JVI Accepts, published online ahead of print on 15 January 2014 J. Virol. doi:10.1128/JVI.03483-13 Copyright © 2014, American Society for Microbiology. All Rights Reserved.

1

Common mechanism for RNA encapsidation by negative strand RNA viruses

2

3

Todd J Green, Robert Cox, Jun Tsao, Michael Rowse, Shihong Qiu, Ming Luo#

4

5

Department of Microbiology

6

University of Alabama at Birmingham

7

Birmingham, Alabama, 35294, USA

8

9

10

(Running title: NSV RNA encapsidation) #

send correspondence to:

11

1720 2nd avenue south

12

Birmingham, Alabama, 35294, USA

13

Phone: 1-205-934-4259

14

Fax: 1-205-934-0480

15

Email: [email protected]

1

16

Abstract

17

The nucleocapsid of a negative strand RNA virus is assembled with a single nucleocapsid

18

protein and the viral genomic RNA. The nucleocapsid protein polymerizes along the length of

19

the single strand genomic RNA (vRNA) or its complementary RNA (cRNA). This process of

20

encapsidation occurs concomitantly with genomic replication. Structural comparisons of several

21

nucleocapsid-like particles show that the mechanism of RNA encapsidation in negative strand

22

RNA viruses has many common features. Fundamentally, there is a unifying mechanism to keep

23

the capsid protein protomer monomeric prior to encapsidation of viral RNA. In the nucleocapsid,

24

there is a cavity between two globular domains of the nucleocapsid protein where the viral RNA

25

is sequestered. The viral RNA must be transiently released from the nucleocapsid in order to

26

reveal the template RNA sequence for transcription/replication. There are cross-molecular

27

interactions among the protein subunits linearly along the nucleocapsid to stabilize its structure.

28

Empty capsids can form in the absence of RNA. The common characteristics of the RNA

29

encapsidation not only delineate the evolutionary relationship of negative strand RNA viruses,

30

but also provide insights to their mechanism of replication.

31

2

32

Importance

33 34

What separates NSVs from the rest of the virosphere is that the nucleocapsid of NSVs serves

35

as the template for viral RNA synthesis. Their viral RNA-dependent RNA polymerase (vRdRp)

36

can induce local conformational changes in the nucleocapsid to temporarily release the RNA

37

genome so that vRdRp can use it as the template for RNA synthesis during both transcription and

38

replication. After RNA synthesis at the local region is completed, vRdRp processes downstream

39

and the RNA genome is restored in the nucleocapsid. We found that the nucleocapsid assembly

40

of all NSVs shares three essential elements: a monomeric capsid protein protomer, parallel

41

orientation of subunits in the linear nucleocapsid, and a 5H+3H motif that forms a proper cavity

42

for sequestration of the RNA. This observation also suggests that all NSVs are evolved from a

43

common ancestor that has this unique nucleocapsid.

3

44

Introduction

45

All viruses contain a protein capsid that encapsidates the genomic polynucleotide. The capsid

46

is assembled with multiple copies of one or a few types of protein subunits following certain

47

symmetry. The most commonly studied symmetry is that of icosahedron which leads to spherical

48

or prolate virus particles (1-3). By contrast, helical symmetry is used for assembly of filamentous

49

capsids (4). The basic fold of the capsid protein subunit is found to be the same for a large

50

number of virus families, even though they may not be related to each other on a genomic basis.

51

The quintessential example of this is that of the β–barrel fold that was first found in small plant

52

RNA viruses, yet now has been discovered in at least 15 different viral families (5). The

53

architecture of the nucleocapsid is closely tied with the mechanism of replication in view of the

54

fact that the assembly of the capsid packages the viral genome.

55

For negative strand RNA viruses (NSVs), eight families have been recognized by the

56

International Committee on Taxonomy of Viruses (ICTV). The unique feature that distinguishes

57

NSV from the rest of the virosphere is that the nucleocapsid, instead of the naked genome, is

58

used as the template for viral nucleotide synthesis. It is indubitable that the assembly of the NSV

59

nucleocapsid is related to the unique mechanism of its viral RNA synthesis. In each of NSVs, the

60

nucleocapsid is packaged inside a lipid envelope. The appearance of the nucleocapsid inside the

61

envelope is different from virus to virus. In rhadboviruses, the nucleocapsid adopts a

62

characteristic bullet shape (6). In paramyxoviruses, the nucleocapsid is filamentous or

63

herringbone-like (7). In orthomyxoviruses, the nucleocapsid has a double helical structure (8, 9).

64

When the nucleocapsid is released from the virion, they all have the appearance of a coil (10).

65

The genomic RNA encapsidated in the nucleocapsid is protected from RNA nucleases to various

66

degrees depending on the structure. This protective structure renders the RNA not readily 4

67

accessible to the viral polymerase. Thus the viral polymerase must gain access to the

68

encapsidated RNA in order to carry out viral RNA synthesis. Since this is a common mechanism

69

of all NSVs, it is likely that the nucleocapsid of NSVs bears characteristic elements shared by all

70

NSV families. Recognition of these elements helps in defining essential viral functions and

71

revealing the underlining mechanism for NSV replication.

72

By systematic analyses of the known structures of the NSV nucleocapsid, we discovered the

73

common mechanism for genomic RNA encapsidation during replication. This unifying

74

mechanism suggests a common origin of NSV families and presents a clear picture for the

75

functions of the nucleocapsid protein.

76 77

Materials and Methods

78

The coordinates with PDB codes of 3PU4 (VSVN), 2WJ8 (RSVN), 4H5P (RVFVN) and

79

4BHH (LACVN) were retrieved from PDB (11-14). The superposition was carried out with Fr-

80

TM-align (15). The results were summarized in Table 1.

81

The NA320-24-2P protein complex was produced in E. coli as reported previously (16). X-ray

82

scattering data on VSV NA320-24-2P samples were collected on the SIBYLS beamline at the

83

Advanced Light Source (Figure 1A). Scattering curves were processed with PRIMUS (17).

84

Protein samples ranging from 0.75 to 4.3 mg/mL showed no sign of aggregation by Guinier plot

85

analysis (Figure 1B). Rg, Dmax and Porod volumes were calculated with the ATSAS package

86

(Table 2) (18). Bead models (ten in total) were generated from the scattering data with

87

DAMMIN (Figure 2) (19) and an averaged bead model was calculated with DAMAVER (20).

88

The average χ2 value for the bead models was 0.986. EOM (RANCH and GAJOE) was used to 5

89

build a multi-domain model of the NA320-24-2P complex against the SAXS data (18, 21).

90

Structures of the VSV P N-terminus (PNT, amino acids 6-34) (22), the dimeric P oligomerization-

91

domain (POD) (23), PCTD/N protein complex with bound PNT (24) and an additional unbound PCTD

92

were used as rigid bodies. The domain structures were derived from the structures with PDB

93

codes: 2FQM, 3HHW, and 3PMK. No partial restraints were imposed on the individual rigid

94

bodies. Fit of the bead and the multi-domain models to the experimental data are shown in

95

Figure 1E. Inspection of the fitted curves showed a dip in the calculated curve at low q values

96

suggesting that a larger organized species was in solution. Two copies of the EOM model were

97

fit to the experimental curve with SASREF (25). This fit to the curve was a remarkable match

98

yielding a χ value of 1.581 (single fit 12.865). The resulting models were superimposed onto the

99

bead models from DAMMIN with SUPCOMB (19, 26). The fit is shown in Figure 2A, B.

100 101

Results and Discussion

102

Capsid protomer

103

The atomic structure of nucleocapsid-like particles (NLPs) has been reported for three virus

104

families of NSV, Rhabdoviridae, Paramyxoviridae (Genus Pneumovirus), and Bunyaviridae

105

(Genera Phlebovirus and Orthobunyavirus) (11-14, 27-31). A comparison of representative

106

structures from each genus was performed. Since the structures in the same genus are highly

107

homologous, vesicular stomatitis virus (VSV) (11) is selected to represent rhabdoviruses;

108

respiratory syncytial virus (RSV) (12), pneumoviruses; rift valley fever virus (RVFV) (13),

109

phleboviruses; and La Crosse virus (LACV) (14), orthobunyaviruses. In each of these structures,

110

it was found that the encapsidated RNA is sequestered in a protein cavity. The capsid protein (N, 6

111

also known as nucleocapsid protein or nucleoprotein) is first synthesized as a monomeric protein

112

(named N0, the capsid protein protomer) before being incorporated in the nucleocapsid. N0

113

remains monomeric through a number of different ways. However, the fundamental requirement

114

to support viral replication is to prevent N0 from oligomerization before encapsidation of viral

115

RNA. The historic description that N0 is prevented from RNA binding is proven incorrect (16,

116

32). N0 is not an RNA binding protein but rather a capsid protein that assembles a capsid to

117

accommodate any RNA sequence. Any reported in vitro RNA binding measurements are mainly

118

for nonspecific electrostatic interactions between the negative charges of RNA phosphate groups

119

and the positively charged residues in N0. The nature of these nonspecific interactions has no

120

difference from that between any other positively charged protein, for instance, the matrix

121

protein of influenza virus, and RNA.

122

For rhabdoviruses, the N0 form is kept monomeric by forming a complex with the

123

phosphoprotein (P). To support viral replication continuously, it is required that the N and P

124

proteins are expressed in a 1:1 molar ratio (33, 34). A complex of N0-P was isolated from insect

125

cell expression of rabies virus (RABV) N and P proteins, which contains a N subunit and a dimer

126

of P (35). The same N0-P complex was also isolated for a mutant of the VSV N protein when the

127

mutant N protein was coexpressed with P in E. coli (16). The mutations in the VSV N protein

128

changed a stretch of five residues to Ala (NA320-24) and prevented formation of NLP. Analytical

129

unltracentrifugation experiments showed that the complex like the insect derived complex has a

130

1:2 (N to P) stoichiometry. This complex was studied by small angle X-ray scattering (SAXS)

131

techniques (Figure 1 and Table 2). The shape of the complex was determined by the ab initio

132

method in DAMMIN and is shown in Figure 2 (19). Independently, previously determined

133

crystal structures of N and domains of P were modeled against the scattering curves with EOM 7

134

(18, 21). In this method, the linkers between P protein domains were also modeled by an ab

135

initio approach yielding a complete model of the No-P2 protein complex. A complete model is

136

presented in Figure 3. The model suggests that the dimeric P is associated with N0 through

137

multiple sites of interactions. The observation that different parts from each monomer of the P

138

dimer contribute to N binding helps to explain studies where functions of the N-terminus

139

truncated P protein mutant can be restored when a C-terminus truncated P protein mutant is

140

provided in trans (36).

141

The P protein has a modular structure with a flexible N-terminal region, a structured

142

oligomerization domain in the middle and a compact C-terminal domain (PCTD). Both the N- and

143

C-terminal regions of P bind to the N protein. In one study, a complex was generated by binding

144

the N-terminal 60 residues of VSV P with the ring structure of a N mutant in which the N-

145

terminal 21 residues were deleted (22). The N-terminus of P interacts with the back of the C-

146

terminal half (C-lobe) of N. An α-helix corresponding to residues 17 to 31 of P was shown to

147

occupy the RNA cavity in the structure. Subsequent studies by Chen et al. (37) using a P3A

148

mutant (triple mutations of Ser or Thr to Ala) showed that phosphorylation of P at positions

149

Ser60, Thr62 and Ser64 is required to prevent N0 from encapsidating cellular RNA, which leads

150

to a dead-end product of N and diminishes viral replication. In the SAXS model, residues Ser60,

151

Thr62 and Ser64 were notably positioned to form charge interactions with Lys398, Arg399,

152

Lys414 and Lys417 of N (named here as the “PO binding site”). The crystal structure of the

153

VSV NLP in complex with PCTD showed P binds to the C-lobe of N, primarily the C-terminal

154

extended loop and α-helix 13 (24). Combining the results of SAXS and crystallography suggests

155

that P forms a stable complex with the monomeric N subunit involving interactions with the C-

156

terminal loop, a positively charged patch at the PO binding site, the RNA cavity, and the back of 8

157

the C-lobe of N. The oligomerization of N is prevented by binding of P whereby the binding site

158

for the N-terminus of P overlaps with that for the N-terminus of N in the nucleocapsid. The

159

region of P bound at the PO binding site can block the side-by-side contact between N subunits

160

found in the nucleocapsid. Mutation studies showed that loss of any of these interaction sites

161

could diminish N oligomerization, thus RNA encapsidation (16). Upon assembly of the

162

nucleocapsid, the N0 bound P is removed. N begins to oligomerize and encapsidate viral RNA.

163

The P protein is then recycled to bind another N0 subunit.

164

For RSV, there is no structural report of an N-P complex. However, it was shown that the N-

165

terminal 119 residues precede the oligomerization domain of P in primary sequence (38). The N-

166

terminus of RSVP is required for P binding to N0 (39). The N-terminal region of RSVP was

167

predicted to be intrinsically disordered, but residues 15-25 are predicted to be an α-helix. A

168

homologous structure was recently reported for a complex between the N-terminus of P and N of

169

Nipah virus (40). The N-terminus of P binds the C-terminal domain of N where it may exclude

170

binding of the C-terminus of a neighboring N subunit to this surface to prevent N

171

oligomerization. RSVP could bind to a similar site in RSVN, or another site of N which is

172

required for binding the N-terminus of a neighboring N subunit. The exact binding site for the N-

173

terminus of RSVP remains to be determined for RSVN.

174

For RVFV, the structure of the N protein has been solved in two forms, as an apo-structure

175

without RNA and an NLP. The first form of RVFVN has its N-terminal helix associated with its

176

own RNA encapsidation cavity (41). Residues 20 to 26 form a helix (α2). Within this helix, the

177

hydrophobic sidechain of Trp24 fits in the pocket where the bases of the RNA are sequestered in

178

the nucleocapsid, as observed in the RVFV NLP. In the oligomeric structure of the RVFV NLP,

179

the N-terminus of N is moved out of the RNA cavity to interact with the neighboring N subunit 9

180

(13). These observations suggest that the RVFVN protein is kept RNA-free by sequestering its

181

own N-terminus. During assembly, the N-terminus of RVFVN rearranges to support capsid

182

formation and RNA encapsidation. In another study, the structures of RVFV empty capsids have

183

also been solved (42). The empty capsids maintain the architecture of capsids with encapsidated

184

RNA. The reason for formation of empty capsids in this case is not understood. It is possible that

185

the N protein is not stable in the over-expressed bacterial milieu or an unknown protein, either

186

viral or cellular is necessary to help keep N from self-assembling. Lastly, this could be an artifact

187

resulted from producing the protein with an N-terminal thioredoxin-fusion.

188

For LACVN, several structures have been produced. The N-terminal 17 residues of N are

189

highly flexible and could be in a fold-back conformation in the N monomer, based on structures

190

of the N protein complex after RNA was removed (14). The fold-back conformation is similar to

191

sequestering the N-terminus in N of phleboviruses, but to a lesser extent. In LACV NLP, the N-

192

terminus of N assumes a conformation that extends to its neighboring N subunit. The extended

193

N-terminus has interactions with the bases and backbone of the encapsidated RNA, as well as

194

amino acid interactions with its neighboring N subunit. The C-terminus (residues 218-235) is

195

extended from the core of N and shows some conformational flexibility though not as much as

196

the N-terminus. The elements required for keeping the orthobunyaviruses N in a monomeric

197

form could be similar to those for N of phleboviruses.

198

The essential requirement for N to be competent in encapsidating viral RNA is to remain

199

monomeric. Different strategies may be used to achieve this, including occupying the sites

200

required for oligomerization by P binding, such as in the case of rhabdoviruses and

201

pneumoviruses, or sequestering the N-terminus by N itself, such as in the case of bunyaviruses.

202

In rhabdoviruses, the proximity of P binding to the RNA cavity is purely coincidental and only 10

203

for the formation of a stable N0-P complex. This complex is essential to prevent premature N

204

oligomerization rather than prevent RNA binding. This concept is supported by the results in

205

Chen et al. who showed that a phosphorylation deficient triple-mutant of P (P3A) that is mutated

206

outside of the cavity binding region could not prevent encapsidation of cellular RNA by VSVN

207

(37).

208

Architecture of the nucleocapsid

209

The N protein oligomerizes to encapsidate the single strand viral RNA. The N subunits are

210

associated with each other in a parallel orientation. A linear nucleocapsid is assembled with the

211

single strand viral RNA sequestered in the center. In rhabdoviruses, the nucleocapsid is a random

212

coil when it is isolated from the virion (43). The nucleocapsid is packaged into a superhelical

213

structure in the bullet shaped virion (6). The super symmetry is imposed on the nucleocapsid by

214

the matrix protein that has a 1:1 interaction with the N protein in the virion. The matrix protein

215

subunits have direct contacts among themselves to form the 2D helical mesh under the viral

216

membrane envelope. The same helical symmetry is adopted by the nucleocapsid when packaged

217

in the virion. The fact that the single strand viral RNA is completely encapsidated prior to being

218

packaged into the virion defines that the true symmetry of the nucleocapsid is linear. In most

219

NSVs, the nucleocapsid appears to be a random coil (10). The nucleocapsid in members of the

220

Paramyxovirinae contains a number of helical segments, but the helical symmetry is not strict in

221

terms of pitch and rotation of the N subunits (44). The Sendai virus nucleocapsid exists in at least

222

four different helical states (45). Since the repeating unit in the nucleocapsid is linear along the

223

encapsidated RNA and the helical symmetry is not required for RNA encapsidation, the exact

224

symmetry of the nucleocapsid of members in Paramyxovirinae strictly speaking must be

225

considered linear. In influenza virus (IFV), the linear nucleocapsid is twisted into a rough double 11

226

helix, with interactions between the two associated strings (8, 9). The helical superstructure is a

227

way to condense the nucleocapsid for packaging in the virion, but not essential for viral RNA

228

encapsidation.

229

The forces that stabilize the nucleocapsid involve extensive cross-molecular interactions

230

among the N subunits. There are side-by-side interactions between N subunits aligned in parallel.

231

These interactions are critical to capsid formation. The N protein can no longer assemble the

232

nucleocapsid if these interactions are disrupted by mutation (16). The extent of the side-by-side

233

interactions is different from virus to virus ranging from below 200 Å2 for LACVN up to ~2200

234

Å2 for VSV. There is also a disparity to the amount of buried surface between adjacent C-

235

terminal domains versus N-terminal domains of each viral N (11). The contact areas between

236

neighboring domains in different NLPs are listed in Table 3. The contact areas must have

237

plasticity because these calculated contact areas are derived from an artificial ring structure. In

238

the authentic linear nucleocapsid, these contact areas are likely to be different. The difference

239

between the two halves of the N protein may have some functional implications because the N

240

protein needs to undergo a conformational change to reveal the sequestered template RNA

241

during viral RNA synthesis. It is likely that the domain that has a lesser contact area with the

242

neighboring domains should be opened to reveal the RNA. As shown in Table 3, the N-terminal

243

domain of VSVN has a lesser contact area and is likely to be open during viral RNA synthesis,

244

while the C-terminal domain remains associated to maintain the integrity of the nucleocapsid.

245

Rearrangement of the N-terminal domain to open the cavity is further supported in lieu of the

246

greater association between the C-terminal domains due to additional interactions, as noted

247

below.

12

248

In addition to the side-by-side contacts, the most obvious interactions between the N subunits

249

are from the structural elements extended from the core of the N subunit. The core of N and the

250

nucleotides associated with the core act as one structural unit. The number of nucleotides

251

accommodated in the core is listed in Table 3. The registration of each N subunit may be defined

252

by choosing one subunit as the origin (position 0). If the access to the RNA cavity faces away

253

from the reader, -1 refers to the subunit on the left of the origin subunit (5’ end of the

254

encapsidated RNA), and +1, on the right (3’end of the encapsidated RNA). In rhabdoviruses, the

255

N-terminal arm (21 residues) extends away from the core to interact with the -1 N-terminal

256

domain (N-lobe), whereas an extended loop in the C-lobe interacts with the +1 C-lobe. The N-

257

terminal arm also interacts with the extended loop in the -2 C-lobe. In RSVN, the N-terminal arm

258

(28 residues) interact with both N- and C-terminal domains of the -1 subunit, whereas the

259

extended C-terminal arm, rather than a loop in this case, interacts with the C-terminal domain of

260

the +1 N subunit. In RVFVN, only the N-terminal arm (32 residues) interacts with the -1 N-

261

terminal domain. In LACVN, the N-terminal arm (17 residues) interacts with the -1 N-terminal

262

domain involving only the first 8 residues. The remaining residues interact with the encapsidated

263

RNA. The C-terminal arm of LACVN (18 residues) contains an α-helix that interacts with the C-

264

terminal domain of subunit +1. Residues that are involved in capsid formation are illustrated in

265

Figure 4.

266

In IFVN, there is also an extended loop in the C-terminal region that interacts with the C-

267

terminal domain in the +1 N subunit. It is not clear if the N-terminal region is involved in the

268

interactions with the neighboring subunits in the nucleocapsid. An atomic model of the

269

nucleocapsid or an NLP is needed to fully address this. It is also not clear how IFVN is

270

maintained in a monomeric form prior to RNA encapsidation. There is no report of any viral 13

271

protein that is involved in this function. In the reported structure of IFVN, the N-terminus of N

272

could be in a sequestered conformation. If this is the case, IFVN may employ a method similar to

273

that used by RVFVN and LACVN to remain monomeric. The extended loop in the C-terminal

274

domain folds back on the surface of its own monomer as shown by the structure of an obligate

275

monomeric mutant of IFVN (46). This folded conformation of the C-terminal extended loop

276

should also contribute to maintaining a monomeric mode of IFVN prior to RNA encapsidation.

277

The surface contact and domain-swap interactions between neighboring N subunits are

278

essential for the assembly of NSV nucleocapsids. Similar interactions are also common in the

279

nucleocapsid of other viruses, such as picornaviruses and adenovirus (47, 48). The only unique

280

feature in the NSV nucleocapsid is that these interactions are linear along the encapsidated single

281

strand viral RNA.

282

Sequestered RNA in the cavity

283

The nucleocapsid of NSVs needs to be capable of encapsidating all possible RNA sequences.

284

The mechanism for specific encapsidation of viral RNA may be that viral RNA encapsidation is

285

concomitant with viral replication, likely to be at the site of the viral RNA replication complex

286

(24). At this point, the monomeric N protein assembles the nucleocapsid simultaneous with

287

encapsidation of the genome. No consensus interactions with the bases of the encapsidated RNA

288

were identified among all the reported NLP structures, including the cases in which

289

homogeneous sequences were encapsidated in the NLP (49). Interactions of the backbone

290

phosphate groups with positively charged sidechains were found for a fraction of the sequestered

291

phosphate groups, but there is not a conserved pattern among NLPs of different viruses. This

292

holds true for even the closely related NLP of VSV and RABV (32) and suggests that the

14

293

positively charged residues in the RNA cavity are not the key factor responsible for RNA

294

encapsidation.

295

What is unique about the viral RNA in the NSV nucleocapsid is that the sequestered bases

296

are stacked to form a motif similar to one-half of the A-form double helix of RNA (Figure 5A).

297

In rhabdoviruses, the bases of nucleotides 1-4 are stacked and face the solvent side of the RNA

298

cavity, whereas bases of nucleotides 5, 7 and 8 are stacked and face the interior of the N subunit.

299

Similarly in RSV, bases of nucleotides 2-4 are stacked and face the solvent side of the RNA

300

cavity, whereas bases of nucleotides 5-7 are stacked and face the interior of the N subunit. There

301

is an additional “linker” nucleotide between the N subunits in rhabdoviruses and RSV. This

302

nucleotide is still sequestered from solvent accessibility when N subunits oligomerize in the

303

nucleocapsid, but it may allow some flexibility between the N subunits. The base of this linker

304

nucleotide can be stacked with the other bases, or an aromatic sidechain will take its place to

305

stack with the other bases when the linker nucleotide is in a transitional conformation (11, 12). In

306

the RVFV NLP, four bases are sequestered by the core domains each facing the interior of the N

307

subunit. Among these four bases, the two in the center are stacked. The bases of the three

308

“linker” nucleotides are also in a stacked conformation and protected by the capsid structure. In

309

the LACV NLP, seven bases near the 5’ end (nucleotides -11 and 1-6) of the RNA strand are

310

stacked and face the entrance to the RNA cavity. The aromatic sidechain of tyrosine-177 is

311

intercalated between the bases of nucleosides 7 and 8, and the base of nucleotide 8 is further

312

stacked with the base of nucleotide 9. The bases in this triple stacking face the RNA cavity.

313

Lastly, the linker nucleotide 10 is sequestered in the RNA cavity, but its base is not stacked. The

314

unique stacking arrangements observed in NSVs allow for maximized packaging of the RNA in

315

the capsid. 15

316

The base stacking is only stabilized when the viral RNA is in the cavity of the nucleocapsid.

317

The N subunits can form a stable empty capsid without RNA (16).The encapsidated RNA is a

318

resident in the nucleocapsid, perhaps also contributing to the overall stability of the

319

nucleocapsid. The shape and the electrostatic environment of the cavity define how base stacking

320

occur. The stability of the base-stacking motifs in the RNA cavity is dependent on the RNA

321

sequence, with poly(rA) being the most stable and poly(rU) being the least stable (49). Specific

322

sequences in the sequestered RNA genome regulate many viral functions, such as transcriptional

323

initiation and termination. In terms of transcriptional termination, there is a highly conserved U7

324

track at the end of each coding region in rhabdoviruses. The structure in this region is the least

325

stable in the nucleocapsid. It is conceivable that such instability could promote dissociation of

326

the viral transcription complex. Mutations in the N protein have been shown to alter the level of

327

mRNA synthesis. In this case, genomic mutations have subsequently occurred resulting in an

328

extension of the U track to U8 in rescued VSV (50). U rich sequences are also found in the

329

intergenic regions in other NSV genomes, but not all.

330

The RNA cavity is formed by secondary structural elements from the N- and C-domains in

331

the N protein. A highly conserved (5H+3H) motif consisting of 5 helices from the N-domain and

332

3 helices from the C-domain has been identified to constitute the RNA cavity (51). When the

333

structures of VSVN and RSVN are superimposed using the Fr-TM-align method (15), the

334

(5H+3H) motif from each N protein (α4-α8 in the N-domain, and α9-α11 in the C-domain of the

335

VSVN protein (11); αN3, αN5-αN8, and αC1, η2, and αC3 of the RSVN protein (12)) could be

336

superimposed as a rigid body (Figure 5B, 5C and Table 1). η indicates a 310 helix. In addition,

337

two helices following the (5H+3H) motif in the C-domain could also be superimposed as part of

338

the core in these two N proteins (Figure 5B, 5C). In the case of RVFVN, only the first four 16

339

helices (α3, α5-α7 in RVFVN (13)) in the (5H+3H) motif may be aligned with those in VSVN or

340

RSVN (Figure 5B, 5C and Table 1). The 5th helix of the motif is present but moved away from

341

the RNA cavity, inducing reorientation of the following three helices in the (5H+3H) motif, as

342

well as helices following the motif. As a result, this RNA cavity is almost completely collapsed

343

compared to that in VSVN or RSVN. This small RNA cavity could only accommodate the bases

344

of four nucleotides, while excluding the ribose-phosphate backbone. LACVN (14) has the same

345

topology as RVFVN (Figure 5B, 5C and Table 1). The first four helices of LACVN in the

346

(5H+3H) motif (α2, η1, α3, and α4) correspond to those in RVFVN. However, there are two

347

noticeable differences between the two structures. One is that the linker between the first and

348

second helices is changed from a helix (α4) to a β–hairpin (β1-4). The other is that the third helix

349

in RVFVN (α4) is much shorter than that in LACVN (α3). The three helices in the C-terminal

350

domain of the (5H+3H) motif are the same in RVFVN and LACVN, as well as three additional

351

helices following the motif, in terms of topology. In addition, LACVN has an extra helix at the

352

C-terminus (α11). This helix is involved in interactions with the neighboring N subunit on the

353

right side (+1), as described above.

354

Summary

355

The nucleocapsid of NSVs is assembled by the universal assembly principle of all viral

356

nucleocapsids. Oligomerization of the protein subunits forms the capsid that encapsidates the

357

viral polynucleotide. Extensive interactions cross protein subunits, including domain-swaps.

358

These interactions are involved in the stabilization of the nucleocapsid. What is unique about

359

NSVs is that their nucleocapsid has a linear symmetry. The viral RNA is a string in the center of

360

the nucleocapsid ribbon. The nucleocapsid assembly of all NSVs shares three essential elements:

361

a monomeric capsid protein protomer, parallel orientation of subunits in the nucleocapsid, and 17

362

the 5H+3H motif that forms a proper cavity for sequestration of the RNA. During viral RNA

363

synthesis, the nucleocapsid does not disassemble to completely release the viral RNA template.

364

Instead, conformational changes must occur in protein domains to temporarily release the RNA

365

template on a local rather than global basis. The viral polymerase complex is a likely candidate

366

to induce this conformational change in order to gain access to the sequestered bases, some of

367

which face the interior of the N subunits. Once revealed, the RNA template is available for

368

transcription and genomic replication. After the viral polymerase complex passes, the RNA is

369

tidily repositioned in the encapsidation cavity and the integrity of the nucleocapsid is restored.

370

This new mechanism of RNA encapsidation allows the N protein to play a role in viral

371

replication and transcription. Mutations in the N protein may increase or decrease the level of

372

viral replication (52, 53). More interestingly, some of these mutations in the N protein did not

373

change the level of viral replication, but changed the level of transcription. This suggests that the

374

structure of the nuclecapsid itself has a role in viral transcription that is independent of

375

replication. For viral replication, the viral polymerase complex needs to load at the 3’ end of the

376

genome and process through the nucleocapsid. For viral transcription, on the other hand, the viral

377

polymerase complex needs to recognize the promoter and the transcription termination

378

sequences, in addition to loading and processivity. There may be a unique structural feature

379

associated with the regions where the promoter or the transcription termination sequence is

380

located in the nucleocapsid to allow such recognition.

381

Acknowledgement

382

We thank Dr. Mark Walter for assistance in collecting SAXS data. SAXS experiments were

383

conducted at the Advanced Light Source (ALS), a national user facility operated by Lawrence

18

384

Berkeley National Laboratory on behalf of the Department of Energy, Office of Basic Energy

385

Sciences, through the Integrated Diffraction Analysis Technologies (IDAT) program, supported

386

by DOE Office of Biological and Environmental Research. Additional support comes from the

387

National Institute of Health project MINOS (R01GM105404). The work is supported in part by a

388

NIH grant 1R01AI10630 to ML and 1R56AI01087 to TG. References

389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422

1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12.

13. 14. 15. 16.

Prasad BV, Schmid MF. 2012. Principles of virus structural organization. Advances in experimental medicine and biology 726:17-47. Poranen MM, Bamford DH. 2012. Assembly of large icosahedral double-stranded RNA viruses. Advances in experimental medicine and biology 726:379-402. Ganser-Pornillos BK, Yeager M, Pornillos O. 2012. Assembly and architecture of HIV. Advances in experimental medicine and biology 726:441-465. Stubbs G, Kendall A. 2012. Helical viruses. Advances in experimental medicine and biology 726:631-658. Abrescia NG, Bamford DH, Grimes JM, Stuart DI. 2012. Structure unifies the viral universe. Annual review of biochemistry 81:795-822. Ge P, Tsao J, Schein S, Green TJ, Luo M, Zhou ZH. 2010. Cryo-EM model of the bullet-shaped vesicular stomatitis virus. Science 327:689-693. Loney C, Mottet-Osman G, Roux L, Bhella D. 2009. Paramyxovirus ultrastructure and genome packaging: cryo-electron tomography of sendai virus. Journal of virology 83:8191-8197. Arranz R, Coloma R, Chichon FJ, Conesa JJ, Carrascosa JL, Valpuesta JM, Ortin J, Martin-Benito J. 2012. The structure of native influenza virion ribonucleoproteins. Science 338:1634-1637. Moeller A, Kirchdoerfer RN, Potter CS, Carragher B, Wilson IA. 2012. Organization of the influenza virus replication machinery. Science 338:1631-1634. Ruigrok RW, Crepin T, Kolakofsky D. 2011. Nucleoproteins and nucleocapsids of negative-strand RNA viruses. Current opinion in microbiology 14:504-510. Green TJ, Zhang X, Wertz GW, Luo M. 2006. Structure of the vesicular stomatitis virus nucleoprotein-RNA complex. Science 313:357-360. Tawar RG, Duquerroy S, Vonrhein C, Varela PF, Damier-Piolle L, Castagne N, MacLellan K, Bedouelle H, Bricogne G, Bhella D, Eleouet JF, Rey FA. 2009. Crystal structure of a nucleocapsidlike nucleoprotein-RNA complex of respiratory syncytial virus. Science 326:1279-1283. Raymond DD, Piper ME, Gerrard SR, Skiniotis G, Smith JL. 2012. Phleboviruses encapsidate their genomes by sequestering RNA bases. Proc Natl Acad Sci U S A 109:19208-19213. Reguera J, Malet H, Weber F, Cusack S. 2013. Structural basis for encapsidation of genomic RNA by La Crosse Orthobunyavirus nucleoprotein. Proc Natl Acad Sci U S A 110:7246-7251. Pandit SB, Skolnick J. 2008. Fr-TM-align: a new protein structural alignment method based on fragment alignments and the TM-score. BMC bioinformatics 9:531. Zhang X, Green TJ, Tsao J, Qiu S, Luo M. 2008. Role of intermolecular interactions of vesicular stomatitis virus nucleoprotein in RNA encapsidation. J Virol 82:674-682.

19

423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469

17. 18.

19. 20. 21.

22.

23. 24.

25. 26. 27.

28.

29.

30.

31.

32. 33. 34.

Konarev PV, Volkov VV, Sokolova AV, Koch MHJ, Svergun DI. 2003. PRIMUS - a Windows-PC based system for small-angle scattering data analysis. J. Appl Cryst 36:1277-1282. Petoukhov MV, Franke D, Shkumatov AV, Tria G, Kikhney AG, Gajda M, Gorba C, Mertens HDT, Konarev PV, Svergun DI. 2012. New developments in the ATSAS program package for smallangle scattering data analysis. J. Appl. Cryst. 45:342-350. Svergun DI. 1999. Restoring low resolution structure of biological macromolecules from solution scattering using simulated annealing. Biophys J 76:2879-2886. Volkov VV, Svergun DI. 2003. Uniqueness of ab initio shape determination in small-angle scattering. J. Appl. Cryst. 36:860-886. Bernado P, Mylonas E, Petoukhov MV, Blackledge M, Svergun DI. 2007. Structural characterization of flexible proteins using small-angle X-ray scattering. Journal of the American Chemical Society 129:5656-5664. Leyrat C, Yabukarski F, Tarbouriech N, Ribeiro EA, Jr., Jensen MR, Blackledge M, Ruigrok RW, Jamin M. 2011. Structure of the vesicular stomatitis virus N0-P complex. PLoS Pathog 7:e1002248. Ding H, Green TJ, Lu S, Luo M. 2006. Crystal structure of the oligomerization domain of the phosphoprotein of vesicular stomatitis virus. J Virol 80:2808-2814. Green TJ, Luo M. 2009. Structure of the vesicular stomatitis virus nucleocapsid in complex with the nucleocapsid-binding domain of the small polymerase cofactor, P. Proc Natl Acad Sci U S A 106:11713-11718. Petoukhov MV, Svergun DI. 2005. Global rigid body modeling of macromolecular complexes against small-angle scattering data. Biophys J 89:1237-1250. Kozin M, Svergun DI. 2001. Automated matching of high- and low-resolution structural models. J. Appl. Cryst. 34:33-41. Albertini AA, Wernimont AK, Muziol T, Ravelli RB, Clapier CR, Schoehn G, Weissenhorn W, Ruigrok RW. 2006. Crystal structure of the rabies virus nucleoprotein-RNA complex. Science 313:360-363. Li B, Wang Q, Pan X, Fernandez de Castro I, Sun Y, Guo Y, Tao X, Risco C, Sui SF, Lou Z. 2013. Bunyamwera virus possesses a distinct nucleocapsid protein to facilitate genome encapsidation. Proc Natl Acad Sci U S A 110:9048-9053. Jiao L, Ouyang S, Liang M, Niu F, Shaw N, Wu W, Ding W, Jin C, Peng Y, Zhu Y, Zhang F, Wang T, Li C, Zuo X, Luan CH, Li D, Liu ZJ. 2013. Structure of severe fever with thrombocytopenia syndrome virus nucleocapsid protein in complex with suramin reveals therapeutic potential. J Virol 87:6829-6839. Ariza A, Tanner SJ, Walter CT, Dent KC, Shepherd DA, Wu W, Matthews SV, Hiscox JA, Green TJ, Luo M, Elliott RM, Fooks AR, Ashcroft AE, Stonehouse NJ, Ranson NA, Barr JN, Edwards TA. 2013. Nucleocapsid protein structures from orthobunyaviruses reveal insight into ribonucleoprotein architecture and RNA polymerization. Nucleic Acids Res 41:5912-5926. Dong H, Li P, Bottcher B, Elliott RM, Dong C. 2013. Crystal structure of Schmallenberg orthobunyavirus nucleoprotein-RNA complex reveals a novel RNA sequestration mechanism. Rna 19:1129-1136. Luo M, Green TJ, Zhang X, Tsao J, Qiu S. 2007. Conserved characteristics of the rhabdovirus nucleoprotein. Virus Res 129:246-251. Peluso RW, Moyer SA. 1988. Viral proteins required for the in vitro replication of vesicular stomatitis virus defective interfering particle genome RNA. Virology 162:369-376. Howard M, Wertz G. 1989. Vesicular stomatitis virus RNA replication: a role for the NS protein. J Gen Virol 70:2683-2694.

20

470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516

35.

36.

37.

38.

39.

40.

41.

42.

43.

44.

45. 46. 47.

48. 49. 50.

51.

Mavrakis M, Iseni F, Mazza C, Schoehn G, Ebel C, Gentzel M, Franz T, Ruigrok RW. 2003. Isolation and characterisation of the rabies virus N degrees-P complex produced in insect cells. Virology 305:406-414. Chen M, Ogino T, Banerjee AK. 2007. Interaction of vesicular stomatitis virus P and N proteins: identification of two overlapping domains at the N terminus of P that are involved in N0-P complex formation and encapsidation of viral genome RNA. J Virol 81:13478-13485. Chen L, Zhang S, Banerjee AK, Chen M. 2013. N-terminal phosphorylation of phosphoprotein of vesicular stomatitis virus is required for preventing nucleoprotein from binding to cellular RNAs and for functional template formation. J Virol 87:3177-3186. Llorente MT, Garcia-Barreno B, Calero M, Camafeita E, Lopez JA, Longhi S, Ferron F, Varela PF, Melero JA. 2006. Structural analysis of the human respiratory syncytial virus phosphoprotein: characterization of an alpha-helical domain involved in oligomerization. J Gen Virol 87:159-169. Mallipeddi SK, Lupiani B, Samal SK. 1996. Mapping the domains on the phosphoprotein of bovine respiratory syncytial virus required for N-P interaction using a two-hybrid system. J Gen Virol 77 ( Pt 5):1019-1023. Yabukarski F, Leyrat C, Tarbouriech N, Jensen MR, Blackledge M, Ruigrok R, Jamin M. 2013. Crystal structure of the N0-P complex of Nipah virus and of VSV provide new insights into the encapsidation mechanism. Proceedings of XV International Conference on Negative Strand RNA Viruses:84. Raymond DD, Piper ME, Gerrard SR, Smith JL. 2010. Structure of the Rift Valley fever virus nucleocapsid protein reveals another architecture for RNA encapsidation. Proc Natl Acad Sci U S A 107:11769-11774. Ferron F, Li Z, Danek EI, Luo D, Wong Y, Coutard B, Lantez V, Charrel R, Canard B, Walz T, Lescar J. 2011. The hexamer structure of Rift Valley fever virus nucleoprotein suggests a mechanism for its assembly into ribonucleoprotein complexes. PLoS Pathog 7:e1002030. Blocquel D, Bourhis J-M, Éléouët J-F, Gerlier D, Habchi J, Jamin M, Longhi S, Yabukarski F. 2012. Transcription et réplication des Mononegavirales : une machine moléculaire originale. Virologie 16:3. Bhella D, Ralph A, Yeo RP. 2004. Conformational flexibility in recombinant measles virus nucleocapsids visualised by cryo-negative stain electron microscopy and real-space helical reconstruction. J Mol Biol 340:319-331. Egelman EH, Wu SS, Amrein M, Portner A, Murti G. 1989. The Sendai virus nucleocapsid exists in at least four different helical states. J Virol 63:2233-2243. Chenavas S, Estrozi LF, Slama-Schwok A, Delmas B, Di Primo C, Baudin F, Li X, Crepin T, Ruigrok RW. 2013. Monomeric nucleoprotein of influenza A virus. PLoS Pathog 9:e1003275. Rossmann MG, Arnold E, Erickson JW, Frankenberger EA, Griffith JP, Hecht HJ, Johnson JE, Kamer G, Luo M, Mosser AG, et al. 1985. Structure of a human common cold virus and functional relationship to other picornaviruses. Nature 317:145-153. Liu H, Jin L, Koh SB, Atanasov I, Schein S, Wu L, Zhou ZH. 2010. Atomic structure of human adenovirus by cryo-EM reveals interactions among protein networks. Science 329:1038-1043. Green TJ, Rowse M, Tsao J, Kang J, Ge P, Zhou ZH, Luo M. 2011. Access to RNA encapsidated in the nucleocapsid of vesicular stomatitis virus. J Virol 85:2714-2722. Harouaka D, Wertz GW. 2012. Second-site mutations selected in transcriptional regulatory sequences compensate for engineered mutations in the vesicular stomatitis virus nucleocapsid protein. J Virol 86:11266-11275. Luo M, Green TJ, Zhang X, Tsao J, Qiu S. 2007. Structural comparisons of the nucleoprotein from three negative strand RNA virus families. Virol J 4:72-78.

21

517 518 519 520 521 522 523 524 525 526 527 528 529

52.

53.

54. 55. 56. 57.

Nayak D, Panda D, Das SC, Luo M, Pattnaik AK. 2009. Single-amino-acid alterations in a highly conserved central region of vesicular stomatitis virus N protein differentially affect the viral nucleocapsid template functions. J Virol 83:5525-5534. Harouaka D, Wertz GW. 2009. Mutations in the C-terminal loop of the nucleocapsid protein affect vesicular stomatitis virus RNA replication and transcription differentially. J Virol 83:1142911439. Svergun DI. 1992. Determination of the regularization parameter in indirect-transform methods using perceptual criteria. . J. Appl. Crystal. 25:495-503. Volkov VV, Svergun DI. 2003. Uniqueness of ab initio shape determination in small-angle scattering. Journal of Applied Crystallography 36:860-864. Kozin MB, Svergun DI. 2001. Automated matching of high- and low-resolution structural models. Journal of Applied Crystallography 34:33-41. PyMol. The PyMOL Molecular Graphics System, Version 1.3, Schrödinger, LLC.

530 531 532

22

533

Figure legends

534

Figure 1. Small-angle X-ray scattering (SAXS) analysis of the N0–P complex from VSV. A)

535

The experimental SAXS profile for three concentrations of the protein complex, 0.75 (black), 2.4

536

(blue) and 4.3 (red) mg/mL. Scattering curves were scaled with PRIMUS (17). The

537

concentrations and associated coloring schemes are used in (A) – (D). B) A Guinier plot of the

538

experimental SAXS profile with the fit shown. Plots are shown on a relative scale. C) A Kratky

539

plot. D) The pair-distribution function as calculated in GNOM (54).

540

Figure 2. Ab initio shape modeling of the VSV N0-P complex. A) DAMMIN (19) was used

541

to produce 10 bead models. These 10 models were used to produce an averaged model with

542

DAMAVER (55). The averaged model (white spheres) is shown with aligned “dimer” of the N0-

543

P complex (red and blue). The dimer was determined with SASREF (25) and superimposed with

544

SUPCOMB (56). 90° rotations relating each orientation are noted. B) A representative single

545

bead model from DAMMIN is shown in beads of cyan. The orientations are the same as in (A).

546

The protein model is not adjusted to fit the single DAMMIN bead model. C) The plot of a single

547

DAMMIN model against the experimental scattering curve. D) Fit of the N0-P model shown in

548

Figure 3 to the experimental SAXS curve. E) SASREF (25) was used to rigid body fit two copies

549

of the complex against the SAXS data. The fit of the SASREF derived dimer is shown.

550 551 552

Figure 3. Maintaining a monomeric N. A complete multi-domain model of theN0-P complex

553

from VSV as determined with EOM is presented. The N protein is shown in red, while the

554

monomers of the P protein dimer are colored, yellow and green. Previously determined domains 23

555

are shown in cartoon representation, while ab initio generated loops are shown as C-α ribbons.

556

The three orange spheres represent positions of Ser60, Thr62 and Ser64 (sites of

557

phosphorylation) of the P protein. These residues have a proximity to the negatively charged

558

patch formed by residues Lys398, Arg399, Lys414 and Lys417 of the N protein (colored dark

559

blue on the surface). This region is circled and denoted the “PO binding site”. All cartoon

560

drawings in this and following figures were prepared with PyMol (57).

561 562

Figure 4. Oligomerization of the N subunits in the nucleocapsid. Surface representations of

563

the nucleocapsid proteins of (A) VSV, (B) RSV, (C) RVFV and (D) LACV are shown for three

564

subunits. Each subunit has been radially spaced to expose the surfaces that contribute to subunit

565

interactions. The monomers are colored red, green and yellow while residues that contact the

566

adjacent protomers are shaded in black.

567 568

Figure 5. Comparisons of the RNA cavity. (A) Ribbon drawings that show the N subunit

569

with encapsidated RNA for VSV, RSV, RVFV and LACV. The rainbow color code of the

570

polypeptide is blue for the N-terminus to red for the C-terminus. The encapsidated RNA is

571

shown again below each N subunit as stick and ribbon models. The nucleotides are numbered

572

from 5’ to 3’. The number 1’ for RVFV means that this nucleotide is the equivalent of nucleotide

573

1 in the next N subunit in the NLP. (B) Superposition of the N proteins. Cα tracing of the aligned

574

coordinates was shown. In the first three panels, the Cα tracing for VSVN was shown in gray.

575

The superimposed Cα tracing was colored in rainbow from N-terminus (blue) to C-terminus

576

(red). In the fourth panel, the Cα tracing for LACVN was shown in skyblue; RVFVN, in 24

577

rainbow. Only the aligned portions are shown (see Table 1). (C) Topological cartoons

578

representing the helices in the N protein core. Each circle represents a helix that is labeled

579

according to their secondary structure assignment in the references (11-14).

580

25

Table 1. Superposition of the N proteins

RSVN(2-375) vs VSVN(2-422) RVFVN(49-126) vs VSVN(131-206) LACVN(35-124) vs VSVN(131-206) LACVN(35-124) vs RVFVN(49-126)

Residues aligned

RMSD (Å2)

TM-score

292(421) 53(76) 50(76) 55(78)

4.97 3.13 4.08 3.88

0.52 0.44 0.35 0.40

Residue number in parenthesis is the length of the reference structure. Table 2. Radius of gyration (Rg), maximum size (Dmax) and Porod volume (Vp) as calculated from the SAXS curves for the VSV N0-P complex. concentration (mg/mL) 0.75 2.4 4.3

Rg (Å) 60.94 65.07 66

Vp (Å3) 5.82 x 105 6.27 x 105 6.52 x 105

Dmax (Å) 209 227.5 229

Rg was calculated with AUTORG, Dmax with DATGNOM and Vp with DATPOROD. All of which are from the ATSAS package (18). Table 3. Area of contacts between the N proteins and number of nucleotides bound in the N protein Surface Virus Area/Protomer 23093.9 VSV 20134.4 RSV 13884.2 RVFV 13775.4 LACV

Buried Interface Complex (Å2) 5648.0 5117.6 3576.6 1982.7

Buried Interface Core Buried Interface N-lobe (Å2) Core C-lobe (Å2) 552.4 1572.4 997.1 248.0 344.6 165.0 63.9

581

26

Buried Interface Arms/Loops (Å2) 3523.2 3872.5 3232.0 1753.8

# of nucleotides in the core 8 6 4 10

Common mechanism for RNA encapsidation by negative-strand RNA viruses.

The nucleocapsid of a negative-strand RNA virus is assembled with a single nucleocapsid protein and the viral genomic RNA. The nucleocapsid protein po...
644KB Sizes 3 Downloads 0 Views