Insights intothecoiled-coilorganizationoftheHendravirus...

14
Insights into the coiled-coil organization of the Hendra virus phosphoprotein from combined biochemical and SAXS studies Matilde Beltrandi a,b,1 , David Blocquel a,b,1 , Jenny Erales a,b , Pascale Barbier c , Andrea Cavalli d,e,n , Sonia Longhi a,b,nn a Aix-Marseille University, Architecture et Fonction des Macromolécules Biologiques (AFMB) UMR 7257, 13288 Marseille, France b CNRS, AFMB UMR 7257, 13288 Marseille, France c Aix-Marseille University, INSERM, CRO2 UMR_S911, Faculté de Pharmacie, 13385 Marseille, France d Institute for Research in Biomedicine, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland e Department of Chemistry, University of Cambridge, Cambridge CB2 1EW, United Kingdom article info Article history: Received 20 September 2014 Returned to author for revisions 21 October 2014 Accepted 19 December 2014 Keywords: Henipavirus Hendra virus Phosphoprotein PMD P multimerization domain Coiled-coil Homotrimer Small-angle X-ray scattering Cross-linking abstract Nipah and Hendra viruses are recently emerged paramyxoviruses belonging to the Henipavirus genus. The Henipavirus phosphoprotein (P) consists of a large intrinsically disordered domain and a C-terminal domain (PCT) containing alternating disordered and ordered regions. Among these latter is the P multimerization domain (PMD). Using biochemical, analytical ultracentrifugation and small-angle X-ray scattering (SAXS) studies, we show that Hendra virus (HeV) PMD forms an elongated coiled-coil homotrimer in solution, in agreement with our previous ndings on Nipah virus (NiV) PMD. However, the orientation of the N-terminal region differs from that observed in solution for NiV PMD, consistent with the ability of this region to adopt different conformations. SAXS studies provided evidence for a trimeric organization also in the case of PCT, thus extending and strengthening our ndings on PMD. The present results are discussed in light of conicting reports in the literature pointing to a tetrameric organization of paramyxoviral P proteins. & 2014 Elsevier Inc. All rights reserved. Introduction The Hendra (HeV) and Nipah (NiV) viruses are recently emerged BSL4 pathogens gathered within the Henipavirus genus within the Paramyxoviridae family (Eaton et al., 2007; Wang et al., 2000). The newly identied Cedar virus has also been classied within this genus (Marsh et al., 2012). In henipaviruses, the genome is encapsi- dated by the nucleoprotein (N) within a helical nucleocapsid that is the substrate used by the polymerase for both transcription and replication. The viral polymerase is a complex made of the L protein and the P protein. The latter is an essential polymerase cofactor as it allows the recruitment of L onto the nucleocapsid template. Beyond its role in tethering the L protein, P is also supposed to serve as a chaperone for N, i.e. to prevent illegitimate self-assembly of N (see Albertini et al. (2005), Blocquel et al. (2012a), Lamb and Parks (2007) and Roux (2005) for reviews on paramyxovirus transcription and replication). Henipavirus N and P proteins were shown to interact with each other (Habchi et al., 2011), being able to form both homologous and heterologous NP complexes (Blocquel et al., 2012b; Chan et al., 2004; Omi-Furutani et al., 2010). Using both computational and experimental approaches, we have previously shown that Henipavirus P proteins possess a modular Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/yviro Virology http://dx.doi.org/10.1016/j.virol.2014.12.029 0042-6822/& 2014 Elsevier Inc. All rights reserved. Abbreviations: HeV, Hendra virus; NiV, Nipah virus; N, nucleoprotein; L, large protein; P, phosphoprotein; IDP, intrinsically disordered protein; IDR, intrinsically disordered region; PNT, P N-terminal domain; PCT, P C-terminal domain; PMD, P multimerization domain; XD, X domain of P; MeV, measles virus, SeV, Sendai virus; MuV, mumps virus; RDV, Rinderpest virus; RSV, respiratory syncytial virus; NMR, nuclear magnetic resonance; PCR, polymerase chain reaction; OD, optical density; IMAC, immobilized metal afnity chromatography; GF, gel ltration; AUC, analy- tical ultracentrifugation; SEC, size-exclusion chromatography; SDS-PAGE, sodium dodecyl sulfate polyacrylamide electrophoresis; R s , Stokes radius; R g , radius of gyration; MALDITOF, matrix-assisted laser desorption ionization/time of ight; CD, circular dichroism; TFE, 2,2,2-triuoroethanol; MRE, mean residue ellipticity values per residue; SAB, suberic acid bis (N-hydroxy-succinimide ester); SAXS, small angle X-ray scattering; T m , melting temperature n Corresponding to: Institute for Research in Biomedicine, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland. Tel.: þ41 0 91 820 0303. nn Correspondence to: AFMB, UMR 7257 CNRS and Aix-Marseille University, 163, avenue de Luminy, Case 932, 13288 Marseille Cedex 09, France. Tel.: þ33 4 91 82 55 80; fax: þ33 4 91 26 67 20. E-mail addresses: [email protected] (A. Cavalli), [email protected] (S. Longhi). 1 These authors have equally contributed to the work. Virology 477 (2015) 4255

Transcript of Insights intothecoiled-coilorganizationoftheHendravirus...

Insights into the coiled-coil organization of the Hendra virusphosphoprotein from combined biochemical and SAXS studies

Matilde Beltrandi a,b,1, David Blocquel a,b,1, Jenny Erales a,b, Pascale Barbier c,Andrea Cavalli d,e,n, Sonia Longhi a,b,nn

a Aix-Marseille University, Architecture et Fonction des Macromolécules Biologiques (AFMB) UMR 7257, 13288 Marseille, Franceb CNRS, AFMB UMR 7257, 13288 Marseille, Francec Aix-Marseille University, INSERM, CRO2 UMR_S911, Faculté de Pharmacie, 13385 Marseille, Franced Institute for Research in Biomedicine, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerlande Department of Chemistry, University of Cambridge, Cambridge CB2 1EW, United Kingdom

a r t i c l e i n f o

Article history:Received 20 September 2014Returned to author for revisions21 October 2014Accepted 19 December 2014

Keywords:HenipavirusHendra virusPhosphoproteinPMDP multimerization domainCoiled-coilHomotrimerSmall-angle X-ray scatteringCross-linking

a b s t r a c t

Nipah and Hendra viruses are recently emerged paramyxoviruses belonging to the Henipavirus genus.The Henipavirus phosphoprotein (P) consists of a large intrinsically disordered domain and a C-terminaldomain (PCT) containing alternating disordered and ordered regions. Among these latter is the Pmultimerization domain (PMD). Using biochemical, analytical ultracentrifugation and small-angle X-rayscattering (SAXS) studies, we show that Hendra virus (HeV) PMD forms an elongated coiled-coilhomotrimer in solution, in agreement with our previous findings on Nipah virus (NiV) PMD. However,the orientation of the N-terminal region differs from that observed in solution for NiV PMD, consistentwith the ability of this region to adopt different conformations. SAXS studies provided evidence for atrimeric organization also in the case of PCT, thus extending and strengthening our findings on PMD. Thepresent results are discussed in light of conflicting reports in the literature pointing to a tetramericorganization of paramyxoviral P proteins.

& 2014 Elsevier Inc. All rights reserved.

Introduction

The Hendra (HeV) and Nipah (NiV) viruses are recently emergedBSL4 pathogens gathered within the Henipavirus genus within theParamyxoviridae family (Eaton et al., 2007; Wang et al., 2000). Thenewly identified Cedar virus has also been classified within thisgenus (Marsh et al., 2012). In henipaviruses, the genome is encapsi-dated by the nucleoprotein (N) within a helical nucleocapsid that isthe substrate used by the polymerase for both transcription andreplication. The viral polymerase is a complex made of the L proteinand the P protein. The latter is an essential polymerase cofactor as itallows the recruitment of L onto the nucleocapsid template. Beyondits role in tethering the L protein, P is also supposed to serve as achaperone for N, i.e. to prevent illegitimate self-assembly of N (seeAlbertini et al. (2005), Blocquel et al. (2012a), Lamb and Parks (2007)and Roux (2005) for reviews on paramyxovirus transcription andreplication). Henipavirus N and P proteins were shown to interactwith each other (Habchi et al., 2011), being able to form bothhomologous and heterologous N–P complexes (Blocquel et al.,2012b; Chan et al., 2004; Omi-Furutani et al., 2010).

Using both computational and experimental approaches, we havepreviously shown that Henipavirus P proteins possess a modular

Contents lists available at ScienceDirect

journal homepage: www.elsevier.com/locate/yviro

Virology

http://dx.doi.org/10.1016/j.virol.2014.12.0290042-6822/& 2014 Elsevier Inc. All rights reserved.

Abbreviations: HeV, Hendra virus; NiV, Nipah virus; N, nucleoprotein; L, largeprotein; P, phosphoprotein; IDP, intrinsically disordered protein; IDR, intrinsicallydisordered region; PNT, P N-terminal domain; PCT, P C-terminal domain; PMD, Pmultimerization domain; XD, X domain of P; MeV, measles virus, SeV, Sendai virus;MuV, mumps virus; RDV, Rinderpest virus; RSV, respiratory syncytial virus; NMR,nuclear magnetic resonance; PCR, polymerase chain reaction; OD, optical density;IMAC, immobilized metal affinity chromatography; GF, gel filtration; AUC, analy-tical ultracentrifugation; SEC, size-exclusion chromatography; SDS-PAGE, sodiumdodecyl sulfate polyacrylamide electrophoresis; Rs, Stokes radius; Rg, radius ofgyration; MALDI–TOF, matrix-assisted laser desorption ionization/time of flight;CD, circular dichroism; TFE, 2,2,2-trifluoroethanol; MRE, mean residue ellipticityvalues per residue; SAB, suberic acid bis (N-hydroxy-succinimide ester); SAXS,small angle X-ray scattering; Tm, melting temperature

n Corresponding to: Institute for Research in Biomedicine, Via Vincenzo Vela 6,6500 Bellinzona, Switzerland. Tel.: þ41 0 91 820 0303.

nn Correspondence to: AFMB, UMR 7257 CNRS and Aix-Marseille University, 163,avenue de Luminy, Case 932, 13288 Marseille Cedex 09, France.Tel.: þ33 4 91 82 55 80; fax: þ33 4 91 26 67 20.

E-mail addresses: [email protected] (A. Cavalli),[email protected] (S. Longhi).

1 These authors have equally contributed to the work.

Virology 477 (2015) 42–55

organization consisting of both ordered and intrinsically disorderedregions (IDRs) (Habchi et al., 2010; Karlin et al., 2003). Intrinsicallydisordered proteins (IDPs) or IDRs are functional proteins/regions thatlack highly populated uniform secondary and tertiary structure underphysiological conditions of pH and salinity in the absence of a partner(for reviews on IDPs, see (Chouard, 2011; Dunker et al., 2008a, 2008b;Habchi et al., 2014; Turoverov et al., 2010; Uversky, 2010, 2013;Uversky and Dunker, 2010)). The Henipavirus P proteins contain infact a spectacularly long N-terminal disordered region (PNT, aa 1–404in HeV) (Habchi et al., 2010) and a C-terminal moiety (PCT, aa 405–707in HeV) (Fig. 1). Contrary to PNT, which is common to both P and theauxiliary viral protein V, PCT is unique to the P protein. PCT has amodular organization being composed of alternating disordered andordered regions: it embraces a predicted disordered region that partlyoverlaps the V ORF (aa 405–468 in HeV) referred to as spacer, astructured region (aa 469–578 in HeV) referred to as PMD (for Pmultimerization domain), a disordered linker (aa 579–656 in HeV) anda globular region (aa 657–707 in HeV) referred to as X domain (XD)(Fig. 1) (Habchi et al., 2010). The crystal structure of HeV XD has beensolved recently and was shown to consist of a triple α-helical bundle(Communie et al., 2013b), in agreement with previous spectroscopicand modeling studies (Habchi et al., 2011).

In Paramyxovirinae, sequence analyses predict a coiled-coil regionwithin PMD (Habchi et al., 2010; Karlin et al., 2003) (see Fig. 1). Thecoiled-coil organization has been experimentally confirmed in the caseof SeV (Tarbouriech et al., 2000b), rinderpest virus (RDV) (Rahamanet al., 2004), MeV (Blocquel et al., 2014; Communie et al., 2013a), MuV(Cox et al., 2013) and NiV (Blocquel et al., 2013; Bruhn-Johannsen et al.,2014). PMDs from SeV (Tarbouriech et al., 2000b) and MeV wereshown to form a tetrameric coiled-coil (Blocquel et al., 2014;Communie et al., 2013a). The tetrameric coiled-coil organization ofPMD has also been experimentally confirmed in the case of RDV(Rahaman et al., 2004), RSV (Llorente et al., 2006, 2008) and MuV (Coxet al., 2013). In this latter case, however, the tetramer was found toconsist of two sets of parallel helices in opposite orientation, i.e. to be adimer of two antiparallel coiled-coil dimers (Cox et al., 2013) instriking contrast with all Paramyxoviridae P proteins characterized sofar that were all shown to possess a parallel organization.

Strikingly, studies focused on NiV PMD yielded different resultsdepending on whether the protein was studied in solution or by

X-ray crystallography. Indeed, several independent biochemicaland biophysical approaches, including size-exclusion chromato-graphy (SEC), SDS-PAGE, cross-linking, analytical ultracentrifuga-tion and small-angle X ray scattering (SAXS), consistently conve-rged to show that NiV PMD adopts a trimeric organization insolution (Blocquel et al., 2013). By contrast, the crystallographicstructure of NiV PMD reported by Bruhn et al. revealed a tetra-meric organization (Bruhn-Johannsen et al., 2014), thus raisinginteresting questions as to the origin of these discrepancies.

In order to solve this conundrum and to shed light onto thestructural organization of Henipavirus phosphoproteins, we hereinhave undertaken the characterization of both HeV PMD and PCTregions. Using a combination of biochemical and biophysicalapproaches, we show that like in the case of NiV, HeV PMD formsa homotrimeric coiled-coil structure with an overall elongatedshape. Notably, SAXS studies provided evidence for a trimericorganization also in the case of PCT, thus extending and strength-ening conclusions based on PMD.

Results and discussion

Cloning, expression and purification of P constructs

The P gene fragments encoding PMD or PCT with a C-terminalhistidine-tag were cloned into pDEST14. In both cases, the recom-binant product was purified from the soluble fraction of thebacterial lysate (see insets in Fig. 2). The proteins were purifiedin two steps: IMAC and gel filtration (see insets in Fig. 2). BothPMD and PCT were found to migrate in SDS-PAGE with anapparent molecular mass higher than that expected from theiramino acid sequence: indeed PMD was found to have an apparentmolecular mass close to 17 kDa (expected mass of 13,524 Da) andPCT a mass of approximately 40 kDa (expected mass of 34,913 Da)(see insets in Fig. 2). In both cases, however, the identity of thepurified protein was confirmed by mass spectrometry analysis ofthe tryptic fragments obtained after digestion of the purifiedprotein band excised from SDS-polyacrylamide gels. This aberrantelectrophoretic migration may reflect either an unusual amino

PCTPNT

XDspacer1

PMD

538 548478469 558528518508488 498 578568

Fig. 1. Domain organization of HeV P. Domain organization of P showing that it is composed of two moieties, PNT (aa 1–404) and PCT (aa 405–707), with this consisting of adisordered region (“spacer”, aa 405–468), a structured region (PMD, aa 469–578), a disordered linker (aa 579–656) and a globular region (XD, aa 657–707). The ribbonrepresentation of the crystal structure of HeV XD (pdb code 4HEO) (Communie et al., 2013b) is shown. The structure was drawn using Pymol (DeLano, 2002). The HCA plot ofPMD, displayed on the top, points out the presence of a region encompassing residues 512–578 (underlined by a green bar) with the typical texture of a coiled-coil region(see Ferron et al. (2005) and Blocquel et al. (2013)).

M. Beltrandi et al. / Virology 477 (2015) 42–55 43

acid sequence and/or composition, or the presence of predicteddisordered regions.

Intriguingly, and as already observed for the cognate NiV PMD(Blocquel et al., 2013), when HeV PMD was analyzed by SDS-PAGE,an additional band of approximately 34 kDa was observed (Fig. 3A).This band was no longer detected when the protein was subjectedto prolonged thermal denaturation (Fig. 3A). Furthermore, thisadditional band was found to be resistant to the addition of20 mM DTT (Fig. 3A), arguing for a very stable oligomer that resistsunder denaturing conditions (i.e. in spite of the presence of SDS)and that is not mediated by disulfide bridges.

Size-exclusion chromatography and cross-linking studies

SEC experiments yielded an apparent molecular mass of 43 kDa forHeV PMD, with the elution profile being independent from proteinconcentration (Fig. 2A). This value is not consistent with a monomericform of the protein (13.5 kDa) and is rather close to the value expectedfor a trimer (40.5 kDa). Further support to this conclusion comes fromthe observation that HeV PMD has the same elution profile of NiVPMD (Fig. 2A), a protein that has the same molecular mass as HeVPMD and has already been shown to be an elongated trimer (Blocquelet al., 2013).

120

RS: 51.0 ÅMM: 207 kDa

mAU

SN fl-th

ENiM EGF

141825354566116

0 20 40 60 80 mL 0 20 40 60 80 mL

HeV PMD NiV PMD

RS: 28.5 ÅMM: 43 kDa50

0

40

30

20

10

60

30

90

mAU

0.0

TF

Fig. 2. Gel filtration elution profile and purification of HeV PMD (A) and PCT (B). The arrows point the peak containing the protein of interest. The corresponding apparentmolecular mass (MM) and Stokes radius (RS) are indicated. In panel A, the elution profile of NiV PMD (blue) is overlaid. Insets: Coomassie blue staining of an SDS-PAGEshowing the bacterial lysate (total fraction, TF), the clarified supernatant (soluble fraction, SN), the non-retained fraction (fl-th), the eluent from Immobilized Metal AffinityChromatography (ENi) and the eluent from Gel Filtration (EGF). M: molecular mass markers (kDa).

M

monomer

dimer

DTT Therm. den.

66.4

55.6

42.7

34.6

27

2017

26

34

43

957255

0 10 33 100 333 1000 3333 10000M

trimer

Fig. 3. Effect of thermal denaturation and cross-linking experiments. (A) Coomassie stained 15% SDS-PAGE showing the electrophoretic migration of purified HeV PMD asobtained either with or without a prolonged thermal denaturation at 95 1C. The migration of a HeV PMD sample not subjected to thermal denaturation and containing20 mM DTT is also shown. (B) Coomassie stained 15% SDS-PAGE showing the results of cross-linking with SAB (see Materials and methods). The concentration of SAB (μM) isindicated above the gel. Samples in panel B, were thermally denatured for 5 minutes prior to electrophoretic migration.

M. Beltrandi et al. / Virology 477 (2015) 42–5544

The Stokes radius (Rs) of HeV PMD, as inferred from itsapparent molecular mass (Uversky, 2002), is 28.5 Å, a value closeto that expected for a globular trimer (27.9 Å) (Uversky, 2002). Theslightly higher Rs value with respect to the value expected for aglobular trimer could reflect an elongated form of the protein, asalready observed for SeV, RDV and NiV PMDs (Blocquel et al., 2013;Rahaman et al., 2004; Tarbouriech et al., 2000a).

In the case of PCT, SEC analyses showed that the proteinwas elutedin the void volume (data not shown) unless 0.2 M arginine was addedto both the protein and the elution buffer (Fig. 2B). In this latter case,its elution profile was borderline with respect to the exclusion volumeallowing only a rough estimation of themolecular mass. The estimatedvalue (207 kDa) is twice the value expected for a trimer and 1.5 timesgreater than the value expected for a tetramer, which can be indicativeeither of a higher oligomeric state or of an elongated shape combinedto the presence of disordered regions that are known to impart an

apparently higher molecular mass during SEC (Uversky, 2002). Thecorresponding RS is 51.0 Å, a value much higher than those expectedfor a trimeric or tetrameric globular form of the protein (39.6 Å and44.1 Å, respectively), again suggesting that the protein either formshigher-order oligomers and/or has an extended shape.

In order to shed light into the oligomeric state of PMD and PCT,we performed cross-linking experiments using SAB, a bifunctionalreagent of fixed size (13.1 Å) that reacts with lysines. Experimentscarried out with PCT failed to provide insights, as addition of cross-linker resulted in the formation of aggregates of high molecularmass that could not enter into the gel (data not shown). In the caseof PMD, the addition of increasing amounts of SAB triggers theprogressive accumulation of two additional bands, the apparentmolecular mass of which is close to 34 and 55 kDa (Fig. 3B). Takinginto account the fact that the monomeric form aberrantly migratesas a 17-kDa protein, these two bands, which correspond to twice

1.5

1

0.5

0

-0.5

-1

0.5 1 1.5 2 2.5 3 3.5

In I(

q)

0.06

0.04

0.02

00 6

P(r

)

r(nm)

120

1

2

3

4

5

I(q)*

q2

0.02 0.04 0.06

In I(

q)

q2 (nm-2)

0.08

q (nm-1)

q (nm-1)

0 0.5 1 1.5 2 2.5 32 8 10

132 Å

90°

0.77 g/L 1.36 g/L 3.31 g/L

180°

4

3.8

3.6

3.4

34 Å

Fig. 4. Small-angle X-ray scattering experiments of HeV PMD. (A) Experimental SAXS data recorded for q values up to 3.5 nm�1. The curves obtained for three proteinconcentrations (0.77 g/L, blue; 1.36 g/L, red; and 3.31 g/L, light blue) are represented after correction for concentration. (B) Representation of the Guinier plot for the proteinat 3.31 g/L. (C) Pair distance distribution, P(r), function of the data for the 3.31 g/L concentration. (D) Kratky plot. (E) Three views of the ab initio envelope calculated withDAMAVER (Volkov and Svergun, 2003). 20 DAMMIF calculations (Franke and Svergun, 2009) were performed and averaged with DAMAVER to produce the average andfiltered shape shown in light blue. The structure of a trimeric coiled-coil model of HeV PMD is shown, with the three chains being displayed in three different colors. Dockingof the model in the envelope was done manually using the program Chimera. The structural model fits well in the density of the SAXS-derived model.

M. Beltrandi et al. / Virology 477 (2015) 42–55 45

and three times the apparent mass of the monomer, couldcorrespond to a dimeric and trimeric form (Fig. 3B). Surprisingly,the band close to 34 kDa is also discernible in the absence of SAB(Fig. 3B), a finding in line with the above described propensity ofHeV PMD to give rise to a very stable (i.e. SDS-resistant) oligomerunless subjected to prolonged thermal denaturation (Fig. 3A).

The lack of detection of higher order oligomers even at thehighest cross-linker concentration suggests that the protein doesnot form tetramers. In this regard, it should be pointed out thatalthough cross-linking experiments carried out by another groupand making use of a different cross-linker (i.e. glutaraldehyde)revealed a tetrameric organization for NiV PMD (Salvamani et al.,2013), the very high cross-linker concentrations used in thosestudies may have generated non-specific association.

Although the present cross-linking experiments on their own arenot sufficient to draw definite conclusions about the oligomeric stateof HeV PMD, they suggest that the protein could adopt a trimeric statein solution like its NiV counterpart. In further support of thishypothesis, it is noteworthy that when similar experiments wereperformed with tetrameric MuV, RPV SeV and MeV PMDs, a clear anddiscrete band corresponding to the tetrameric form could be readilydetected, even in our hands (Blocquel et al., 2014; Cox et al., 2013;Rahaman et al., 2004; Tarbouriech et al., 2000b).

Far-UV CD studies

In order to confirm that HeV PMD adopts a coiled-coil con-formation like in the case of its most related paramyxoviralmembers, we used far-UV spectroscopy. The spectrum of PMD isindicative of a folded protein with an estimated α-helical contentof 61% (see inset in Supplementary Fig. S1A). The ratio of theellipticities at 222 and 208 nm (Θ222/208) is greater than 1.0 (seeinset in Supplementary Fig. S1A), a property indicative of thepresence of interacting helices and already observed also in thecase of PMDs from SeV (Tarbouriech et al., 2000a), RDV (Rahamanet al., 2004), MeV (Blocquel et al., 2014) and NiV (Blocquel et al.,2013). Upon addition of 70% of TFE the Θ222/208 ratio drops below 1(Supplementary Fig. S1A). Taking into account the fact that high

TFE concentrations are known to disrupt tertiary structure andquaternary structure and to stabilize secondary structure (Lauet al., 1984), these results indicate that HeV PMD forms oligomersthrough coiled-coil interactions.

To investigate the thermal stability of PMD oligomers, werecorded the CD spectrum at 100 1C along with that obtained upona stepwise cooling down to 20 1C (Supplementary Fig. S1A). Thegood superimposition between the initial spectrum and the spec-trum obtained after cooling points out the reversibility of thermalunfolding and argues for refolding of the main structural motif (i.e.coiled-coil) (Supplementary Fig. S1A). We also monitored the meanresidue ellipticity at 222 nm (MRE222), indicative of the α-helicalcontent, at increasing temperatures and plotted it as a function oftemperature (Supplementary Fig. S1B). The unfolding profile showsa single cooperative transition with an inflection point at approxi-mately 65 1C (Supplementary Fig. S1B). Analysis of the thermalunfolding profile yielded an apparent melting temperature (Tm) of66 1C, a value lower than that determined for NiV PMD (76 1C)(Blocquel et al., 2013).

These results support the conclusion that HeV PMD forms acoiled-coil oligomer that features a relatively high thermal stabi-lity, in line with previous observations on related PMDs (Blocquelet al., 2013, 2014; Communie et al., 2013a; Llorente et al., 2008).

The far-UV CD spectrum of PCT is indicative of a proteincontaining a higher amount of disorder and a lower α-helicalcontent as compared to PMD (Supplementary Fig. S1C and inset),in agreement with the presence of predicted disordered regions. Inthe case of PCT, the Θ222/208 ratio is lower than 1.0, a valuereflecting the absence of interacting helices. In addition, and bycontrast with HeV PMD, the addition of TFE does not significantlyaffect this ratio (see inset of Supplementary Fig. S1C), againsuggesting the lack of interacting helices through a quaternarystructure organization. The lower than 1.0 value of Θ222/208

probably reflects the contribution of disordered regions, whichare typified by very low ellipticities values at 222 nm (Woody,2010). Thermal denaturation experiments similar to thosedescribed above, yielded comparable results in that thermalunfolding was found to be reversible, thus ruling out possibleheat-induced protein aggregation and arguing for proper refolding

Table 1SAXS data collection and scattering-derived structural parameters (at the highest concentration) for PMD and PCT.

PMD PCT

Data-collection parametersDetector Pilatus (1 M) Pilatus (1 M)Beam geometry Bending magnet (BM29) Bending magnet (BM29)Wavelength (Å) 0.992 0.992q Range (Å�1) 0.028–4.525 0.028–4.525Exposure time (s) 1 1Concentration range (g/L) 0.77–3.31 0.44–2.48Temperature (1C) 10 10

Structural parametersI(0) (cm�1) (from P(r)) 42.070.1 115.070.2I(0) (cm�1) (from Guinier) 41.670.3 114.570.3Rg (Å) (from P(r)) 39.170.2 86.770.5Rg (Å) (from Guinier) 37.270.1 87.070.1Dmax (Å) 132 302

Molecular-mass determination (kDa)Molecular mass (MM) (from I(0)) 41.0 114Calculated MM from sequence 40.5 104.7

Software employedPrimary data reduction PRIMUS PRIMUSData processing GNOM GNOMAb initio analysis DAMMIF DAMMIFValidation and averaging DAMAVER DAMAVERValidation of structural models CRYSOL/FOXS CRYSOL/FOXS

M. Beltrandi et al. / Virology 477 (2015) 42–5546

of the structural motifs of PCT (i.e. PMD and XD). Plotting theMRE222 as a function of temperature yielded a typical two-stateprofile, indicative of a single cooperative transition with anapparent Tm of 68.6 1C (Supplementary Fig. S1D). These resultssupport the conclusion that HeV PCT encompasses both disorderedand structured regions adopting a prevalently α-helical conforma-tion. Its behavior in terms of reversibility of thermal unfolding andhigh Tm are reminiscent of that observed in the case of HeV PMD,suggesting that the folding/unfolding behavior of PCT is domi-nated by PMD and that this latter maintains its structural organi-zation in the context of PCT.

SAXS studies of PMD

In order to achieve insights into the shape and oligomeric state ofHeV PMD in solution, we performed SAXS studies. The shapes of the

SAXS curves (Fig. 4A) and the Guinier plots obtained (Fig. 4B) areindependent of protein concentration, indicating the absence ofsignificant aggregation. Each curve can be well approximated by astraight line in the Guinier region (qRgo1.0). The slope gives the valueof the radius of gyration, Rg, while the intercept of the straight linegives the I(0) which is proportional to the molecular mass of thescatterer. Guinier analysis in the low q region gave an Rg of37.270.1 Å, which is in good agreement with the value of 39.170.2determined from the pair distance distribution function P(r) (Tables 1and 2). The molecular mass determined from the extrapolatedscattering intensity at zero angle I(0) is 41.0 kDa, a value in very goodagreement with the molecular mass expected for a trimeric form(40.5 kDa) (Table 1). Note that the reliability of this estimation issupported by the precision with which we could estimate proteinconcentration. Indeed, accurate determination of molecular mass bySAXS analysis is known to critically depend on accurate estimation ofparticle concentration (Mylonas and Svergun, 2007). Herein, ourestimation of HeV PMD concentration based on the theoreticalabsorption coefficient was found to be in very good agreement withthe actual protein concentration determined through analysis of theamino acid composition (see Materials and methods).

The experimentally observed Rg value is 1.7 times larger thanthat expected (22.1 Å) for a globular protein with an Rs equal tothat experimentally observed for HeV PMD (28.5 Å) (Wilkins et al.,1999). It is even much higher than the value expected (17.3 Å)for a sphere with an Rs of 22.4 Å, as determined from the volumeof a sphere with 117�3 residues (see Materials and methods).The strong discrepancy between the experimentally observed Rgand the value expected for a globular/spherical form indicates that

Table 2Rg (from Guinier), I(0) (from Guinier) and Dmax for PMD and PCT at the variousprotein concentrations.

Protein concentration (g/L) Rg (Å)(Guinier)

I(0) (Guinier)(cm�1)

Dmax

(Å)

PMD 0.77 36.570.3 41.1 127.6PMD 1.36 37.070.1 41.8 130.8PMD 3.31 37.270.1 41.6 132.3PCT 0.44 86.470.5 114.4 300.8PCT 1.25 86.870.2 114.7 301.1PCT 2.48 87.070.1 114.5 301.8

LgI

2.5

2

1.5

0.5

0

1

0 0.250.20.150.10.05

Exp Fit Re1Err

P1

s=4pi*sin(θ)/λ

LgI

2.5

2

1.5

0.5

0

1

Exp Fit Re1Err

P2

0 0.250.20.150.10.05

s=4pi*sin(θ)/λ

LgI

2.5

2

1.5

0.5

0

1

0 0.250.20.150.10.05

s=4pi*sin(θ)/λ

Exp Fit Re1Err

P3

LgI

2.5

2

1.5

0.5

0

1

0 0.250.20.150.10.05

s=4pi*sin(θ)/λ

Exp +++Fit ------Re1Err °°°°°°

P4

0.77 g/L 1.041.36 g/L 1.643.31 g/L 1.11

0.77 g/L 1.321.36 g/L 1.373.31 g/L 1.24

0.77 g/L 0.911.36 g/L 1.323.31 g/L 1.02

0.77 g/L 1.431.36 g/L 2.393.31 g/L 1.56

Fig. 5. Comparison between the experimental (green) and calculated (red) HeV PMD scattering curve obtained either with no forced symmetry (P1) or by imposing a P2, P3or P4 symmetry. The calculated curve corresponds to the average profile provided by DAMMIF. Experimental data presented here were recorded at 3.31 g/L. Values in theinsets show the χ values calculated for the comparison between the predicted SAXS profiles and the experimental data as obtained at different protein concentrations.

M. Beltrandi et al. / Virology 477 (2015) 42–55 47

the overall structure of HeV PMD is very elongated. The higherthan expected Rg confirms that the higher apparent molecularmass observed upon SEC does indeed reflect an elongated trimer,as already observed for the cognate NiV PMD (Blocquel et al.,2013).

The distances distribution function inferred from the scatteringcurve of HeV PMD exhibits a maximum at 20 Å, a shoulder at100 Å and a long tail up to 132 Å (Dmax) typical of an elongatedobject (Fig. 4C). The Rg of a thin rod of length L¼132 Å is 38.1 Å(see Materials and methods), a value in very good agreement withthe values estimated from the Guinier plot (37.2 Å) and from thepair distance distribution function (39.1 Å).

The Kratky plot presents a maximum at qE0.83 nm�1 and a flatregion for q41.75 nm�1 (Fig. 4D). The shape of the plot isindicative of a structured protein with possibly an at least partlydisordered appendage, as judged from the bell-shape nature of thecurve that displays a clear maximum and from the presence of theflat region.

Next, we employed the program DAMMIF to carry out ab initioshape reconstruction from the SAXS data. To this purpose, we used thescattering data obtained at the highest concentration so as to achievemaximal resolution. Several series of independent runs were carriedout either with no forced symmetry or by imposing P2, P3 or P4symmetries. Within each symmetry class, models were very repro-ducible with an average normalized spatial discrepancy (NSD) well

below 1.0, indicating structurally similar solutions (Supplementary Fig.S2). The lowest average NSD values were obtained either with noimposed symmetry (NSD¼0.5870.02) or by imposing a P3 symmetry(NSD¼0.6270.64). All generated bead models appeared as elongatedcylinders, compatible with a coiled-coil structure (Supplementary Fig.S3 and data not shown). Supplementary Fig. S4 shows, for eachsymmetry class, the distribution of χ values for the 20models. For eachsymmetry class, the models resulting from 20 independent DAMMIFruns were superimposed using the DAMAVER suite. After havingrejected three and two outliers (see Materials and methods) withinthe P2 and P3 symmetry class, respectively, the program built the finalaverage filteredmodels that were all found to have an elongated shape(Supplementary Fig. S5). Best results were obtained with the P3symmetry, as judged from the DAMMIF average χ parameter (Fig. 5).Of note, best average χ values were systematically obtained byimposing a P3 symmetry even at the intermediate and low proteinconcentration (see insets in Fig. 5). Global statistical analysis of theobtained χ values using Bayesian-Factors (see Materials and methods)showed that differences in χ values are meaningful (SupplementaryTable S1), thereby supporting the conclusion that the P3 symmetry isthe one that best fits the experimental data. As shown in Fig. 4E, theaverage filtered model generated by DAMAVER by imposing the P3symmetry has an elongated dumbbell shape.

Interestingly, the shape of the HeV PMD envelope is somehowdifferent from that of NiV PMD (Blocquel et al., 2013). Another

1

0

-1

-2

1

0

-1

-2

Lgl

Lgl

1

0

-1

-2

1

0

-1

-2

Lgl

Lgl

1

0

-1

-2

1

0

-1

-2

Lgl

Lgl

0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4

0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4

s=4pi*sin(θ)/λ s=4pi*sin(θ)/λ s=4pi*sin(θ)/λ

s=4pi*sin(θ)/λ s=4pi*sin(θ)/λ s=4pi*sin(θ)/λ

Fig. 6. Comparison of experimental (red) and predicted (green) SAXS data generated by CRYSOL for a dimeric, trimeric and tetrameric model of HeV PMD with the “head”either in the “down” (A) or “up” conformation (B). Experimental data presented here were recorded at 3.31 g/L. Values in the insets show the χ values calculated for thecomparison between the predicted SAXS profiles of the models and the experimental data as obtained at different protein concentrations (0.77, 1.36 and 3.31 g/L). In allpanels models are shown embedded in the ab initio SAXS envelope. Docking of the models in the envelope was done manually using the program Chimera.

M. Beltrandi et al. / Virology 477 (2015) 42–5548

unexpected difference between the two PMDs resides in theiroverall dimensions: indeed, although the two domains differ inlength by only one residue, HeV PMD was found to have an Rg anda Dmax much smaller than those of its NiV counterpart (compare37.2 Å and 132 Å, respectively, to 50.8 Å and 182 Å) (Blocquel et al.,2013). Taking into account the high sequence identity (91%) andthe highly similar size of the two P multimerization domains,these results indicate that the amino acid sequence does notuniquely dictates the structure adopted in solution, thereby lead-ing to unexpected differences in the structural parameters. Thesestructural discrepancies between the envelopes of NiV and HeVPMD may reflect the ability of Henipavirus PMDs to undergoconformational changes resulting in forms of different lengths.

Similarly to what we did in the case of NiV PMD (Blocquel et al.,2013), we generated various structural models of HeV PMD differing intheir oligomeric state (see Materials and methods) and assessed howthey fit into the SAXS envelope. As shown in Fig. 4E, the trimericmodel can easily be accommodated in the SAXS envelope. To quantifythe extent to which the different models fit the experimental scatter-ing profiles, we used the program CRYSOL (Svergun et al., 1995), whichcalculates a theoretical SAXS profile from each model and thencompares it to the experimental SAXS profile (Fig. 6A). As we usedmodels and not crystal structures, the best fit was observed in thelower values of q (0–1.5 nm�1), which represent the global shape ofHeV PMD, whatever the model used (see Fig. 6A). At the lowest andhighest protein concentrations, the calculated scattering profile of thetrimeric model was found to fit better the experimental profile ascompared to the other models, especially in the low q values, therebyyielding the lowest χ value (see insets in Fig. 6A). At the intermediateprotein concentration, a slightly lower χ value was obtained for thedimeric form (see insets in Fig. 6A). Global statistical analysis of theobtained χ values using Bayesian-Factors (see Materials and methods)unambiguously showed that the trimeric model is much more likelythan both the dimeric and tetrameric models (Supplementary TableS2). Altogether these results indicate that HeV PMD adopts a trimericorganization in solution, as already observed for its close NiV relative.

Interestingly, the N-terminal “head” region is oriented differently inour previous NiV PMD model (hereafter referred to as “up” conforma-tion), being solvent exposed and not packed back onto the coiled-coilas in the models herein described (Blocquel et al., 2013). We thusreasoned that a different orientation of the head could be responsiblefor the observed differences in the length of the NiV and HeV PMDenvelopes. We therefore generated models of a dimeric, trimeric andtetrameric HeV PMD form in which the head is in the “up” configura-tion (see Materials and methods). We then assessed how these modelsfit the experimental SAXS profile using CRYSOL (Fig. 6B). At the lowestand highest protein concentrations, the lowest χ values were obtainedfor the trimer, while at the intermediate concentration a lower χ valuewas obtained for the dimer (Fig. 6B). Global statistical analysisperformed on all χ values allowed us to conclude that the trimer isthe model that is in best agreement with experimental data (seeSupplementary Table S2). Interestingly, comparison of the Chi valuesobtained with the “head-up” and “head-down” trimer clearly desig-nates the latter as the best fitting model (Fig. 6 and Supple-mentary Table S2).

Altogether, these data support the ability of Henipavirus PMDsto adopt different conformations differing in the orientation of thehead and hence in their length. Whether the “up” and “down”conformations are unique to NiV and HeV, respectively, orwhether both domains are able to adopt an equilibrium betweenthe two forms remains to be established.

Sedimentation analysis

In view of gaining additional experimental support for a trimericorganization of HeV PMD in solution, we performed sedimentation

equilibrium experiments (Fig. 7). Sedimentation equilibrium wasperformed with various loading protein concentrations (0.6, 1.6 and2.1 g/L) at 10,000, 12,000, 17,000 and 40,000 rpm at 4 1C. Analysis ofthe three concentrations at 280 and 257 nm showed a best residualsdistribution for a single species model. Multispeed global analysis foreach concentration at the two wavelengths lead to a mean molecularmass of 43,5007600 Da, a value close to that expected for a trimer.Equilibrium sedimentation experiments thus confirmed that HeVPMD adopts a trimeric conformation in solution.

SAXS studies of PCT

We next studied PCT by SAXS. The scattering profiles (Fig. 8A)and the obtained Guinier plots (Fig. 8B) were found to beconcentration-independent in the 0.44–2.48 g/L range providedthat 5 mM DTT was added. Above this concentration thresholdhowever, the experimental profiles indicated at least partialaggregation of the protein (data not shown). We therefore usedonly the data obtained at concentrations up to this thresholdvalue. Under these conditions, all the scattering curves are linearin the Guinier region. Guinier analysis in the low q region gave anRg of 87.070.1 Å, a value close to that obtained from the pairdistance distribution function (86.770.5 Å) (Tables 1 and 2). Themolecular mass deduced from I(0) is 114 kDa (Table 1). Althoughthis value slightly exceeds the molecular mass expected for atrimeric form (104.7 kDa), it remains nevertheless indicative of atrimeric organization of the protein. In this case, like in the case ofPMD, the experimentally observed Rg value is much larger (3.6times) than the value expected (24.0 Å) for a sphere with an Rs of31.0 Å, as determined from the volume of a sphere with 310�3residues (see Materials and methods). In this case, the discrepancybetween the observed and the expected Rg is even more pro-nounced than in the case of PMD. The finding that the actual Rg

r (cm)6.90 6.95 7.00 7.05 7.10

Abs

orba

nce

at 2

80 n

m

0.0

0.1

0.2

0.3

0.4

0.5

r (cm)6.90 6.95 7.00 7.05 7.10

Res

idua

ls-0.02

-0.01

0.00

0.01

0.02

Fig. 7. Sedimentation equilibrium analysis of HeV PMD at 0.6 g/L and at 4 1C. Thesymbols show the experimental radial distribution of HeV PMD at 10,000 rpm (▼),12,000 rpm (●) and 17,000 rpm (○). The solid lines represent the best fit curves ofthe multispeed global analysis with a single species model. For this concentration, amolecular mass of 42,2007800 Da was obtained.

M. Beltrandi et al. / Virology 477 (2015) 42–55 49

1

2

0

-1

In I(

q)

1 2 3-1q (nm )

0.44 g/L1.25 g/L 2.48 g/L

0.07

0.04

0.03

P(r

)

0.06

0.05

0.02

0.01

00 10

r(nm)3020

I(q)*

q2

1

2

3

4

q (nm-1)0 0.5 1 1.5 2 2.5

0

4.5

4

3.5

)q(InI

0.02 0.04 0.06 0.08

q2 (nm-2)

302 Å

180°129 Å

Fig. 8. Small-angle X-ray scattering experiments of HeV PCT. (A) Experimental SAXS data recorded for q values up to 3.5 nm�1. The curves obtained for three proteinconcentrations (0.44 g/L, light blue; 1.25 g/L, magenta; and 2.48 g/L, yellow) are represented after correction for concentration. (B) Representation of the Guinier plot for theprotein at 2.48 g/L. (C) Pair distance distribution, P(r), function of the data for the 2.48 g/L concentration. (D) Kratky plot. (E) Three views of the ab initio envelope calculatedwith DAMAVER (Volkov and Svergun, 2003). 20 DAMMIF calculations (Franke and Svergun, 2009) were performed and averaged with DAMAVER to produce the average andfiltered shape shown in blue.

M. Beltrandi et al. / Virology 477 (2015) 42–5550

largely exceeds that expected for a sphere reflects the highlyelongated nature of the protein, a property arising from theelongated shape of PMD combined with the presence of disor-dered regions. The very high Rg value also explains the spectacu-larly high apparent molecular mass observed in SEC studies.

The distance distribution function of PCT displays a maxi-mum at 90 Å and a Dmax value of 302 Å typical of an elongatedobject (Fig. 8C). The Rg of a thin rod of length L¼302 Å is 87.2 Å(see Materials and methods), a value in very good agreement withthe values estimated from the Guinier plot (87.0 Å) and from thepair distance distribution (86.7 Å). The shape of the Kratky plot,with a maximum at qE0.27 nm�1 and a flat region forq40.83 nm�1, is indicative of a mixed protein consisting of astructured and a disordered moiety (Fig. 8D), in agreement withthe far-UV CD data.

For ab initio shape reconstruction, we used the scattering dataobtained at the highest concentration so as to achieve maximalresolution, and proceeded as described for PMD. For each sym-metry class, models were found to be very reproducible with anaverage NSD below 1.0, indicating structurally similar solutions(Supplementary Fig. S6 and data not shown). All generated beadmodels have an elongated shape (data not shown). Like in the caseof PMD, best results were obtained with the P3 symmetry (data notshown), thus providing additional support for a trimeric organizationof the protein. Assuming a P3 symmetry, the models resulting from20 independent DAMMIF runs were superimposed using the DAMA-VER suite, which yielded an average NSD of 0.6670.23 after havingdiscarded four outliers. The average filtered model generated byDAMAVER has an elongated shape (Fig. 8E).

Altogether, these results confirm and reinforce conclusionsbased on studies carried out on HeV PMD. They therefore affordadditional reliability to the conclusion that the HeV P proteinadopts a trimeric organization in solution, like its NiV counterpart.

Conclusions

The results herein presented, along with our previouslyreported results on NiV PMD (Blocquel et al., 2013), provideexperimental evidence for a trimeric organization of Henipavirusphosphoproteins in solution. On the other hand, the crystallo-graphic structure of NiV PMD reported by Bruhn et al. unambigu-ously shows a tetrameric organization within the crystal (Bruhn-Johannsen et al., 2014).

It is conceivable that Henipavirus P proteins can form bothtrimers and tetramers depending on the conditions. That coiled-coils are able to modulate their oligomeric state according to thephysico-chemical conditions (pH, temperature.) or depending onwhether they are located inside or outside the cell, has alreadybeen reported (Dutta et al., 2001; Lupas and Gruber, 2005). Ofeven more interest, the GCN4 leucine-zipper domain was shownto adopt different oligomeric states depending on the crystal-lization conditions, implying that the amino acid sequence doesnot specify a unique oligomeric state (Oshaben et al., 2012). It isalso worthy to emphasize that conflicting experimental evidenceare not unique to NiV PMD: indeed SeV PMD had also been shownto form trimers in solution (Curran, 1998; Curran et al., 1995) andto adopt a tetrameric coiled-coil conformation in the crystal(Tarbouriech et al., 2000b). The experimental evidence pointingto a trimeric form of SeV P have been set aside perhaps too rapidlyin light of the crystallographic data pointing to a tetramericorganization. However, the finding that both SeV and NiV PMDcan form trimers in solution and tetramers in the crystal mayreflect their intrinsic ability to adopt different oligomeric statesthat could be related to different functional forms of the P proteinand to the different complexes (i.e. N–P, N1–P, P–L) that this latter

can form within infected cells. Likewise, the herein proposedability of the N-terminal “head” region of Henipavirus PMD toadopt different conformations, might play a functional role byimparting specificity in molecular interactivity.

In the same vein, structural comparison among the differentcrystallographic structures of MeV PMD solved so far unveiledunexpected structural variations (Blocquel et al., 2014; Communieet al., 2013a). Although all the structures have a tetrameric coiled-coil organization, structural comparison unveiled considerabledifferences not only in the quaternary structure but also in theextent of disorder within the C-terminal region of the coiled-coil(Blocquel et al., 2014). We have previously proposed that theunexpected plasticity and flexibility of MeV PMD could be the firsthint of the existence of different functional forms of the P proteinreflecting its multifunctional nature and pivotal role in thereplicative cycle.

In conclusion, the ability of SeV and Henipavirus P proteins toadopt different oligomeric states, together with the ability of MeVPMD to dynamically sample different forms differing in the degreeof compaction and in the extent of disorder, might be the basis forthe ability of the P protein to form different complexes critical fortranscription and replication, with conformational changes possi-bly dictating the ability to form a transcriptase versus a replicasecomplex.

Funding

This work was carried out with the financial support of theAgence Nationale de la Recherche, specific programs “Physico-Chimie du Vivant” (ANR-08-PCVI-0020-01) and “ASTRID”, ANR-11-ASTR-003-01. D.B. was supported by a joint doctoral fellowshipfrom the Direction Générale de l'Armement (DGA) and the CNRS.M.B. was previously supported by an Erasmus Master2 fellowshipfrom the University of Milan and is presently supported by a PhDfellowship from the Italo-French University. J.E. is supported by apost-doctoral fellowship by the Fondation pour la RechercheMédicale (FRM). The funders had no role in study design, datacollection and analysis, decision to publish, or preparation of themanuscript.

Materials and methods

Cloning of the PMD and PCT coding regions

The HeV PMD and PCT constructs, encoding respectively residues469–578 and 404–707 of the HeV P protein (SwissProt sequenceaccession number O55778.2), all with an hexahistidine tag fused totheir C-terminus, were obtained by polymerase chain reaction (PCR).PCR was performed using a synthetic HeV P gene (GenScript),optimized for the expression in Escherichia coli, as template, and Pfupolymerase (Promega). Primers (Operon) were designed to introducean AttB1 and AttB2 site at the 50 and 30 ends, respectively, and toamplify the desired part of the P ORF with a fragment encoding a C-terminal hexahistidine tag. The rationale for the choice of the tagposition was to avoid purification of truncated forms arising frompossible abortive translation. After digestion with DpnI (New EnglandBiolabs) to remove the methylated DNA template, the PCR productswere purified (PCR Purification Kit, Qiagen) and cloned into thepDEST14 vector (Invitrogen) using the Gateways recombinationsystem (Invitrogen). This vector drives the expression of recombinantproducts under the control of the T7 promoter. Selection andamplification of the construct was carried out using CaCl2-competent

M. Beltrandi et al. / Virology 477 (2015) 42–55 51

E. coli TAM1 cells (Active Motif). The sequences of the coding regionsof the constructs were checked by sequencing (GATC Biotech) andfound to conform to expectations.

Expression and purification of PMD and PCT constructs

The E. coli strain Rosetta [DE3] pLysS (Novagen) was used forexpression of all the P constructs. The pLysS (Novagen) plasmidcarries the lysozyme gene, thus allowing tight regulation of theexpression of the recombinant gene, as well as a facilitated lysis.Cultures were grown overnight to saturation in LB mediumcontaining 100 mg/mL ampicilin and 34 mg/mL chloramphenicol.An aliquot of the overnight culture was diluted 1/25 in LB mediumand grown at 37 1C. At an OD600 of 0.7, isopropyl β-D-thiogalacto-pyranoside (IPTG) was added to a final concentration of either0.5 mM (PMD) or 0.2 mM (PCT), and the cells were subsequentlygrown at 37 1C for 4 h. The induced cells were harvested, washedand collected by centrifugation (5000g, 10 min). The resultingbacterial pellets were frozen at �20 1C.

Bacterial pellets, irrespective of whether they contained PMD orPCT, were resuspended in 5 volumes (v/w) of buffer A (50 mM Tris/HCl pH 8, 1 M NaCl, 10 mM Imidazole) supplemented with 0.1 mg/mLlysozyme, 10 μg/mL DNAse I, 20 mM MgSO4 and protease inhibitorcocktail (Sigma) (one tablet for a 2 L-culture). After a 20-min incuba-tion with gentle agitation at 4 1C, the cells were disrupted bysonication (using a 750W sonicator and 5 cycles of 30 s each at 60%power output). The lysates were clarified by centrifugation at 30,000gfor 30 min and then purified by immobilized metal affinity chromato-graphy (IMAC). The clarified supernatant from a 1 L-culture wasinjected onto a 5-mL HisTrap FF column (GE, Healthcare), previouslyequilibrated in buffer A supplemented with 1 M NaCl. Elution wascarried out using a gradient of imidazole (20–500mM) in buffer Asupplemented with 1 M NaCl. Eluents were analyzed by SDS-PAGE forthe presence of the desired product. The fractions containing thedesired recombinant product were combined and loaded onto aSuperdex 200 HR 16/60 column (GE, Healthcare). PMD was elutedin 50 mM Hepes pH 8, and 300 mM NaCl, while PCT was eluted in30 mM Tris/HCl pH 8, 5 mM EDTA, 300 mM NaCl, 1 mM PMSF(phenylmethylsulfonylfluoride), and glycerol 10% supplemented with200mM arginine. Proteins were concentrated using Centricon Plus-20(molecular cutoff of 3000 Da for PMD and of 30,000 Da for PCT)(Millipore) and then stored at �20 1C.

All purification steps, except for IMAC and gel filtrations, werecarried out at 4 1C. Apparent molecular mass of proteins elutedfrom the gel filtration column was deduced from a calibrationcarried out with ferritin (440 kDa), catalase (232 kDa), BSA (132and 66 kDa), Tobacco Etch virus protease (27 kDa), and lysozyme(14 kDa).

Protein concentrations were either calculated using the theoreticalabsorption coefficients at 280 nm, as provided by the programProtParam at the EXPASY server (http://www.expasy.ch/tools), orestimated via quantitative analysis of the amino acid composition. Inthose analyses, PMD samples were dried and hydrolyzed at 110 1C inconstant-boiling HCl containing 1% (v/v) phenol for 24 h underreduced pressure and in the absence of oxygen. Amino acids wereanalyzed on a model Biochrom 30 amino acid analyzer, with thestandard sodium citrate eluting buffer system. Calibrations were madewith standard solutions that contained all the amino acids excepttryptophan. The amino acids' concentrations were estimated throughreaction with ninhydrine. The internal standard was norleucine. Oncethe actual protein concentration has been determined in that manner,the corresponding experimental absorption coefficient at 280 nmcould be derived and found to be very close (4.209 mM�1 cm�1) tothe theoretical one (4.470 mM�1 cm�1).

Mass spectrometry (MALDI–TOF)

The identity of the purified PMD and PCT proteins was confirmedbymass spectral analysis of tryptic fragments. Samples for this analysiswere obtained by digesting (0.25 μg trypsin) 1 μg of purified recombi-nant protein obtained after separation onto SDS-PAGE. Mass analysisof the tryptic peptides was performed using an Autoflex II TOF/TOF(Bruker Daltonics, Bremen, Germany). Peptides were analyzed in theAutoflex matrix-assisted laser desorption ionization/time of flight(MALDI–TOF). Spectra were acquired in the linear mode. Samples(0.7 mL containing 15 pmol) were mixed with an equal volume ofsinapinic acid matrix solution, spotted on the target, then dried atroom temperature for 10 min. The mass standards were eitherautolytic tryptic peptides or peptide standards (Bruker Daltonics).Peptide fingerprints were obtained and compared in silicowith proteindigest (Biotools, Bruker Daltonics, Germany).

Analytical ultracentrifugation (AUC)

Equilibrium sedimentation experiments were performed at4 1C in a Beckman Optima-XL-A analytical ultracentrifuge in sixchannel centerpiece. Measurements were done at three successivespeeds (10,000, 12,000 and 17,000 rpm) by taking scans at 280 nmand 257 nm, when sedimentation equilibrium was reached. High-sedimentation was conducted afterwards for baseline correction.At 4 1C, the partial specific volume of HeV PMD, solvent densityand viscosity calculated with SEDNTERP (Laue et al., 1992) were0.72793 mL/g, 0.99823 g/cm3 and 0.01567 poise, respectively.

For one concentration, a multispeed global analysis was appliedto the data using the SEDPHAT program (Vistica et al., 2004) andthe best fit was obtained for a single solute.

Calculation of the hydrodynamic radius and of radius of gyration

The theoretical Stokes radii (Rs, in Å) expected for a nativelyfolded ðRNF

s Þ protein with an expected molecular mass (MMtheo)(in Daltons) were calculated according to (Uversky, 2002) thefollowing equation:

log ðRNFs Þ ¼ 0:369logðMMtheoÞ�0:254 ð1Þ

The theoretical Stokes radii (Rs) of a natively folded trimerðRTrimer

s Þ or tetramer ðRTetramers Þ was calculated as follows:

log RTrimers

� �¼ 0:369logðMMtheo� 3Þ�0:254 ð2Þ

log RTetramers

� �¼ 0:369logðMMtheo� 4Þ�0:254 ð3Þ

The theoretical radius of gyration (Rg, in Å) expected for aglobular protein with a hydrodynamic radius Rs was calculatedaccording to (Wilkins et al., 1999) the following equation:

Rg ¼ 3=5� �1=2Rs ð4Þ

The Rg of a thin rod with a length L can be calculated as follows:

R2g ¼ L2=12 ð5Þ

PMD and PCT consist of 117 and 310 residues, respectively,including the initial methionine and the hexahistidine tag. Usingan average volume of 134 Å3 per residue for proteins, and con-sidering a trimeric organization for both PMD and PCT, the radiusof a sphere with volume V¼4/3πRS3 would be 22.4 Å in the case ofPMD, and 31.0 Å in the case of PCT. According to Eq. (4), thecorresponding Rg would be 17.3 Å for PMD and 24.0 Å for PCT.

M. Beltrandi et al. / Virology 477 (2015) 42–5552

Cross-linking experiments

A fixed amount of PMD (5 μg) was incubated for 20 h at roomtemperature with increasing amounts (0–10mM) of suberic acid bis(N-hydroxy-succinimide ester) (SAB) in 20 mM Hepes pH 7.0, 150 mMNaCl in a final volume of 30 μL. The reactions were stopped by addingLaemli sample buffer and by heating the sample at 95 1C for 5 min.The samples were analyzed by 15% SDS-PAGE followed by CoomassieBlue staining. SAB was first solubilized in DMSO at a concentration of1% and then diluted in 50mM Hepes pH 7.0, 150 mM NaCl to thedesired concentration.

Far-UV circular dichroism (CD)

For both PMD and PCT, CD spectra were recorded on a Jasco 810dichrograph using 1-mm thick quartz cells in 10 mM sodium phos-phate pH 7 at 20 1C either in the absence or in the presence ofincreasing concentrations of 2,2,2 trifluoroethanol (TFE). Both CDspectra were measured between 185 and 260 nm, at 0.2 nm/minand were averaged from three independent acquisitions. Mean molarellipticity values per residue (MRE) were calculated as MRE¼3300mΔA/(lcn), where l (path length)¼0.1 cm, c (protein concentration in g/l)¼0.1, n (number of residues)¼117 for PMD and 310 for PCT, m(molecular mass in Daltons)¼13,524 for PMD and 34,923 for PCT.Spectra were deconvoluted using the DICHROWEB website (http://dichroweb.cryst.bbk.ac.uk/html/home.shtml) which was supported bygrants to the BBSRC Centre for Protein and Membrane Structure andDynamics (CPMSD) (Whitmore and Wallace, 2004, 2008). The CON-TINLL deconvolution method was used to estimate the α-helicalcontent using the reference protein set 7.

In order to monitor protein unfolding, measurements at fixedwavelength (222 nm) were performed in the temperature range of20–100 1C with data pitch 1 1C and a temperature slope of 1 1C/min.The buffer solution without the protein was used as blank. Fitting ofexperimental data was done using Sigmaplot.

Small angle X-ray scattering (SAXS) measurements and ab initio 3Dshape reconstructions

All small-angle X-ray scattering (SAXS) measurements werecarried out at the ESRF on beamline BM29 at a working energy of12.5 keV. The sample-to-detector distance of the X-rays was 2.847 m,leading to scattering vectors q ranging from 0.028 to 4.525 nm�1

(see Table 1 for details on data collection). The scattering vector isdefined as q¼4π/λ sin θ, where 2θ is the scattering angle. Theexposure time was optimized to reduce radiation damage.

SAXS data were collected at 10 1C using purified protein samples(30 μl each). Protein concentrations and buffers were as follows: 0.77,1.36 and 3.31 g/L in 20 mM Tris/HCl pH 8, 5 mM EDTA, 150 mM NaClfor PMD; 0.44, 1.25, 2.48 and 4.0 g/L in 20 mM Tris/HCl pH 8, 5 mMEDTA, 500mM NaCl, and 5mM DTT for PCT.

Samples were loaded in a fully automated sample charger. Tenexposures of 10 s each were made for each protein concentrationand data were combined to give the average scattering curve foreach measurement. Any data points affected by aggregation, possi-bly induced by radiation damages were excluded. The profilesobtained in the range 0.77–3.31 g/L for PMD, and 0.44–2.48 g/Lfor PCT had the same shape and were flat at low q values indicatingthe absence of significant aggregation. Then, we used the higherconcentration (3.31 g/L for PMD and 2.48 g/L for PCT) to obtainmaximal information at high resolution.

Data reductions were performed using the established proce-dure available at BM29, and buffer background runs were sub-tracted from sample runs. The Rg and forward intensity at zeroangle I(0) were determined with the program PRIMUS (Konarev etal., 2003) according to the Guinier approximation at low values, in

a QRg range up to 1.3 as follows:

ln ½IðQ Þ� ¼ ln ½I0��Q2R2

g

3ð6Þ

The forward scattering intensities were calibrated using bovineserum albumin as reference. The Rg and pair distance distributionfunction, P(r) were calculated with the program GNOM (Svergun,1992). The maximum dimension (Dmax) value was adjusted suchthat the Rg value obtained from GNOM agreed with that obtainedfrom the Guinier analysis.

In the case of PMD, 3D beads models were built by fitting thescattering data with the program DAMMIF (Franke and Svergun,2009). This program restores a low-resolution shape of the proteinas a volume filled with densely packed spheres that reproduces theexperimental scattering curve by a simulated annealing minimizationprocedure. DAMMIF minimizes the interfacial area between themolecule and the solvent by imposing compactness and connectivityconstraints. 20 Independent models were generated with DAMMIFeither without imposing a symmetry or by imposing a P2, P3 and P4symmetry. The models resulting from independent runs were super-imposed using the DAMAVER suite (Volkov and Svergun, 2003). Thisyielded an initial alignment of structures based on their axes of inertiafollowed byminimization of the normalized spatial discrepancy (NSD),which is zero for identical objects and larger than one for system-atically different objects. Models with an NSD value greater than themean value plus twice the standard deviation were discarded.Accordingly, two and four models were discarded in the case ofPMD and PCT, respectively, when imposing the P3 symmetry. Thealigned structures were then averaged, giving an effective occupancyto each voxel in the model, and filtered at half-maximal occupancy toproduce models of the appropriate volume that were used for allsubsequent analyses.

Structural modeling

Structural models of HeV PMD were generated as follows. Themodel of the tetramer was generated using as template thestructure of NiV PMD (pdb code 4N5B) whose sequence is highlysimilar (98%) to that of HeV PMD. The additional N-terminal fiveresidues, including the initial methionine and the first fourresidues of the HeV PMD construct (aa 469–472) whose corre-sponding NiV residues are not defined in the electron density ofthe NiV PMD structure, were modeled using the SAM-T08 server(Karplus et al., 2003, 2005; Katzman et al., 2008). For each chain ofthe tetramer, the two last C-terminal HeV PMD residues (P577 andG578), whose cognate NiV PMD residues are not defined in theelectron density of the NiV PMD structure, and the C-terminalhexahistidine tag (6xHis tag) were modeled by prolonging thecoiled-coil. The final model thus consists of a long coiled-coilregion (residues 506–578 plus the 6xHis tag) and an N-terminalregion (residues 469–505, hereafter referred to as “head”) that isnot involved in formation of the coiled-coil.

To generate the model of the trimer, we first generated a modelof a trimeric coiled-coil of 79 residues in length using the structureof hemagglutinin (pdb code 2HMG) as starting model. Then, usingChimera (Pettersen et al., 2004), we appended at the N-terminusof each of the three chains the “head” fragment in an orientationas much similar as possible to that found in the tetramer whileavoiding obvious steric clashes.

The dimeric model of HeV PMD, with a 79-residue coiled-coilregion, was generated using the structure of Saccharomyces cerevisiaeAtg16e (pdb code 3A7O) as starting model. The “head” was appendedat the N-terminus of each of the two chains as described above.

Models with the head oriented as in our previous NiV PMDmodel (Blocquel et al., 2013), and herein referred to as “head-up”models, were generated from the aforementioned models by

M. Beltrandi et al. / Virology 477 (2015) 42–55 53

appending the “head” in opposite orientation, i.e. exposed to thesolvent rather than packed back onto the coiled-coil region. Notethat in these latter models the “head” conserves the same fold asin the “head-down” models, so as to minimize differences amongprotomers.

All the models were subjected to 10,000 steps of steepest descentenergy minimization using the CHARMM19-EEF1 force field (Lazaridisand Karplus, 1999) implemented in the molecular modeling programALMOST (Fu et al., 2014). They were then further refined to removesteric clashes and improve the agreement with the experimental SAXSprofiles. The refinement was performed using ALMOST (Fu et al., 2014)and consisted of 25 rounds. In each round, 100 structures wheregenerated by performing a simulated annealing cycle of 40,000 (80 ps)steps of molecular dynamics starting from a randomly selectedstructure among the 25 with the lowest energy from the previousround. As energy function we used the CHARMM19-EEF1 force fieldwith an additional term that models the agreement of the structurewith the experimental SAXS profile (Forster et al., 2008). Finally thestructure with lowest energy among all 25,000 models was selected.

Statistical analysis

To estimate the statistical significance of the individual modelswe use Bayesian-Factors (Kass and Raftery, 1995). Bayesian-Factorsprovide a direct way to estimate the posterior odds of two models(hypothesis) A and B by

K ¼ PrðAjdataÞPrðBjdataÞ ð7Þ

where Pr(X|data) is the probability of a model X given a set of data.Appling Bayes theorem, we obtain

K ¼ PðdatajAÞPðAÞP datajBð ÞPðBÞ: ð8Þ

The term P(A)/P(B) is the prior odds of the two models. Thisfactor can be taken to be equal to 1.0 because it is natural toassume that in the absence of any experimental data all modelsare equally likely. From this, it follows that in the case of a singledata point the Bayesian-Factor of two models A and B is

K ¼ e�ð1=2Þχ2A=e�ð1=2Þχ2B ð9Þ

where

χ2 ¼ ðDobs�DcalcÞ2σ2

ð10Þ

is the chi-square between the observed and the calculated dataand σ is the experimental error. In the presence of N independentdata point the previous equation can be easily generalized to

χ2 ¼XNi ¼ 1

ðDobsi �Dcalc

i Þ2σ2i

ð11Þ

The chi-square between an observed and a calculated SAXSprofile is given by

χ2SAXS ¼1M

XMi ¼ 1

ðI qi� �obs� IðqiÞcalcÞ2

σ2i; ð12Þ

where M is the number of sampled points. It would, however, bean error to compute the Bayesian-Factors using χ2 ¼Mχ2SAXSbecause not all M observations are independent. In fact, the SAXSprofile I(q) is the Fourier transform of a band-limited function, thepair-distance distribution P(r), which is different from zero onlyfor 0orodmax. In this case, the sampling theorem guaranteesthe ability to reconstruct the entire SAXS profile from a set of

Shannon-points qn¼nπ/dmax (Rambo and Tainer, 2013) as follows:

I qð Þ ¼X1n ¼ 0

Inπdmax

� �sin ðqdmax�nπÞqdmax�nπ

ð13Þ

Because the scattering intensity at any value of q is theweighted average of the intensities at Shannon-points, it followsthat these points form a set of maximally independent observa-tions. From this, it follows that the number of independentobservations in a scattering profile truncated at qmax is as follows:

n¼ qmaxdmax=π ð14ÞFor HeV PMD the maximal distance obtained from the calcula-

tion of the pair-distance distribution is dmax¼130 Å. With aqmax¼0.5 Å�1 we obtain approximately n¼20 independent obser-vations. Putting everything together, we get that the final equationfor relative probability of two models is

K ¼ e�10χ2Tot A=e�10χ2Tot B ð15Þwhere

χ2Tot A ¼ χ20:77þχ21:36þχ23:31 ð16Þis the sum of the SAXS chi-square at the three measured PMDconcentrations.

Acknowledgments

We wish to thank Petra Pernot and Martha Brennich (ESRF) fortheir help in SAXS data collection, and the ESRF synchrotron forbeamtime allocation. We are also grateful to Marion Sevajol andFrançois Ferron (AFMB lab) for their help in SAXS data collectionand analysis, and Gerlind Sulzenbacher (AFMB lab) for efficientlymanaging the AFMB BAG. We also thank Jean-Pierre Andrieu(Seq3A Platform: sequencing and amino acid compositions ana-lyses, Institut de Biologie Structurale, Grenoble, France) for aminoacid composition analyses of HeV PMD, and Christophe Flaudropsfrom the mass spectrometry platform of the IFR48 of Marseille formass spectrometry analyses. S.L. would like to thank RobertaPierattelli (CERM, Universita’ degli Studi of Florence) for co-supervising with her M.B. Finally, we would like to express ourgratitude to Johnny Habchi (previously within the AFMB lab andpresently at the Dept. of Chemistry, Cambridge, UK) for usefulcomments and suggestions on the manuscript.

Appendix A. Supplementary information

Supplementary data associated with this article can be found inthe online version at http://dx.doi.org/10.1016/j.virol.2014.12.029.

References

Albertini, A.A.V., Schoehn, G., Ruigrok, R.W., 2005. Structures impliquées dans laréplication et la transcription des virus à ARN non segmentés de sens négatif.Virologie 9, 83–92.

Blocquel, D., Beltrandi, M., Erales, J., Barbier, P., Longhi, S., 2013. Biochemical andstructural studies of the oligomerization domain of the Nipah virus phospho-protein: evidence for an elongated coiled-coil homotrimer. Virology 446,162–172.

Blocquel, D., Bourhis, J.M., Eléouët, J.F., Gerlier, D., Habchi, J., Jamin, M., Longhi, S.,Yabukarski, F., 2012a. Transcription et réplication des Mononégavirales: unemachine moléculaire originale. Virologie 16, 225–257.

Blocquel, D., Habchi, J., Durand, E., Sevajol, M., Ferron, F., Erales, J., Papageorgiou, N.,Longhi, S., 2014. Coiled-coil deformations in crystal structures: the measlesvirus phosphoprotein multimerization domain as an illustrative example. ActaCryst. D 70, 1589–1603.

Blocquel, D., Habchi, J., Gruet, A., Blangy, S., Longhi, S., 2012b. Compaction andbinding properties of the intrinsically disordered C-terminal domain of Heni-pavirus nucleoprotein as unveiled by deletion studies. Mol. Biosyst. 8, 392–410.

M. Beltrandi et al. / Virology 477 (2015) 42–5554

Bruhn-Johannsen, J.F., Barnett, K., Bibby, J., Thomas, J., Keegan, R., Rigden, D.,Bornholdt, Z.A., Saphire, E.O., 2014. Crystal structure of the Nipah virusphosphoprotein tetramerization domain. J. Virol. 88, 758–762.

Chan, Y.P., Koh, C.L., Lam, S.K., Wang, L.F., 2004. Mapping of domains responsible fornucleocapsid protein-phosphoprotein interaction of Henipaviruses. J. Gen.Virol. 85, 1675–1684.

Chouard, T., 2011. Structural biology: breaking the protein rules. Nature 471,151–153.

Communie, G., Crepin, T., Maurin, D., Jensen, M.R., Blackledge, M., Ruigrok, R.W.,2013a. Structure of the tetramerization domain of measles virus phosphopro-tein. J. Virol. 87, 7166–7169.

Communie, G., Habchi, J., Yabukarski, F., Blocquel, D., Schneider, R., Tarbouriech, N.,Papageorgiou, N., Ruigrok, R.W., Jamin, M., Ringkjøbing-Jensen, M., Longhi, S.,Blackledge, M., 2013b. Atomic resolution description of the interaction betweenthe nucleoprotein and phosphoprotein of Hendra virus. PLoS Pathog. 9,e1003631.

Cox, R., Green, T.J., Purushotham, S., Deivanayagam, C., Bedwell, G.J., Prevelige, P.E.,Luo, M., 2013. Structural and functional characterization of the mumps virusphosphoprotein. J. Virol. 87, 7558–7568.

Curran, J., 1998. A role for the Sendai virus P protein trimer in RNA synthesis.J. Virol. 72, 4274–4280.

Curran, J., Boeck, R., Lin-Marq, N., Lupas, A., Kolakofsky, D., 1995. Paramyxovirusphosphoproteins form homotrimers as determined by an epitope dilutionassay, via predicted coiled coils. Virology 214, 139–149.

DeLano, W.L., 2002. The PyMOL molecular graphics system. Proteins: Struct., Funct.Bioinform. 30, 442–454.

Dunker, A.K., Oldfield, C.J., Meng, J., Romero, P., Yang, J.Y., Chen, J.W., Vacic, V.,Obradovic, Z., Uversky, V.N., 2008a. The unfoldomics decade: an update onintrinsically disordered proteins. BMC Genomics 9 (Suppl. 2), S1.

Dunker, A.K., Silman, I., Uversky, V.N., Sussman, J.L., 2008b. Function and structureof inherently disordered proteins. Curr. Opin. Struct. Biol. 18, 756–764.

Dutta, K., Alexandrov, A., Huang, H., Pascal, S.M., 2001. pH-induced folding of anapoptotic coiled coil. Protein Sci. 10, 2531–2540.

Eaton, B.T., Mackenzie, J.S., Wang, L.F., 2007. Henipaviruses. In: Fields, B.N., Knipe, D.M., Howley, P.M. (Eds.), Fields Virology, 5th ed. Lippincott-Raven, Philadelphia,pp. 1587–1600.

Ferron, F., Rancurel, C., Longhi, S., Cambillau, C., Henrissat, B., Canard, B., 2005.VaZyMolO: a tool to define and classify modularity in viral proteins. J. Gen.Virol. 86, 743–749.

Forster, F., Webb, B., Krukenberg, K.A., Tsuruta, H., Agard, D.A., Sali, A., 2008.Integration of small-angle X-ray scattering data into structural modeling ofproteins and their assemblies. J. Mol. Biol. 382, 1089–1106.

Franke, D., Svergun, D.I., 2009. DAMMIF, a program for rapid ab-initio shapedetermination in small-angle scattering. J. Appl. Cryst. 42, 342–346.

Fu, B., Sahakyan, A.B., Camilloni, C., Tartaglia, G.G., Paci, E., Caflisch, A., Vendruscolo,M., Cavalli, A., 2014. ALMOST: an all atom molecular simulation toolkit forprotein structure determination. J. Comput. Chem. 35, 1101–1105.

Habchi, J., Blangy, S., Mamelli, L., Ringkjobing Jensen, M., Blackledge, M., Darbon, H.,Oglesbee, M., Shu, Y., Longhi, S., 2011. Characterization of the interactionsbetween the nucleoprotein and the phosphoprotein of Henipaviruses. J. Biol.Chem. 286, 13583–13602.

Habchi, J., Mamelli, L., Darbon, H., Longhi, S., 2010. Structural disorder withinHenipavirus nucleoprotein and phosphoprotein: from predictions to experi-mental assessment. PLoS ONE 5, e11684.

Habchi, J., Tompa, P., Longhi, S., Uversky, V.N., 2014. Introducing protein intrinsicdisorder. Chem. Rev. 114, 6561–6588.

Karlin, D., Ferron, F., Canard, B., Longhi, S., 2003. Structural disorder and modularorganization in Paramyxovirinae N and P. J. Gen. Virol. 84, 3239–3252.

Karplus, K., Karchin, R., Draper, J., Casper, J., Mandel-Gutfreund, Y., Diekhans, M.,Hughey, R., 2003. Combining local-structure, fold-recognition, and new foldmethods for protein structure prediction. Proteins 53 (Suppl. 6), S491–S496.

Karplus, K., Katzman, S., Shackleford, G., Koeva, M., Draper, J., Barnes, B., Soriano, M.,Hughey, R., 2005. SAM-T04: what is new in protein-structure prediction forCASP6. Proteins 61 (Suppl. 7), S135–S142.

Kass, R., Raftery, A., 1995. Bayes factors. J. Am. Stat. Assoc. 90, 773–795.Katzman, S., Barrett, C., Thiltgen, G., Karchin, R., Karplus, K., 2008. PREDICT-2ND: a

tool for generalized protein local structure prediction. Bioinformatics 24,2453–2459.

Konarev, P.V., Volkov, V.V., Sokolova, A.V., Koch, M.H.J., Svergun, D.I., 2003. PRIMUS:a Windows PC-based system for small-angle scattering data analysis. J. Appl.Cryst. 36, 1277–1282.

Lamb, R.A., Parks, G.D., 2007. Paramyxoviridae: the viruses and their replication. In:Knipe, D.M., Howley, P.M. (Eds.), Fields Virology, 5th edition LippincottWilliams & Wilkins, Philadelphia, PA, pp. 1450–1497.

Lau, S.Y., Taneja, A.K., Hodges, R.S., 1984. Synthesis of a model protein of definedsecondary and quaternary structure. Effect of chain length on the stabilizationand formation of two-stranded alpha-helical coiled-coils. J. Biol. Chem. 259,13253–13261.

Laue, T.M., Shah, B.D., Ridgeway, T.M., Pelletier, S.L., 1992. Computer-aided inter-pretation of analytical sedimentation data for proteins in analytical ultracen-trifugation. In: Harding, S.E., Rowe, A.J., Horton, J.C. (Eds.), Biochemistry andPolymer Sciences Royal Society of Chemistry, Cambridge, UK, pp. 90–125.

Lazaridis, T., Karplus, M., 1999. Effective energy function for proteins in solution.Proteins 35, 133–152.

Llorente, M.T., Barreno-Garcia, B., Calero, M., Camafeita, E., Lopez, J.A., Longhi, S.,Ferron, F., Varela, P.F., Melero, J.A., 2006. Structural analysis of the humanrespiratory syncitial virus phosphoprotein: characterization of an a-helicaldomain involved in oligomerization. J. Gen. Virol. 87, 159–169.

Llorente, M.T., Taylor, I.A., Lopez-Vinas, E., Gomez-Puertas, P., Calder, L.J., Garcia-Barreno, B., Melero, J.A., 2008. Structural properties of the human respiratorysyncytial virus P protein: evidence for an elongated homotetrameric moleculethat is the smallest orthologue within the family of paramyxovirus polymerasecofactors. Proteins 72, 946–958.

Lupas, A.N., Gruber, M., 2005. The structure of alpha-helical coiled coils. Adv.Protein Chem. 70, 37–78.

Marsh, G.A., de Jong, C., Barr, J.A., Tachedjian, M., Smith, C., Middleton, D., Yu, M.,Todd, S., Foord, A.J., Haring, V., Payne, J., Robinson, R., Broz, I., Crameri, G., Field,H.E., Wang, L.F., 2012. Cedar virus: a novel Henipavirus isolated from Australianbats. PLoS Pathog. 8, e1002836.

Mylonas, E., Svergun, D., 2007. Accuracy of molecular mass determination ofproteins in solution by small-angle X-ray scattering. J. Appl. Crystallogr. 40,s245–s249.

Omi-Furutani, M., Yoneda, M., Fujita, K., Ikeda, F., Kai, C., 2010. Novelphosphoprotein-interacting region in Nipah virus nucleocapsid protein andits involvement in viral replication. J. Virol. 84, 9793–9799.

Oshaben, K.M., Salari, R., McCaslin, D.R., Chong, L.T., Horne, W.S., 2012. The nativeGCN4 leucine-zipper domain does not uniquely specify a dimeric oligomeriza-tion state. Biochemistry 51, 9581–9591.

Pettersen, E.F., Goddard, T.D., Huang, C.C., Couch, G.S., Greenblatt, D.M., Meng, E.C.,Ferrin, T.E., 2004. UCSF Chimera—a visualization system for exploratoryresearch and analysis. J. Comput. Chem. 25, 1605–1612.

Rahaman, A., Srinivasan, N., Shamala, N., Shaila, M.S., 2004. Phosphoprotein of therinderpest virus forms a tetramer through a coiled coil region important forbiological function. A structural insight. J. Biol. Chem. 279, 23606–23614.

Rambo, R.P., Tainer, J.A., 2013. Accurate assessment of mass, models and resolutionby small-angle scattering. Nature 496, 477–481.

Roux, L., 2005. Dans le génome des Paramyxovirinae, les promoteurs et leursactivités sont façonnés par la⪡règle de six. Virologie 9, 19–34.

Salvamani, S., Goh, Z., Ho, K., Tey, B., Tan, W., 2013. Oligomerization state of themultimerization domain of Nipah virus phosphoprotein. Process Biochem. 48,1476–1480.

Svergun, D., 1992. Determination of the regularization parameters in indirect-trasform methods using perceptual criteria. J. Appl. Cryst. 25, 495–503.

Svergun, D., Barberato, C., Koch, M.H.J., 1995. CRYSOL, a program to evaluate X-raysolution scattering of biological macromolecules from atomic coordinates.J. Appl. Cryst. 28, 768–773.

Tarbouriech, N., Curran, J., Ebel, C., Ruigrok, R.W., Burmeister, W.P., 2000a. On thedomain structure and the polymerization state of the sendai virus P protein.Virology 266, 99–109.

Tarbouriech, N., Curran, J., Ruigrok, R.W., Burmeister, W.P., 2000b. Tetrameric coiledcoil domain of Sendai virus phosphoprotein. Nat. Struct. Biol. 7, 777–781.

Turoverov, K.K., Kuznetsova, I.M., Uversky, V.N., 2010. The protein kingdom extended:ordered and intrinsically disordered proteins, their folding, supramolecular complexformation, and aggregation. Prog. Biophys. Mol. Biol. 102, 73–84.

Uversky, V.N., 2002. What does it mean to be natively unfolded? Eur. J. Biochem.269, 2–12.

Uversky, V.N., 2010. The mysterious unfoldome: structureless, underappreciated,yet vital part of any given proteome. J. Biomed. Biotechnol. 2010, 568068.

Uversky, V.N., 2013. A decade and a half of protein intrinsic disorder: Biology stillwaits for physics. Protein Sci.

Uversky, V.N., Dunker, A.K., 2010. Understanding protein non-folding. Biochim.Biophys. Acta 1804, 1231–1264.

Vistica, J., Dam, J., Balbo, A., Yikilmaz, E., Mariuzza, R.A., Rouault, T.A., Schuck, P.,2004. Sedimentation equilibrium analysis of protein interactions with globalimplicit mass conservation constraints and systematic noise decomposition.Anal. Biochem. 326, 234–256.

Volkov, V.V., Svergun, D.I., 2003. Uniqueness of ab initio shape determination insmall-angle scattering. J. Appl. Cryst. 36, 860–864.

Wang, L.F., Yu, M., Hansson, E., Pritchard, L.I., Shiell, B., Michalski, W.P., Eaton, B.T.,2000. The exceptionally large genome of Hendra virus: support for creation of anew genus within the family Paramyxoviridae. J. Virol. 74, 9972–9979.

Whitmore, L., Wallace, B.A., 2004. DICHROWEB, an online server for proteinsecondary structure analyses from circular dichroism spectroscopic data.Nucleic Acids Res. 32, W668–W673.

Whitmore, L., Wallace, B.A., 2008. Protein secondary structure analyses fromcircular dichroism spectroscopy: methods and reference databases. Biopoly-mers 89, 392–400.

Wilkins, D.K., Grimshaw, S.B., Receveur, V., Dobson, C.M., Jones, J.A., Smith, L.J.,1999. Hydrodynamic radii of native and denatured proteins measured by pulsefield gradient NMR techniques. Biochemistry 38, 16424–16431.

Woody, R.W., 2010. Circular dichroism of intrinsically disordered proteins. In:Uversky, V.N., Longhi, S. (Eds.), Instrumental Analysis of Intrinsically DisorderedProteins: Assessing Structure and Conformation. John Wiley and Sons, Hoboken(NJ), pp. 303–321.

M. Beltrandi et al. / Virology 477 (2015) 42–55 55