JGV Papers in Press. Published February 28, 2014 as doi:10.1099/vir.0.057984-0

1

Journal category: Animal Viruses - Double-strand RNA

2

Genome sequencing identifies genetic and antigenic divergence of porcine

3

picobirnaviruses

4

Krisztián Bányai 1 , Christiaan Potgieter 2,3, Ákos Gellért 4, Balasubramanian Ganesh

5

5

, Maria Tempesta 6, Eleonora Lorusso 6, Canio Buonavoglia 6, Vito Martella 6

6 7

1

8

Academy of Sciences, Budapest, Hungary

9

2

Virology Division, Onderstepoort Veterinary Institute, Onderstepoort, South Africa

10

3

Biochemistry Division, North-West University, Potchefstroom, South Africa

11

4

Agricultural Institute, Centre for Agricultural Research, Hungarian Academy of Sciences,

12

Martonvásár, Hungary

13

5

14

West Bengal, India

15

6

Institute for Veterinary Medical Research, Centre for Agricultural Research, Hungarian

Division of Virology, National Institute of Cholera and Enteric Diseases (NICED), Kolkata,

Dipartimento di Medicina Veterinaria, Università di Bari Aldo Moro, Bari, Italy

16 17 18 19 20

Correspondence to:

Krisztián Bányai, Institute for Veterinary Medical Research, Centre for Agricultural Research, Hungarian Academy of Sciences, Hungária krt 21., H-1143 Budapest, Hungary; E-mail: [email protected] Vito Martella, Dipartimento di Medicina Veterinaria, Università di Bari Aldo Moro, S.p. per Casamassima Km 3, 70010 Valenzano, Bari, Italy; E-mail: [email protected]

21 22 23 24 25 26

Running title: Porcine picobirnavirus genome

27

Word count in summary: 193

28

Word count in text: 2570

29 30

GenBank accession numbers: KF861767-KF861773

1

31

SUMMARY

32

The full-length genome sequence of a porcine picobirnavirus detected in Italy in 2004

33

was determined. The S segment was 1730 nucleotide (nt) in length, coding for a

34

putative RNA-dependent RNA polymerase. Two distinct sub-populations of L

35

segment (LA and LB) were identified in the sample, with the sizes ranging from 2351

36

to 2666 nt. The ORF1 coding for a protein of unknown function, contained repetitions

37

of the ExxRxNxxxE motif in variable number. The capsid protein coding ORF2

38

spanned nt 810-2447 in the LB variants and started at nt 734 in the LA variants.

39

However, a termination codon was present only in one of all LA segment variants.

40

Three dimensional modelling of the porcine PBV capsids suggested structural

41

differences in the protruding domain, tentatively involved as antigens in humoral

42

immune response. Altogether, these findings suggest simultaneous presence of two

43

different picobirnavirus strains sharing the same S segment but displaying genetically

44

diverse L segments. In addition, the sample probably contained a mixture of PBVs

45

with aberrant RNA replication products. Altered structure in the L segments could be

46

tolerated and retained in the presence of functionally integer cognate genes and

47

represent a mechanism of virus diversification.

48

Keywords: Picobirnaviridae, reassortment, pig, zoonosis

2

49

TEXT

50

Picobirnaviruses (PBVs), family Picobirnaviridae, are small (35–41 nm in diameter),

51

non-enveloped viruses displaying a distinctive icosahedral capsid. The PBV genome

52

is bisegmented double-stranded RNA (dsRNA). The larger (L) genome segment is

53

2.2–2.7 kb in length and encodes the capsid protein and a putative protein of

54

unknown function, while the smaller (S) genome segment is 1.2–1.9 kb in length and

55

encodes the viral RNA-dependent RNA polymerase (Delmas, 2011). Nucleotide

56

sequence differences in the S segment permit the classification of PBVs into two

57

major clusters, designated as genogroup I and genogroup II. The intra- and

58

intergenogroup sequence similarities of a short nucleic acid fragment of the S

59

segment used in diagnostic PCR ranges from 49 to 100 % and 28 to 37 %,

60

respectively (Bányai et al., 2003; Rosen et al., 2000).

61

PBVs have been found in the feces and intestinal content and, most recently, in the

62

respiratory tract of vertebrates. While the list of host species with record of PBV

63

infections has increased continuously (Fregolente et al., 2009; Gallimore et al., 1995;

64

Ganesh et al., 2011; Ghosh et al., 2009; Gillman et al., 2013; Green et al., 1999;

65

Haga et al., 1999; Malik et al., 2011; Masachessi et al., 2007, 2012; Pereira et al.,

66

1988; Wang et al., 2007; Woo et al., 2012), an association of PBV infection with

67

enteric or respiratory diseases has not been formally demonstrated. In fact, PBV is

68

often detected together with co-infecting major enteric pathogens or in clinical

69

specimens from inapparently infected hosts. The lack of cell culture or animal model

70

further hinders the recognition of any disease associations. However, studies in

71

immuncompromised hosts have suggested an opportunistic role for PBV in diarrheic

72

infections (Giordano et al., 1998; Grohmann et al., 1993).

73

Transmission of PBV from one host species to another has been suggested in

74

molecular epidemiological and strain characterization studies, based on very close

75

genetic relatedness between human and porcine or human and equine PBV strains

76

(Bányai et al., 2008; Ganesh et al., 2011; Giordano et al., 2011). Given that PBV is

77

frequently found in waste water and surface water, a water-born route of

78

transmission has been also suggested (Ganesh et al., 2011). Indeed, the genetic

79

similarity among heterologous PBV strains has been seen along a short, ~200 bp

3

80

gene fragment of the S segment. Because the PBV genome is segmented and

81

measures ~ 4 kbp, the value of this short fragment to describe molecular

82

epidemiology may be limited. A major obstacle to perform whole genome based

83

studies is the lack of genome sequence data for a reasonably high number of strains.

84

Thus far, the complete genome sequence has been determined only for a human

85

strain, and more recently, for a phocine PBV strain (Wakuda et al., 2005; Woo et al.,

86

2012). In addition, complete or partial L segment sequences have become available

87

for a lapine strain and some human strains (Green et al., 1999; Rosen et al., 2000),

88

while the sequences of a dozen or so S gene fragments longer than 1000 bp has

89

been determined for human, porcine, simian, bovine, and murine PBVs (Ghosh et al.,

90

2009; Phan et al., 2011; Rosen et al., 2000; van Leeuwen et al., 2010; Wang et al.,

91

2012; and unpublished GenBank entries). In this study, sequencing and molecular

92

characterization of the full-length genome of a porcine PBV strain was performed by

93

sequence independent amplification and high throughput sequencing of the genomic

94

dsRNA.

95

During a rotavirus surveillance project in 2004, total RNAs prepared from 37

96

randomly selected fecal samples collected from diarrheic piglets were analyzed by

97

polyacrylamide gel electrophoresis and silver staining to diagnose rotavirus

98

infections. One sample contained bands consistent with picobirnavirus dsRNA (Fig.

99

1a). This porcine PBV strain, designated 221/04-16 was detected from a 45-day-old

100

piglet with diarrhea at a farm in Brescia, Northern Italy, in 2004.

101

Total RNA was extracted from fecal specimen using the commercial TRI-REAGENT–

102

LS (Molecular Research Centre), according to the manufacturer’s protocol. Single

103

stranded RNA (ssRNA) was removed by precipitation with 2 M LiCl (Sigma) at 4 C for

104

16 h followed by centrifugation at 16 000 g for 30 min. dsRNA was purified from the

105

resulting supernatant using a MinElute gel extraction kit (Qiagen). The complete

106

nucleotide sequence was determined for the small (S) and large (L) segments of this

107

strain by sequence-independent amplification and deep sequencing as described

108

previously (Potgieter et al., 2002, 2009; Lambden et al., 1992). Briefly, an anchor

109

primer was ligated to dsRNA and used for reverse transcription and subsequent

110

amplification of virus genome segments. Amplified cDNA products were then

4

111

subjected to sequencing with GS FLX technology at Inqaba biotechTM (Pretoria, South

112

Africa).

113

Lasergene7 software from DNASTAR was used for de novo assembly of the contigs.

114

Files containing the sequence information, quality values and flowgrams (sff files)

115

were loaded into the Seqman 7 programme of the Lasergene software. Contigs

116

resulting from the assembly were checked manually and their consensus sequences

117

were exported as FASTA files. Consensus sequences were aligned to known

118

sequences using MEGALIGN (Lasergene7). Finally, sequences were subjected to

119

BLASTN analysis using the National Center for Biotechnology Information website.

120

The

121

(http://www.ncbi.nlm.nih.gov/gorf/gorf.html). Nucleotide and amino acid sequence

122

alignments were prepared by Multalin (Corpet, 1988) and manually adjusted to

123

codons. Phylogenetic analysis was performed by using MEGA5 (Tamura et al.,

124

2011). Sequences were deposited in GenBank under the following accession

125

numbers: KF861767-KF861773.

126

Small genome segment. The S segment of strain 221/04-16 was 1730 bp in length.

127

In comparison with other full-length PBV S segments, this segment is shorter than

128

those of the human strain Hy005102 (1745 bp), the monkey strain CHN-49/2002

129

(1784 bp), and the bovine strain RUBV-P/IND/20 (1758 bp) and longer than those of

130

the human strains 1-CHN-97 (1696 bp) and GPBV6C1 (1711 bp), and the otarine

131

PBV (1687 bp). The length of the 5’ and 3’ UTRs were 74 and 49 bp, respectively,

132

and were structurally heterogenous when compared with reference strains, except for

133

the common GTAAA and ACTGC pentanucleotides, located at the termini of the RNA

134

segments and conserved among the various PBV strains. The S segment of strain

135

221/04-16 contained a single 1607 nt-long ORF (nt 75 to 1682). Alternative ORFs

136

were also predicted (data not shown). However, the predicted proteins were too short

137

and their presence was not consistently found in the S segments of other strains;

138

therefore these putative ORFs were ignored in this analysis. The protein coded by

139

the S segment ORF was 535 aa long and was predicted to be the viral RNA

140

dependent RNA polymerase (RdRp). It had conservative motifs in aa positions 266-

141

271, 328-333 and 364-366 defining the GDD sequence motif. When comparing the

142

221/04-16 S gene sequence with the Chinese strain SD, a single nucleotide deletion

NCBI’s

ORF

finder

was

used

to

identify

protein

coding

regions

5

143

was observed at position 768 that was compensated at position 848 by a single

144

nucleotide insertion. This insertion caused a codon shift, altering the translation of a

145

28 aa stretch between the two porcine PBVs (Fig. 1c).

146

Phylogenetic analysis was first performed using a ~200 bp fragment of the RdRp

147

used to amplify GGI PBV strains in diagnostic PCR (Rosen et al., 2000). This

148

analysis identified a Chinese porcine PBV strain as presenting the greatest nt

149

sequence identity (95%) with that of the Italian strain. Other relatively closely related

150

sequences (up to 89%) included viral sequences identified from US water samples.

151

The Italian porcine PBV strain shared lower nt similarity to European and American

152

(up to 75%) porcine PBVs and shared comparable similarities with European, Asian,

153

and American human PBVs (up to 88%). Another, ~280 bp long, partially overlapping

154

region of the S segment used to specifically amplify porcine PBVs (Carruyo et al.,

155

2008) was analyzed separately. This analysis demonstrated that the Italian strain

156

shared somewhat greater similarities with Argentinean porcine and human PBV

157

strains (up to 97% nt identity; data not shown). An analysis of the whole ORF of

158

several strains of porcine, human, murine, phocine, and bovine origin identified

159

comparable sequence identity data when these strains were compared along the

160

commonly characterized ~200 bp long fragment. The same observation was made

161

when comparing the Italian strain with two genogroup II strains, 4-GA-91 and

162

GPBV6G2. However, it should be considered that extrapolating nucleotide similarities

163

obtained from a short fragment may not be relevant in each case as exemplified by

164

the codon shift described above. Phylogenetic analysis of an ~1200 bp fragment of

165

selected strains with longer S genome segment sequence confirmed the similarity

166

based prediction of the genetic relatedness between the Italian strain and the

167

Chinese SD strain (Fig. 1b).

168

Large genome segment. Amplification and cloning of the L segment yielded a

169

heterogeneous population of cDNA products, comprising two major types and

170

several minor sequence variants within each type. The major sequence types were

171

designated LA and LB. The LB variants (ie LB-1, LB-3, and LB-4) were 2524 bp in

172

length and had two predicted ORFs at nt 195-794 (ORF1) and 810-2447 (ORF2),

173

encoding proteins of 199 and 545 aa in length, respectively. The LA variants, LA-2,

174

LA-5 and LA-7, respectively, were 2361, 2666 and 2351 bp long. The ORF1 started

6

175

at nt 185 and ended at nt 721, coding for a polypeptide of 175 aa in length. The

176

ORF2 started at nt 734 in all LA sequences. However, while LA-5 was terminated

177

with a stop codon, LA-2 and LA-7 seemed to be pseudogenes without translation

178

termination signals. The protein encoded by the LA-5 sequence was longer (619 aa)

179

than those of other PBV capsid proteins (ie, 590 aa in the lapine PBV strain, 552 aa

180

in the human strain Hy005102 and 575 aa in the otarine strain). The 3’ UTRs ranged

181

from 184 (LA segments) to 194 nt (LB segments). The 5’ UTRs ranged from 76 (LB

182

segments) to 84 nt (LA-5).

183

The ORF1-encoded proteins of the LA and LB clones were extremely diverse, yet

184

retained common genetic signatures. One shared feature was the recently

185

recognized 10-aa long sequence motif (Da Costa et al., 2011), ExxRxNxxxE, where

186

x, can be any amino acid. Four and 6 repeats of this motif was identified in LB clones

187

(e.g. LB-3) and LA clones (e.g. LA-5), respectively. The amino acid distance between

188

two adjacent motifs were variable. When comparing the ORF1s with other reference

189

strains, we found similar diversities in the number and position of the ExxRxNxxxE

190

motif (Fig 1d).

191

Due to the low number of sequence data derived from cognate genes, a meaningful

192

phylogenetic analysis of the capsid protein coding genes could not be performed.

193

However, as the three dimensional structure of a rabbit PBV has been determined by

194

using recombinant capsid protein (Duquerroy et al., 2009), attempts were made to

195

gain insight into the structure of the capsid protein of porcine PBVs by homology

196

modelling. The deduced aa sequences of the LB-3 and LA-7 clones were subjected

197

to protein modeling employing the Prime module of Schrödinger Suite (Schrödinger,

198

LLC, New York, NY, 2013) (www.schrodinger.com) molecular modelling software

199

package using the rabbit picobirnavirus capsid protein X-ray structure (PDB ID code:

200

2VF1) as homo-dimer template. Sequence similarity values of LB-3 and LA-7 with the

201

rabbit capsid protein were 38.3 % and 39.6 %, respectively. The models were refined

202

with the MacroModel energy minimization module of the Schrödinger Suite to

203

eliminate the steric conflicts between the side-chain atoms. The reconstructed

204

porcine PBV virion-like-particles (VLPs) were generated using the Oligomer

205

Generator

206

Molecular graphics and sequence alignment visualization were prepared using VMD

service

of

VIPERdb

(http://viperdb.scripps.edu/oligomer_multi.php).

7

207

version 1.9.1 (Humphrey et al., 1996) and the Multiple Sequence Viewer of the

208

Schrödinger Suite (Schrödinger, LLC, New York, NY, 2013), respectively. Structural

209

differences between the rabbit and the porcine PBV capsid protein (CP) models are

210

visualized in Fig. 2. The homology model for LB-3 proved to be very similar to the

211

rabbit PBV capsid protein (Fig. 2, panels a, c and d). Interestingly, we identified four

212

additional external surface exposed loop regions on the LA-7 pig PBV CP model (Fig.

213

2, panels b, c, d and g). These additional surface loops are located on the most

214

protruded edge of the coat protein, raising the possibility that these peptide insertions

215

may act as potential immune epitopes for B-cell mediated response.

216

The porcine PBV genome described represents the third complete sequence of this

217

highly diverse virus family. Until now only one human and one phocine PBV genome

218

sequence had been determined (Wakuda et al., 2005; Woo et al., 2012). In this study

219

we used the sequence independent amplification of genomic dsRNA and the next

220

generation sequencing strategy. This seemed important to maximize the depth of

221

sequence information and minimize any bias inherent to sequence variation seen in

222

PBV genomes, given that porcine PBVs frequently occur as mixed strain population

223

in the same host and there is evidence indicative of the quasi species nature of PBVs

224

(Bányai et al., 2008).

225

The porcine stool specimen from which strain 221/04-16 was isolated, contained two

226

major PBV subpopulations, each sharing the same RdRp gene and expressing

227

different capsid genes. The conservation of the RdRp gene sequence suggested that

228

one strain may be a reassortant. The likelihood that another RdRp gene was also

229

present but remained hidden in the sample is low, given that sequence independent

230

amplification coupled with deep sequencing were employed which, on one hand

231

shows no bias in recognition of the template genomic dsRNA segment and on the

232

other hand should detect minor sequence variants. This is further illustrated with the

233

truncated L segments of type LA clones. In this study we did not determine whether

234

pseudogenes among the LA sequences were generated as an artifact during RNA

235

processing or the shorter L segments encoding pseudogenes were incorporated in

236

the capsid. However, PBV genes with altered structure (pseudogenes) may be

237

tolerated/maintained if functionally integer cognate genes are retained (Bányai et al.,

8

238

2008) and we were able to determine at least one LA segment variant (LA-5) with a

239

correct ORF architecture in the porcine PBV genome.

240

Sequencing of the whole genome of strain 221/04-16 uncovered another key

241

mechanism likely contributing to the great sequence diversity seen even among PBV

242

strains collected from the same host species (Bányai et al., 2003, 2008; Carruyo et

243

al., 2008; Fregolente et al., 2009; Phan et al., 2011; Rosen et al., 2000; van Leeuwen

244

et al., 2010; Wang et al., 2012). We found further proof that insertions and deletions

245

(in/del) occur in the PBV genome, as evidenced by sequence length heterogeneity

246

seen in the UTRs and the coding regions of both S and L segments. The effect of

247

such mutations on phylogenetic analysis and protein evolution was demonstrated by

248

the observation of a deleterious mutation in the RdRp gene, which was compensated

249

by a single nucleotide insertion several nucleotides downstream (Fig. 1c). This

250

mutation resulted in a highly diverse deduced amino acid sequence in the flanking

251

region. It is possible that, in addition to the accumulation of single point mutations,

252

these compensatory in/del mutations are also common in the PBV genome,

253

generating variants of viral proteins with modified antigenic structure, more efficient

254

replication and/or increased viral fitness (Ganesh et al., 2012).

255

In conclusion, this study provides a more detailed insight into the evolutionary

256

mechanisms of PBVs. How different mechanisms contribute to better adaptation to a

257

homologous host or allow interspecies transmission should be the subject of future

258

studies. Until adequate cell culture and model animals are not established, whole

259

genome sequencing of PBVs from various host species remains the only alternative

260

to investigate this intriguing question.

261

Acknowledgements

262

K. B. was supported by the Momentum program (Hungarian Academy of Sciences).

263

Á. G. was the recipient of a János Bolyai fellowship from the Hungarian Academy of

264

Sciences. The licensing of the Schrödinger Suite software package was funded from

265

the OTKA grant under agreement No.:108793.

266 267

References

9

268

Bányai, K., Jakab, F., Reuter, G., Bene, J., Uj, M., Melegh, B. & Szucs, G. (2003).

269

Sequence

heterogeneity

among

human

picobirnaviruses

270

gastroenteritis outbreak. Arch Virol 148, 2281–2291.

detected

in

a

271

Bányai, K., Martella, V., Bogdán, Á., Forgách, P., Jakab, F., Meleg, E., Bíró, H.,

272

Melegh, B., & Szucs, G. (2008). Genogroup I picobirnaviruses in pigs: evidence

273

for genetic diversity and relatedness to human strains. J Gen Virol 89, 534–539.

274

Carruyo, G. M., Mateu, G., Martínez, L. C., Pujol, F. H., Nates, S. V., Liprandi, F.,

275

& Ludert, J. E. (2008). Molecular characterization of porcine picobirnaviruses and

276

development of a specific reverse transcription-PCR assay. J Clin Microbiol 46,

277

2402–2405.

278 279 280 281

Corpet, F. (1988). Multiple sequence alignment with hierarchical clustering. Nucleic Acids Res 16, 10881–10890. Da Costa, B., Duquerroy, S., Tarus, B., & Delmas, B. (2011). Picobirnaviruses encode a protein with repeats of the ExxRxNxxxE motif. Virus Res 158, 251–256.

282

Delmas, B. (2011). Picobirnaviridae In Virus Taxonomy. Ninth Report of the

283

International Committee on Taxonomy of Viruses, pp. 535-539. Edited by King, A.

284

M. Q., Adams, M. J., Carstens, E. B., Lefkowitz, E. J. Elsevier Inc.

285

Duquerroy, S., Da Costa, B., Henry, C., Vigouroux, A., Libersou, S., Lepault, J.,

286

Navaza, J., Delmas, B., & Rey, F. A. (2009). The picobirnavirus crystal structure

287

provides functional insights into virion assembly and cell entry. EMBO J 28,

288

1655–1665.

289

Fregolente, M. C., de Castro-Dias, E., Martins, S. S., Spilki, F. R., Allegretti, S.

290

M., & Gatti, M. S. (2009). Molecular characterization of picobirnaviruses from new

291

hosts. Virus Res 143, 134–136.

292

Gallimore, C. I., Appleton, H., Lewis, D., Green, J., & Brown, D. W. (1995).

293

Detection and characterisation of bisegmented double-stranded RNA viruses

294

(picobirnaviruses) in human faecal specimens. J Med Virol 45, 135–140.

295

Ganesh, B., Banyai, K., Masachessi, G., Mladenova, Z., Nagashima, S., Ghosh,

296

S., Nataraju, S. M., Pativada, M., Kumar, R., & Kobayashi, N. (2011).

297

Genogroup I picobirnavirus in diarrhoeic foals: can the horse serve as a natural

298

reservoir for human infection? Vet Res 42, 52.

299

Ganesh, B., Bányai, K., Martella, V., Jakab, F., Masachessi, G., & Kobayashi, N.

300

(2012). Picobirnavirus infections: viral persistence and zoonotic potential. Rev

301

Med Virol 22, 245–256. 10

302

Ghosh, S., Kobayashi, N., Nagashima, S., & Naik, T. N. (2009). Molecular

303

characterization of full-length genomic segment 2 of a bovine picobirnavirus

304

(PBV) strain: evidence for high genetic diversity with genogroup I PBVs. J Gen

305

Virol 90, 2519–2524.

306 307

Gillman, L., Sánchez, A. M., & Arbiza, J. (2013). Picobirnavirus in captive animals from Uruguay: identification of new hosts. Intervirology 56, 46-49.

308

Giordano, M. O., Martinez, L. C., Rinaldi, D., Gúinard, S., Naretto, E., Casero, R.,

309

Yacci, M. R., Depetris, A. R., Medeot, S. I., & Nates, S. V. (1998). Detection of

310

picobirnavirus in HIV-infected patients with diarrhea in Argentina. J Acquir

311

Immune Defic Syndr Hum Retrovirol 18, 380–383.

312

Giordano, M. O., Martinez, L. C., Masachessi, G., Barril, P. A., Ferreyra, L. J.,

313

Isa, M. B., Valle, M. C., Massari, P. U., & Nates, S. V. (2011). Evidence of

314

closely related picobirnavirus strains circulating in humans and pigs in Argentina.

315

J Infect 62, 45–51.

316

Green, J., Gallimore, C. I., Clewley, J. P., & Brown, D. W. (1999). Genomic

317

characterisation of the large segment of a rabbit picobirnavirus and comparison

318

with the atypical picobirnavirus of Cryptosporidium parvum. Arch Virol 144, 2457–

319

2465.

320

Grohmann, G. S., Glass, R. I., Pereira, H. G., Monroe, S. S., Hightower, A. W.,

321

Weber, R., & Bryan, R. T. (1993). Enteric viruses and diarrhea in HIV-infected

322

patients. Enteric Opportunistic Infections Working Group. N Engl J Med 329, 14–

323

20.

324

Haga, I. R., Martins, S. S., Hosomi, S. T., Vicentini, F., Tanaka, H., & Gatti, M. S.

325

(1999).

326

(Picobirnavirus) in faeces of giant anteaters (Myrmecophaga tridactyla). Vet J

327

158, 234–236.

328 329

Identification

of

a

bisegmented

double-stranded

RNA

virus

Humphrey, W., Dalke, A., & Schulten, K. (1996). VMD - Visual Molecular Dynamics. J Mol Graphics 14, 33–38.

330

Lambden, P. R., Cooke, S. J., Caul, E. O., & Clarke, I. N. (1992). Cloning of

331

noncultivatable human rotavirus by single primer amplification. J Virol 66, 1817–

332

1822.

333

Malik, Y. S., Chandrashekar, K. M., Sharma, K., Haq, A. A., Vaid, N., Chakravarti,

334

S., Batra, M., Singh, R., & Pandey, A. B. (2011). Picobirnavirus detection in

11

335

bovine and buffalo calves from foothills of Himalaya and Central India. Trop Anim

336

Health Prod 43, 1475–1478.

337

Masachessi, G., Martínez, L. C., Giordano, M. O., Barril, P. A., Isa, B. M.,

338

Ferreyra, L., Villareal, D., Carello, M., Asis, C., & Nates, S. V. (2007).

339

Picobirnavirus (PBV) natural hosts in captivity and virus excretion pattern in

340

infected animals. Arch Virol 152, 989–998.

341

Masachessi, G., Martinez, L. C., Ganesh, B., Giordano, M. O., Barril, P. A., Isa,

342

M. B., Ibars, A., Pavan, J. V., & Nates, S. V. (2012). Establishment and

343

maintenance of persistent infection by picobirnavirus in greater rhea (Rhea

344

Americana). Arch Virol 157, 2075–2082.

345

Pereira, H. G., Flewett, T. H., Candeias, J. A., & Barth, O. M. (1988). A virus with a

346

bisegmented double-stranded RNA genome in rat (Oryzomys nigripes) intestines.

347

J Gen Virol 69, 2749–2754.

348 349

Phan, T. G., Kapusinszky, B., Wang, C., Rose, R.K., Lipton, H.L., & Delwart, E.L. (2011). The fecal viral flora of wild rodents. PLoS Pathog 7, e1002218.

350

Potgieter, A. C., Steele, A. D., & van Dijk, A. A. (2002). Cloning of complete

351

genome sets of six dsRNA viruses using an improved cloning method for large

352

dsRNA genes. J Gen Virol 83, 2215–2223.

353

Potgieter, A. C., Page, N. A., Liebenberg, J., Wright, I. M., Landt, O., & van Dijk,

354

A. A. (2009). Improved strategies for sequence-independent amplification and

355

sequencing of viral double-stranded RNA genomes. J Gen Virol 90, 1423–1432.

356

Rosen, B. I., Fang, Z. Y., Glass, R. I. & Monroe, S. S. (2000). Cloning of human

357

picobirnavirus genomic segments and development of an RT-PCR detection

358

assay. Virology 277, 316–329.

359

Tamura, K., Peterson, D., Peterson, N., Stecher, G., Nei, M., & Kumar, S. (2011).

360

MEGA5: molecular evolutionary genetics analysis using maximum likelihood,

361

evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28,

362

2731–2739.

363

Van Leeuwen, M., Williams, M. M., Koraka, P., Simon, J. H., Smits, S. L., &

364

Osterhaus, A. D. (2010). Human picobirnaviruses identified by molecular

365

screening of diarrhea samples. J Clin Microbiol 48, 1787–1794.

366

Wakuda, M., Pongsuwanna, Y., & Taniguchi, K. (2005). Complete nucleotide

367

sequences of two RNA segments of human picobirnavirus. J Virol Methods 126,

368

165–169. 12

369

Wang, Y., Tu, X., Humphrey, C., McClure, H., Jiang, X., Qin, C., Glass, R. I., &

370

Jiang, B. (2007). Detection of viral agents in fecal specimens of monkeys with

371

diarrhea. J Med Primatol 36, 101–107.

372

Wang, Y., Banyai, K., Tu, X., & Jiang, B. (2012). Simian genogroup I

373

picobirnaviruses: prevalence, genetic diversity, and zoonotic potential. J Clin

374

Microbiol 50, 2779–2782.

375

Woo, P. C., Lau, S. K., Bai, R., Teng, J. L., Lee, P., Martelli, P., Hui, S. W., &

376

Yuen, K. Y. (2012). Complete genome sequence of a novel picobirnavirus,

377

otarine picobirnavirus, discovered in California sea lions. J Virol 86, 6377–6378.

378 379 380

Figure legends

381 382

Fig. 1. (a) RNA profile of the PBV strain 221/04-16 identified in an Italian pig with

383

diarrhea. The genomic pattern of a Rotavirus A strain electrophoresed in parallel with

384

the PBV strain is also shown. (b) Nucleotide (NT) and amino acid (AA) sequence

385

based phylogeny of the partial S genome segment. The gene fragment used in the

386

phylogeny encodes a ~ 400 aa long polypeptide (position nt 290-1414, reference

387

strain Hy005102). Bootstrap values ≥50 are indicated. GGI and GGII refer to the

388

genogroup assignments of PBV strains. The scale bar (5 nt changes per 100

389

positions) is proportional to the genetic distance. (c) Partial S genome segment of

390

two porcine PBV strains, the Italian 221/04-16 and the Chinese SD. The alignment

391

shows the relative position of the single nucleotide insertion and deletion resulting in

392

remarkable amino acid change within the affected region. Numbers above the

393

alignment refer to the nucleotide positions of 221/04-16. (d) Number and position of

394

the sequence motif ExxRxNxxxE in the ORF1-encoded protein (L segment) of

395

various PBV strains (x can be any amino acid; Hu, human, La, lapine, Ot, otarine, Po,

396

porcine; Hu-1 is VS10, Hu-2 is Hy005102, Po-1 is LA-5, Po-2 is LB-3). Numbers on

397

the right indicate the length of the protein encoded by ORF1, ranging from 106

398

(lapine PBV) to 224 (Hy005102) aa in length. The protein sequences are not aligned.

399 400

Fig. 2. Structure comparison of PBV capsid (ORF2, large segment) proteins and

401

virions. Structure based alignment of the newly described LB-3 and LA-7 pig

402

picobirnavirus mature capsid protein amino acid sequences (a) and (b). The 13

403

background of the sequence alignments reflects the homology levels of the two

404

related PBV capsid protein sequences: identical amino acids are red, similar amino

405

acids are light orange while different amino acids are light pink. The four additional

406

external surface exposed loop regions on LA-7 PBV capsid protein are indicated

407

consequently on panel (b, c, d and g): loop region from aa 262 to 268 is orange, loop

408

region form aa 405 to 411 is green, loop region from aa 454 to 466 is red, while loop

409

region from aa 558 to 563 is purple. The superimposed LB-3 (pink) and LA-7 (cyan)

410

porcine PBV capsid protein and the template rabbit (grey) PBV capsid protein dimers

411

are in cartoon representation (c) and (d). The extra four outer loops on LA-7 PBV

412

VLP are illustrated in licorice representation. Molecular surface representation of the

413

rabbit (e), and the porcine LB-3 (f) and LA-7 PBV (g) virions. The molecular surface

414

is colored by radial extension of the amino acids from the virion centre. Dark blue

415

represents the most protruding capsid protein parts.

416

14

Fig. 1. (b)

(a)

768 |

(c)

848 |

221/04-16 SD

221/04-16 SD

(d)

Fig. 2.

Genome sequencing identifies genetic and antigenic divergence of porcine picobirnaviruses.

The full-length genome sequence of a porcine picobirnavirus (PBV) detected in Italy in 2004 was determined. The smaller (S) genome segment was 1730 nt...
4MB Sizes 0 Downloads 3 Views