3aNC template using eitherNC1 or NC3 (corresponds to 130–150 and 530–550 nucleotides of Ngene) as forward primer and DJC2 (Majumder et al., 2001) as reverseprimer. The respective PCR products were cloned into pET3a vectorsubsequent to restriction digestion.

Further, pET3a-NC was digested with NdeI/EcoRI or BalI/BamHI toobtain fragments corresponding to the N-terminal 541 nucleotidesand C-terminal 472 nucleotides, respectively. Both the fragmentswere cloned into pGEM3z vector sequentially. The resulting N gene

Page 9: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

41A. Mondal et al. / Virology 407 (2010) 33–42

lacking nucleotides 542-796 was restriction digested out of thepGEM3z using NdeI and BamHI and ligated into an NdeI/BamHIdouble digested pET3a vector to generate the middle deletant N(1/180–265/422).

Expression and purification of full-length and truncated N proteins

All full-length and deletion mutants of N were expressed in E. coli,BL21(DE3) or in BL21(DE3) pLysS. The expression and purification offull-length untagged protein was carried out as described (Majumderet al., 2001). His-tagged version of full-length Nwas purified using Ni–NTA agarose (QIAGEN). Different deletion mutants of N wereexpressed in E. coli. using 1 mM IPTG at 37 °C for 3 h. Aftersolublization with 0.8% N-Lauroylsarcosine (Sigma) and sonication,the proteins were purified by anion exchange chromatography usingMono-Q (5/50 GL) FPLC column (Amersham Biosciences, Sweden)with a sodium chloride gradient of 100 mM to 1 M in TET buffer(50 mM Tris pH−8.0, 1 mM EDTA and 0.1% Triton X-100). Respectivefractions containing the desired purified proteins were concentratedusing centricon and dialyzed overnight against TET buffer containing100 mM NaCl. Quantifications were made with BIORAD DC proteinassay reagents.

Synthesis of RNA probes

RNAs used in EMSA were synthesized in-vitro by run-offtranscription using T7 RNA polymerase as mentioned (Basak et al.,2004, 2003). In-vitro synthesized RNA was eluted from urea-polyacrylamide gels, precipitated twice with ethanol and suspendedin RNase free water. Radioactivity was measured in liquid scintillationcounter. Unlabelled RNAs used in this study were quantified by theirabsorbance at 260 nm. Non-viral RNA was derived from linearizedpGEM-3z vector.

Electrophoretic mobility shift assay (EMSA)

Electrophoretic mobility shift assays were carried out as described(Bhattacharya et al., 2006). Briefly, different concentrations of full-lengthor truncated N proteins, untreated or treated with 1% deoxycholate(DOC) at room temperature for 15 min, were further incubated withradiolabelled leader (l*)-RNA probe for 10 min. For competition gel shifts,the DOC treated or untreated proteins were first incubated in thepresence of molar excess of cold pGEM3z vector derived non-viral RNAfor 15 min. Subsequently, l*-RNA probe was added to the reactionmixture and incubated for an additional 10 min. RNA–protein complexeswere resolved at 4 °C through 4% native PAGE.

His-tag co-elution assay

His-tagged N was mixed with each of the truncated proteinsseparately in equimolar ratio and the mix was treated with deoxycholatefor 15–20 min. Subsequently it was dialyzed against 100 mM NaCl-TETbuffer containing 10 mM Imidazole. Ni–NTA agarose resins were pre-equilibrated with the same buffer and protein mixes were next allowedto bind to the Ni–NTA agarose in the same buffer. After consecutivewashings with 10 mM, 20 mM and 50mM Imidazole the proteins werefinally eluted with 250 mM Imidazole.

Density gradient assay

10–60% sucrose step gradients of full-length and truncated N proteinswere performed as mentioned previously (Bhattacharya et al., 2006).Briefly the gradients were centrifuged for 16 h at 32,000 rpm in an SW44Beckman rotor at 4 °C; 0.5 ml fractionswere collected from the bottom ofthe gradients, precipitated using TCA and run in 12% SDS-PAGE. RelativeBand intensity was measured by densiometric scanning using Image

quant TL software (Amarsham Biosciences), and plotted against thefraction number (not shown). Sedimentation coefficients were deter-mined by comparing with protein standards (Bhattacharya et al., 2006).

Gel filtration analysis

Gel filtration chromatography was carried out as described(Bhattacharya et al., 2006) using Superdex 200 (10/300 GL) FPLCcolumn (Amersham Biosciences, Sweden) equilibrated with TETbuffer (50 mM Tris (pH 8), 1 mM EDTA, 0.1% TritonX-100 and100 mM NaCl) at 4 °C. Protein concentrations were kept at 0.4 mg/ml. The column was calibrated using Gel filtration calibration kit(Amersham Biosciences). The molecular weight for the proteins wasestimated from the calibration curve.

Transmission electron microscopy

For electron microscopy 0.4 mg/ml of purified full-length andtruncated N proteins were absorbed for 10–15 min on 400 meshcarbon coated copper grids pre-coatedwith Alcian Blue. Samplesweresubsequently washed with water for two to three times. Staining wascarried out with 0.1% uranyl acetate for 1–2 min, dried at roomtemperature and imaged under low beam current b5 uA under120 KV operating voltage in a TECNAI 12 BIOTWIN transmissionelectron microscope (FEI Netherlands). Electron micrographs wereobtained using a side mounted 1 K×1 K CCD camera (Megaview III,SIS Germany). For the N(1–320) mutant, bacitracin was used toremove large aggregates.

Acknowledgments

We acknowledge Arunava Roy and Hirak Patra, University of Calcutta,for the helpful comments on themanuscript and Saswata SahooW.B.U.T.,Kolkata, Biswanath Jana, Bose Institute, Kolkata, and Tuhin Subhra SarkarDepartment of Biochemistry, University of Calcutta, for technicalassistance. AM acknowledges the University Grant Commission forfellowship. The investigation was supported by a grant from theDepartment of Science and Technology (SR/SO/BB-67-2008), the Govt.of India and the CAS program of Department of Biochemistry, UniversityGrants Commission, India.

Appendix A. Supplementary data

Supplementary data associated with this article can be found, inthe online version, at doi:10.1016/j.virol.2010.07.032.

References

Albertini, A.A., Wernimont, A.K., Muziol, T., Ravelli, R.B., Clapier, C.R., Schoehn, G.,Weissenhorn, W., Ruigrok, R.W., 2006. Crystal structure of the rabies virusnucleoprotein–RNA complex. Science 313 (5785), 360–363.

Banerjee, A.K., Roy, J., Chattopadhyay, D.J., 1987. Effect of proteolytic digestion on thefunction of vesicular stomatitis virus ribonucleoprotein complex. J. Biosci. 11 (1–4),515–523.

Bankamp, B., Horikami, S.M., Thompson, P.D., Huber, M., Billeter, M., Moyer, S.A., 1996.Domains of the measles virus N protein required for binding to P protein and self-assembly. Virology 216 (1), 272–277.

Basak, S., Raha, T., Chattopadhyay, D., Majumder, A., Shaila, M.S., Chattopadhyay, D.J.,2003. Leader RNA binding ability of Chandipura virus P protein is regulated by itsphosphorylation status: a possible role in genome transcription–replication switch.Virology 307 (2), 372–385.

Basak, S., Polley, S., Basu, M., Chattopadhyay, D., Roy, S., 2004. Monomer and dimer ofChandipura virus unphosphorylated P-protein binds leader RNA differently:implications for viral RNA synthesis. J. Mol. Biol. 339 (5), 1089–1101.

Basak, S., Mondal, A., Polley, S., Mukhopadhyay, S., Chattopadhyay, D., 2007. ReviewingChandipura: a vesiculovirus in human epidemics. Biosci. Rep. 27 (4–5), 275–298.

Bhattacharya, R., Basak, S., Chattopadhyay, D.J., 2006. Initiation of encapsidation asevidenced by deoxycholate-treated Nucleocapsid protein in the Chandipura viruslife cycle. Virology 349 (1), 197–211.

Blumberg, B.M., Leppert, M., Kolakofsky, D., 1981. Interaction of VSV leader RNA andnucleocapsid protein may control VSV genome replication. Cell 23 (3), 837–845.

Page 10: Elucidation of functional domains of Chandipura virus Nucleocapsid protein involved in oligomerization and RNA binding: Implication in viral genome encapsidation

42 A. Mondal et al. / Virology 407 (2010) 33–42

Blumberg, B.M., Giorgi, C., Kolakofsky, D., 1983. N protein of vesicular stomatitis virusselectively encapsidates leader RNA in-vitro. Cell 32 (2), 559–567.

Buchholz, C.J., Spehner, D., Drillien, R., Neubert, W.J., Homann, H.E., 1993. The conservedN-terminal region of Sendai virus nucleocapsid protein NP is required fornucleocapsid assembly. J. Virol. 67 (10), 5803–5812.

Chadha, M.S., Arankalle, V.A., Jadi, R.S., Joshi, M.V., Thakare, J.P., Mahadev, P.V., Mishra,A.C., 2005. An outbreak of Chandipura virus encephalitis in the eastern districts ofGujarat state, India. Am. J. Trop. Med. Hyg. 73 (3), 566–570.

Chen, Z., Green, T.J., Luo, M., Li, H., 2004. Visualizing the RNAmolecule in the bacteriallyexpressed vesicular stomatitis virus nucleoprotein–RNA complex. Structure 12 (2),227–235.

Chen, M., Ogino, T., Banerjee, A.K., 2007. nteraction of vesicular stomatitis virus P and Nproteins: identification of two overlapping domains at the N terminus of P that areinvolved in N0–P complex formation and encapsidation of viral genome RNA. J.Virol. 81 (24), 13478–13485.

Elton, D., Medcalf, E., Bishop, K., Digard, P., 1999. Oligomerization of the influenza virusnucleoprotein: identification of positive and negative sequence elements. Virology260 (1), 190–200.

Ge, P., Tsao, J., Schein, S., Green, T.J., Luo, M., Zhou, Z.H., 2010. Cryo-EM model of thebullet-shaped vesicular stomatitis virus. Science 327 (5966), 689–693.

Green, T.J., Luo, M., 2009. Structure of the vesicular stomatitis virus nucleocapsid incomplex with the nucleocapsid-binding domain of the small polymerase cofactor,P. Proc. Natl. Acad. Sci. U. S. A. 106 (28), 11713–11718.

Green, T.J., Macpherson, S., Qiu, S., Lebowitz, J., Wertz, G.W., Luo, M., 2000. Study of theassembly of vesicular stomatitis virus N protein: role of the P protein. J. Virol. 74(20), 9515–9524.

Green, T.J., Zhang, X., Wertz, G.W., Luo, M., 2006. Structure of the vesicular stomatitisvirus nucleoprotein–RNA complex. Science 313 (5785), 357–360.

Gubbay, O., Curran, J., Kolakofsky, D., 2001. Sendai virus genome synthesis andassembly are coupled: a possible mechanism to promote viral RNA polymeraseprocessivity. J. Gen. Virol. 82 (Pt 12), 2895–2903.

Iseni, F., Barge, A., Baudin, F., Blondel, D., Ruigrok, R.W., 1998. Characterization of rabiesvirus nucleocapsids and recombinant nucleocapsid-like structures. J. Gen. Virol. 79(Pt 12), 2909–2919.

Kho, C.L., Tan, W.S., Tey, B.T., Yusoff, K., 2003. Newcastle disease virus nucleocapsid protein:self-assembly and length-determination domains. J. Gen. Virol. 84 (Pt 8), 2163–2168.

Lamb, R.A., Kolakofsky, D., 2001. Paramyxoviridae: the viruses and their replication, In:Knipe, D.M., Howley, P.M. (Eds.), Fields Virology, 4th ed. Lippincott Williams &Wilkins, Philadelphia.

Lesbats, P., Metifiot, M., Calmels, C., Baranova, S., Nevinsky, G., Andreola, M.L., Parissi, V.,2008. In-vitro initial attachment of HIV-1 integrase to viral ends: control of the DNAspecific interaction by the oligomerization state. Nucleic Acids Res. 36 (22), 7043–7058.

Liston, P., Batal, R., DiFlumeri, C., Briedis, D.J., 1997. Protein interaction domains of themeasles virus nucleocapsid protein (NP). Arch. Virol. 142 (2), 305–321.

Luo, M., Green, T.J., Zhang, X., Tsao, J., Qiu, S., 2007a. Conserved characteristics of therhabdovirus nucleoprotein. Virus Res. 129 (2), 246–251.

Luo, M., Green, T.J., Zhang, X., Tsao, J., Qiu, S., 2007b. Structural comparisons of thenucleoprotein from three negative strand RNA virus families. Virol. J. 4, 72.

Majumdar, A., Bhattacharya, R., Basak, S., Shaila, M.S., Chattopadhyay, D., Roy, S., 2004.P-protein of Chandipura virus is an N-protein-specific chaperone that acts at thenucleation stage. Biochemistry 43 (10), 2863–2870.

Majumder, A., Basak, S., Raha, T., Chowdhury, S.P., Chattopadhyay, D., Roy, S., 2001.Effect of osmolytes and chaperone-like action of P-protein on folding ofnucleocapsid protein of Chandipura virus. J. Biol. Chem. 276 (33), 30948–30955.

Marriott, A.C., 2005. Complete genome sequences of Chandipura and Isfahanvesiculoviruses. Arch. Virol. 150 (4), 671–680.

Masters, P.S., Banerjee, A.K., 1987. Sequences of Chandipura virus N and NS genes:evidence for high mutability of the NS gene within vesiculoviruses. Virology 157(2), 298–306.

Masters, P.S., Banerjee, A.K., 1988. Complex formation with vesicular stomatitis virusphosphoprotein NS prevents binding of nucleocapsid protein N to nonspecific RNA.J. Virol. 62 (8), 2658–2664.

Moyer, S.A., Smallwood-Kentro, S., Haddad, A., Prevec, L., 1991. Assembly andtranscription of synthetic vesicular stomatitis virus nucleocapsids. J. Virol. 65 (5),2170–2178.

Myers, T.M., Pieters, A., Moyer, S.A., 1997. A highly conserved region of the Sendai virusnucleocapsid protein contributes to the NP–NP binding domain. Virology 229 (2),322–335.

Myers, T.M., Smallwood, S., Moyer, S.A., 1999. Identification of nucleocapsid proteinresidues required for Sendai virus nucleocapsid formation and genome replication.J. Gen. Virol. 80 (Pt 6), 1383–1391.

Nishio, M., Tsurudome, M., Ito, M., Kawano, M., Kusagawa, S., Komada, H., Ito, Y., 1999.Mapping of domains on the human parainfluenza virus type 2 nucleocapsid protein(NP) required for NP–phosphoprotein or NP–NP interaction. J. Gen. Virol. 80 (Pt 8),2017–2022.

Pattnaik, A.K., Ball, L.A., LeGrone, A., Wertz, G.W., 1995. The termini of VSV DI particleRNAs are sufficient to signal RNA encapsidation, replication, and budding togenerate infectious particles. Virology 206 (1), 760–764.

Popuri, V., Bachrati, C.Z., Muzzolini, L., Mosedale, G., Costantini, S., Giacomini, E.,Hickson, I.D., Vindigni, A., 2008. The Human RecQ helicases, BLM and RECQ1,display distinct DNA substrate specificities. J. Biol. Chem. 283 (26), 17766–17776.

Rainsford, E.W., Harouoka, D., Wertz, G.W., 2010. Importance of hydrogen bondcontacts between the N protein and RNA genome of vesicular stomatitis virus inencapsidation and RNA synthesis. J. Virol. 84 (4), 1741–1751 Epub 2009 Dec 9.(Feb84(4)):1741-51.

Sarkar, A., Chattopadhyay, S., Cox, R., Luo, M., Banerjee, A.K., 2010. Structural andfunctional properties of the vesicular stomatitis virus nucleoprotein–RNA complexas revealed by proteolytic digestion. Virology 401 (1), 61–69.

Takemori, Y., Enoki, Y., Yamamoto, N., Fukai, Y., Adachi, K., Sakurai, H., 2009. Mutationalanalysis of human heat-shock transcription factor 1 reveals a regulatory role foroligomerization in DNA-binding specificity. Biochem. J. 424 (2), 253–261.

Tandale, B.V., Tikute, S.S., Arankalle, V.A., Sathe, P.S., Joshi, M.V., Ranadive, S.N., Kanojia,P.C., Eshwarachary, D., Kumarswamy, M., Mishra, A.C., 2008. Chandipura virus: amajor cause of acute encephalitis in children in North Telangana, Andhra Pradesh,India. J. Med. Virol. 80 (1), 118–124.

Tawar, R.G., Duquerroy, S., Vonrhein, C., Varela, P.F., Damier-Piolle, L., Castagne, N.,MacLellan, K., Bedouelle, H., Bricogne, G., Bhella, D., Eleouet, J.F., Rey, F.A., 2009.Crystal structure of a nucleocapsid-like nucleoprotein–RNA complex of respiratorysyncytial virus. Science 326 (5957), 1279–1283.

Thomas, D., Newcomb, W.W., Brown, J.C., Wall, J.S., Hainfeld, J.F., Trus, B.L., Steven, A.C.,1985. Mass and molecular composition of vesicular stomatitis virus: a scanningtransmission electron microscopy analysis. J. Virol. 54 (2), 598–607.

Yang, J., Hooper, D.C., Wunner, W.H., Koprowski, H., Dietzschold, B., Fu, Z.F., 1998. Thespecificity of rabies virus RNA encapsidation by nucleoprotein. Virology 242 (1),107–117.

Zhang, X., Green, T.J., Tsao, J., Qiu, S., Luo, M., 2008. Role of intermolecular interactions ofvesicular stomatitis virus nucleoprotein in RNA encapsidation. J. Virol. 82 (2),674–682.