JVI Accepted Manuscript Posted Online 21 January 2015 J. Virol. doi:10.1128/JVI.03328-14 Copyright © 2015, American Society for Microbiology. All Rights Reserved.

1

1 2

Identification of PB2 mutations responsible for the efficient replication

3

of H5N1 influenza viruses in human lung epithelial cells

4 5 6 7

Reina Yamaji1, Shinya Yamada1##, Mai Q. Le2, Chengjun Li3, Hualan Chen3, Ema Qurnianingsih4, Chairul A. Nidom4, Mutsumi Ito1, Yuko Sakai-Tagawa1, Yoshihiro Kawaoka1, 5, 6 #

8 9 10

1

Division of Virology, Department of Microbiology and Immunology, Institute of Medical Science, University of Tokyo, Tokyo 108-8639, Japan

11 12

2

National Institute of Hygiene and Epidemiology, Hanoi, Vietnam

13 14 15 16

3

State Key Laboratory of Veterinary Biotechnology, Harbin Veterinary Research Institute, Chinese Academy of Agricultural Sciences, 427 Maduan Street, Nangang District, Harbin, China 150001.

17 18

4

Avian Influenza Research Center, Airlangga University, Surabaya, Indonesia

19 20

5

21

Wisconsin-Madison, WI 53711

Department of Pathobiological Sciences, School of Veterinary Medicine, University of

22 23 24

6

Department of Special Pathogens, International Research Center for Infectious Diseases, Institute of Medical Science, University of Tokyo, Tokyo 108-8639, Japan

25 26

Running title: Adaptive mutations of H5N1 virus

27

Document totals: Word count for abstract: 190 Word count for main text: 3622

28 29 30

#

31

##

32

Address for correspondence: Division of Virology, Department of Microbiology and Immunology, Institute of Medical Science, University of Tokyo, Tokyo 108-8639, Japan. E-mail: [email protected]

33 34

Corresponding author. Tel.: +81-3-5449-5504; fax: +81-3-5449-5408 Co-corresponding author. Tel.: +81-3-5449-5504; fax: +81-3-5449-5408

2 35 36 37 38

Address for co-correspondence: Division of Virology, Department of Microbiology and Immunology, Institute of Medical Science, University of Tokyo, Tokyo 108-8639, Japan. E-mail: [email protected]

3

39

Abstract

40

Highly pathogenic H5N1 avian influenza viruses have caused outbreaks among poultry

41

worldwide, resulting in sporadic infections in humans, approximately 60% mortality of which

42

have been fatal. However, efficient transmission of H5N1 viruses among humans has yet to

43

occur, suggesting that further adaptation of H5N1 viruses to humans is required for their

44

efficient transmission among humans. Viral determinants for efficient replication in humans

45

are currently poorly understood. Here, we report that the polymerase PB2 protein of an

46

H5N1 influenza virus isolated from a human in Vietnam (A/Vietnam/UT36285/2010, 36285)

47

increased the growth ability of an avian H5N1 virus (A/wild bird/Anhui/82/2005, Wb/AH82) in

48

human lung epithelial A549 cells (although, the reassortant virus did not replicate more

49

efficiently than human 36285 virus). Furthermore, we demonstrate that amino acid residues

50

at positions 249, 309, and 339 of the PB2 protein from this human isolate were responsible

51

for its efficient replication in A549 cells. PB2-249G and 339M, which are found in the human

52

H5N1 virus, are rare in H5N1 viruses from both human and avian sources. Interestingly,

53

PB2-249G is found in over 30% of human seasonal H3N2 viruses, which suggests that

54

H5N1 viruses may replicate well in human cells when they acquire this mutation. Our data

55

are of value to H5N1 virus surveillance.

56

4 57

Importance

58

Highly pathogenic H5N1 avian influenza viruses must acquire mutations to overcome the

59

species barrier between avian species and humans. When H5N1 viruses replicate in human

60

respiratory cells, they can acquire amino acid mutations that allow them to adapt to humans

61

through continuous selective pressure. Several amino acid mutations have been shown to

62

be advantageous for virus adaptation to mammalian hosts. Here, we found that amino acid

63

changes at positions 249, 309, and 339 of PB2 contribute to efficient replication of avian

64

H5N1 viruses in human lung cells. These findings are beneficial for evaluating the pandemic

65

risk of circulating avian viruses and for further functional analysis of PB2.

66 67

5 68

Introduction

69

Highly pathogenic H5N1 influenza A viruses continue to circulate among avian species. As a

70

result, sporadic avian-to-human transmission has occurred and the threat of a pandemic has

71

persisted. The first human case caused by a highly pathogenic H5N1 influenza A virus was

72

reported in Hong Kong in 1997 (1, 2). Although the mass culling ordered by the Hong Kong

73

government temporarily ended the outbreak, a second outbreak of H5N1 viruses occurred in

74

2003. With the spread of H5N1 viruses among migrating birds and poultry, these viruses

75

quickly spread beyond East Asia (3). This spread of H5N1 viruses has been accompanied

76

by increasing reports of cases of avian-to-human transmission. As of January 2015, the total

77

number of confirmed human cases of highly pathogenic H5N1 influenza A virus infection had

78

reached 694, with 402 deaths ( http://www.who.int/, 4, 5). However, the human cases of

79

H5N1 infection have been limited mainly to individuals in close contact with infected poultry,

80

and reports of human-to-human transmission have been extremely rare (6, 7). For an H5N1

81

virus to overcome the host barrier, the following conditions must be met: efficient

82

transmission via droplets, efficient replication in the host, and immune vulnerability of the

83

infected populace. Consequently, viruses harboring mutations that promote efficient

84

transmission and replication in humans could lead to a pandemic.

85

A number of viral determinants have been identified that contribute to adaptation of avian

86

influenza viruses to mammalian hosts (8-28), and several large-scale comparative analyses

87

of proteins from avian and human viruses have been performed to catalog the amino acids

6

88

that are conserved in each host species (29-31). The PB2, PB1, and PA subunits of the

89

polymerase complex play a major role in virus pathogenicity and efficient viral growth in

90

mammals. PB2 is particularly important in overcoming species barriers. For example, lysine

91

(K) at residue 627 of PB2 enhances the polymerase activity, replication efficiency, and

92

virulence of H5N1 viruses in mammals (8-10). In addition, arginine at residue 591, which is

93

located close to the amino acid at position 627 of PB2 in the three-dimensional structure of

94

the protein, compensates for the lack of lysine at residue 627 and confers efficient viral

95

replication to pandemic H1N1 viruses in mammals (11). Other amino acid substitutions, such

96

as 701N, 271A, and 158G, in PB2 have also been identified as important markers that

97

confer a replicative advantage and high pathogenicity to influenza viruses in mammals

98

(12-15). Recently, we showed that the combination of amino acid substitutions 147T, 339T,

99

and 588T in PB2 confer a high polymerase activity to avian viruses in human cells (28). It is

100

not clear, however, whether we have identified all of the possible amino acid changes that

101

are needed for efficient growth of H5N1 viruses in mammals.

102

In this study, we attempted to identify amino acids that support the efficient growth of H5N1

103

viruses in human lung cells. Accumulation of information on adaptive mutations to humans

104

helps us to prepare for an unpredictable but potential H5N1 pandemic and to assess the risk

105

of future isolates with pandemic potential. Here, we attempted to identify new molecular

106

determinants associated with the efficient growth of a human-isolated H5N1 virus in human

7

107

lung cells by generating reassortant and mutant viruses in the genetic background of an

108

avian H5N1 virus (with low replicative ability in human lung cells) and a human H5N1 virus

109

(with high replicative ability in human lung cells).

110

8 111

Materials and Methods

112

Cells

113

Madin-Darby canine kidney (MDCK) cells were grown in minimal essential medium (MEM)

114

containing 5% newborn calf serum, vitamins, essential amino acids, and antibiotics. Human

115

embryonic kidney 293T cells, A549 human lung adenocarcinoma epithelial cells, and DF-1

116

chicken embryo fibroblast cells were grown in Dulbecco modified Eagle medium (DMEM)

117

containing 10% fetal bovine serum (FBS) and antibiotics. MDCK, 293T, and A549 cells were

118

cultured at 37℃ with 5% CO2. DF-1 cells were cultured at 39℃ with 5% CO2 unless

119

otherwise stated.

120 121

Viruses

122

In this study, we used the following ten H5N1 viruses: A/wild bird/Anhui/82/2005 (Wb/AH82)

123

(accession numbers: PB2 KP638500, PB1 KP638559, PA KP638509, HA KP638516, NP

124

KP638524, NA KP638533, M KP638542, NS KP638551), A/chicken/Vietnam/TY31/2005

125

(Ck/TY31) (accession numbers: PB2 KP638493, PB1 KP638558, PA KP638501, HA

126

EU118136.1, NP KP638525, NA EU118127.1, M KP638534, NS KP638550), A/chicken/Central

127

Java/UT3091/2005

128

GQ122415, and GQ122395),

129

PB2 KP638492, PB1 KP638552, PA KP638507, HA KP638560, NP KP638523, NA

130

KP638530, M KP638539, NS KP638547), A/Vietnam/UT36282/2010 (36282) (accession

(Ck/UT3091)

(accession

numbers:

GQ122490

to

GQ122495,

A/Vietnam/UT36285/2010 (36285) (accession numbers:

9

131

numbers: PB2 KP638494, PB1 KP638557, PA KP638506, HA KP638515, NP KP638522,

132

NA KP638532, M KP638541, NS KP638549), A/Vietnam/UT36236/2010 (36236) (accession

133

numbers: PB2 KP638495, PB1 KP638556, PA KP638505, HA KP638514, NP KP638520,

134

NA KP638531, M KP638540, NS KP638548), A/Vietnam/HN31676DH/2009 (31676)

135

(accession numbers: PB2 KP638496, PB1 KP638553, PA KP638504, HA KP638513, NP

136

KP638519, NA KP638526, M KP638535, NS KP638543), A/Vietnam/UT31641 Ⅱ /2008

137

(31641) (accession numbers: PB2 KP638497, PB1 KP638554, PA KP638503, HA

138

KP638512,

139

A/Vietnam/UT31604Ⅰ/2009 (31604) (accession numbers: PB2 KP638498, PB1 KP638561,

140

PA KP638502, HA KP638511, NP KP638517, NA KP638527, M KP638536, NS KP638544),

141

and A/Vietnam/UT36250 Ⅰ /2010 (36250) (accession numbers: PB2 KP638499, PB1

142

KP638555, PA KP638508, HA KP638510, NP KP638521, NA KP638529, M KP638538, NS

143

KP638546). Throat swabs were collected from H5N1 influenza virus-infected patients in

144

northern Vietnam and were sent to the National Institute of Hygiene and Epidemiology in

145

Hanoi, Vietnam. To isolate H5N1 virus, clinical specimens were inoculated to MDCK cells in

146

MEM containing 0.3% BSA and incubated at 37℃ for 48 hours. Stock viruses were

147

propagated in MDCK cells at 37℃ and stored at -80℃. 36285 was isolated from a patient

148

who exhibited influenza symptoms and recovered (a two-year-old female, onset on April 2,

149

2010, hospitalized on April 4, 2010). All experiments with H5N1 viruses were performed in a

NP

KP638518,

NA

KP638528,

M

KP638537,

NS

KP638545),

10

150

biosafety level 3 containment laboratory approved for such use by the Ministry of Agriculture,

151

Forestry, and Fisheries, Japan.

152 153

Isolation of viral RNA, RT-PCR, and generation of viruses by use of reverse genetics.

154

Viral RNA was extracted from the supernatants of virus-infected MDCK cells by using a

155

QIAamp viral RNA minikit (Qiagen, Hilden, Germany). Extracted RNA was reverse

156

transcribed with Superscript Ⅲ (Invitrogen, Carlsbad, CA) and the universal primers specific

157

for influenza A virus genes to generate cDNA. The resulting products were PCR-amplified by

158

using KOD Fx DNA polymerase (Toyobo, Osaka, Japan) with specific primers for each virus

159

gene and cloned into the RNA polymerase I plasmid pHH21 (ref). Mutations in the PB2 gene

160

of Wb/AH82 were generated by PCR amplification of the respective PB2 construct with

161

primers possessing the desired mutations. Primer sequences are available upon request. All

162

constructs were sequenced to ensure the absence of unwanted mutations. All avian,

163

reassortant, and mutant viruses were generated by use of plasmid-based reverse genetics,

164

as described previously (32). The culture supernatant derived from transfected cells was

165

amplified in DF-1 cells grown in DMEM containing 0.3% bovine serum albumin. At 48 h

166

post-infection, the culture supernatant was harvested, clarified, divided into aliquots, and

167

stored at -80℃. The virus titers of all human and avian viruses were determined by using

168

plaque assays in MDCK cells.

11

169 170

Mouse experiments

171

Six-week-old female BALB/c mice (Japan SLC) were used for these experiments. To

172

determine the dose lethal to 50% of mice (MLD50), 5 mice/group were anesthetized with

173

isoflurane and inoculated intranasally with 101 to 105 plaque-forming units (PFU) in a 50-µl

174

volume. The mice were monitored daily for clinical signs of infection and checked for

175

changes in body weight and mortality for 14 days post-infection. MLD50 values were

176

calculated by using the method of Reed & Muench (1938). All experiments with mice were

177

performed in accordance with the University of Tokyo's Regulations for Animal Care and Use

178

and were approved by the Animal Experiment Committee of the Institute of Medical Science,

179

the University of Tokyo.

180 181

Viral replication assay

182

Triplicate wells of confluent A549 cells or DF-1 cells were infected with viruses at a

183

multiplicity of infection (MOI) of 0.0002, and incubated for 1 h at 37℃ or 39℃, respectively.

184

After the 1-h incubation, A549 cells were further incubated in MEM containing 0.3% bovine

185

serum albumin (BSA) at 33℃ and 37℃. DF-1 cells were further incubated in DMEM

186

containing 0.3% bovine serum albumin (BSA) at 39℃. Aliquots of supernatants were

187

harvested at 1, 24, 48, 72, and 96 h post-infection, and frozen at -80℃. Virus titers in the

12

188

culture supernatants at each time point were determined by plaque assays in MDCK cells.

189 190

Statistical analysis

191

Differences in the virus titers of the supernatants were statistically analyzed by using the

192

Student’s t-test. Differences in mean maximum body weight loss in mice were also

193

statistically analyzed by using the Student’s t- test.

194 195

13 196

Results

197

Comparison of the growth ability of H5N1 viruses in A549 and DF-1 cells

198

To identify human adaptive mutations in H5N1 influenza viruses, we first compared the

199

growth properties at a multiplicity of infection (MOI) of 0.0002 in carcinomic human alveolar

200

basal

201

A/Vietnam/UT36285/2010

202

A/Vietnam/UT36236/2010

203

A/Vietnam/UT31641 Ⅱ /2008

204

A/Vietnam/UT36250Ⅰ/2010 (36250). All seven human H5N1 viruses grew well in A549 cells

205

at 37˚C with maximum titers of over 106 PFU/ml, except for 31676 which grew to slightly

206

lower titers (Fig. 1A). Sequence analysis revealed that 36250, 36236, 31604, and 31641

207

possess lysine at position 627 of the PB2 protein (PB2-627K), whereas 36282 possesses

208

PB2-591R. However, viruses 31676 and 36285 do not have any known PB2 markers that

209

could account for the efficient replication in mammals (Table 1). These findings suggest that

210

the human H5N1 viruses 31676 and 36285 may have unreported amino acids that enhance

211

their viral growth ability in A549 cells. Therefore, in this study we focused on the replicative

212

efficiency of 36285. First, we generated the 36285 virus by use of reverse genetics

213

(36285-RG) and confirmed that 36285-RG grew as well as wild-type 36285 in A549 cells

214

(Fig.1B). We next compared the growth property of 36285-RG in A549 cells at 37˚C with that

215

of three avian H5N1 viruses generated by use of reverse genetics: A/wild bird/Anhui/82/2005

epithelial

A549

cells

of

seven

(36285), (36236), (31641),

H5N1

viruses

isolated

from

humans:

A/Vietnam/UT36282/2010

(36282),

A/Vietnam/HN31676DH/2009

(31676),

A/Vietnam/UT31604 Ⅰ /2009

(31604),

and

14

216

(Wb/AH82-RG), A/chicken/Vietnam/TY31/2005 (Ck/TY31-RG), and A/chicken/Central

217

Java/UT3091/2005 (Ck/UT3091-RG) (Fig. 1C). Ck/TY31-RG and Wb/AH82-RG grew poorly

218

with maximum titers of less than 104.5 PFU/ml (Fig. 1C). Of note, both 36285 and Wb/AH82

219

belong to the same H5 HA sub-clade 2.3.4 and are genetically closely related (Table 2).

220

Therefore, we compared 36285 and Wb/AH82 to elucidate the mechanism of H5N1

221

adaptation to humans. In chicken embryo fibroblast DF-1 cells, 36285-RG and Wb/AH82-RG

222

both grew well at 39˚C (Fig. 2A), demonstrating their potential to replicate in avian cells; in

223

contrast, 36285-RG grew much better than Wb/AH82-RG in A549 cells at 33˚C (Fig. 2B),

224

demonstrating its potential to replicate in mammalian cells. We compared the growth

225

capability of H5N1 viruses in A549 cells at both 37℃ and 33℃ because in humans, the

226

temperature in the lungs is 37℃, and that in the upper airway is 33℃. However, the body

227

temperature of birds is 39℃; therefore, we compared the growth capability of H5N1 viruses

228

in chicken DF-1 cells only at 39℃.

229 230

Comparison of the pathogenicity of Wb/AH82-RG and 36285-RG in mice

231

We next examined the pathogenicity of Wb/AH82-RG and 36285-RG in mice. Inoculation of

232

103, 104, and 105 PFU of 36285-RG virus resulted in 100% mortality, and a 40% fatality rate

233

was observed in the mice that received 101 and 102 PFU of virus. In contrast, inoculation of

234

104 and 105 PFU of Wb/AH82-RG resulted in 100% mortality and a 60% fatality rate was

15

235

observed in mice that received 102 and 103 PFU of virus. The MLD50 values for 36285-RG

236

and Wb/AH82-RG were 101.5 PFU and 102.7 PFU, respectively (Fig. 3C, 3D). Thus,

237

36285-RG and Wb/AH82-RG differed in their mouse virulence by only one log MLD50 unit.

238 239

The PB2 gene segment from 36285 is responsible for its efficient growth in human

240

cells.

241

To elucidate the molecular basis for the replicative difference between Wb/AH82 and 36285

242

in A549 cells, we generated a series of reassortant viruses (as illustrated in Fig. 4A) and

243

compared their growth properties in A549 cells. The reassortant viruses were named

244

according to the origin of their Wb/AH82 or 36285 genes. For instance, “36285(PB2)”

245

indicates a virus possessing PB2 from 36285 and the rest of its gene segments from

246

Wb/AH82. Three reassortant viruses possessing the PB2, PB1, PA, and NP of 36285 [i.e.,

247

36285(3P+NP), 36285(3P+NP, HA, NA), and 36285(3P+NP, M, NS)] were comparable in

248

their growth to 36285-RG; none of these viruses replicated more efficiently than 36285-RG

249

(Fig. 4B). The replicative ability of 36285(M, NS) was similar to that of Wb/AH82-RG,

250

indicating that the M and NS proteins of 36285 do not have a large impact on the difference

251

in the growth capabilities of Wb/AH82-RG and 36285-RG. Of note, the HA and NA of 36285

252

also contributed, to some extent, to the difference in the growth properties of Wb/AH82-RG

253

and 36285-RG; however, they enhanced viral replication to a much lesser extent than did

16

254

the polymerase subunits and NP of 36285. These results suggest that the PB2, PB1, PA,

255

and/or NP gene segments of 36285 were the most responsible for the difference in growth

256

capability between Wb/AH82-RG and 36285-RG in A549 cells.

257

To determine which viral segment(s) among PB2, PB1, PA, and NP contributed to the high

258

replication of 36285-RG, we generated a range of reassortants (Fig. 5A) and compared their

259

growth properties in A549 cells. Only viruses possessing 36285(PB2) grew well (Fig. 5B). By

260

contrast, 36285(PB1, PA, NP) grew poorly, similarly to Wb/AH82-RG. These results

261

demonstrate that the PB2 of 36285 makes an important contribution to the difference in

262

growth capability between Wb/AH82-RG and 36285-RG in A549 cells.

263 264

A single G-to-D mutation at 309 and the double mutations E249G and T339M enhance

265

the replication of Wb/AH82-RG.

266

Sequence comparison of the PB2 proteins of Wb/AH82 and 36285 revealed twenty-one

267

amino acid differences (Table 3). To identify the specific changes that give rise to efficient

268

viral replication, we generated mutant viruses possessing a chimeric PB2 protein (Fig. 6A);

269

the remaining gene segments were derived from Wb/AH82. In A549 cells, the mutant

270

viruses possessing chimera 1, 5, or 6 grew poorly, whereas those with chimera 2 or 3 grew

271

as well as that with 36285 PB2. These results suggest that residues 221 to 460 contribute

272

most to the difference between 36285-RG and Wb/AH82-RG in terms of growth in A549

17

273

cells.

274

To further define which amino acid residues enhance the viral replication, we generated

275

viruses with additional chimeric PB2 proteins (Fig.7A). The mutant viruses possessing

276

chimera 2, 7, 8, and 11 grew well, whereas the growth of the viruses with chimera 1, 9, and

277

10 was lower compared with that of the viruses possessing residues 221 to 345 from 36285.

278

These results indicate that residues 221 to 345 from 36285 provide the replicative

279

advantage to 36285-RG (Fig. 7B).

280

We then tried to determine which amino acids among residues 221 to 345 of PB2 enhanced

281

the growth properties of 36285-RG. There are three amino acid differences in this region

282

between Wb/AH82 and 36285: 249E/G, 309G/D, and 339T/M. We singly or in combination

283

introduced the three residues of 249G, 309D, and 339M into the PB2 of Wb/AH82 (Fig. 8A)

284

and generated viruses possessing the mutant Wb/AH82 PB2 gene in the background of the

285

remaining

286

changes found among circulating influenza viruses). Indeed, the combination of PB2-249G,

287

-309D, and -339M markedly increased the growth capability of Wb/AH82-RG by around 102

288

fold compared with Wb/AH82-RG (Fig. 8B, 8C). Although the single introduction of

289

PB2-249G, -PB2-339M, or -PB2-309D enhanced the growth capability of Wb/AH82-RG,

290

PB2-339M was least effective (Fig. 8B). All mutants that possessed combinations of two of

291

the PB2 mutations tested grew better than Wb/AH82-RG by more than 101.5-fold (Fig. 8C).

Wb/AH82 genes (note that all mutant Wb/AH82 viruses possess amino acid

18

292

These results indicate that PB2-249G, PB2-339M, and PB2-309D support efficient viral

293

growth in A549 cells.

294 295

The amino acid mutations E249G and T339M may be associated with the adaptation of

296

H5N1 virus to humans.

297

To further evaluate the potential role of the PB2 amino acid residues 249, 309, and 339 in

298

adaptation to humans, we collected full-length PB2 sequences of influenza viruses of

299

various subtypes from the Influenza sequence database (ISD) (Table 4). Almost all of H5N1

300

viruses contained PB2 amino acid 309D, which is highly conserved among influenza viruses

301

of various subtypes regardless of the host species, implying that this residue does not need

302

to change for a virus to break through the species barrier. In contrast, PB2 amino acids

303

249G and 339M were rare among the avian H5N1 viruses. Intriguingly, no less than 49.2%

304

of the human seasonal H3N2 viruses had PB2 amino acid 249G. This finding supports the

305

hypothesis that PB2 amino acid 249G is the amino acid mutation that the virus gains during

306

replication in human respiratory cells and increases its fitness for human adaptation.

307 308

19 309

Discussion

310

Several amino acid mutations associated with the adaptation of H5N1 virus to humans have

311

been identified, including 627K (8-10), 591R, 591K (11), 701N (9, 13), 271A (14), and 158G

312

(15) of PB2. We thought that further analysis of H5N1 viruses isolated from humans could

313

identify new mutations needed for their efficient replication in human cells. Even though the

314

36285 H5N1 virus did not have any amino acid markers that have been previously identified

315

to provide a replicative advantage in mammalian cells, it grew well in human lung cells. By

316

using a viral replication assay, we identified three amino acids in PB2 that are responsible for

317

the replicative efficiency of H5N1 virus in A549 cells. Viruses with these amino acids in PB2

318

circulate in nature; we did not generate viruses possessing ‘novel’ amino acid changes.

319

These data will provide information of value for pandemic preparedness and for use in

320

evaluating the pandemic risk potential of future isolates.

321

Pflug et al. (33) recently determined the complete crystal structure of the heterotrimeric

322

influenza A polymerase bound to the vRNA. This structure revealed that all three amino acid

323

substitutions identified in this study (i.e., PB2-E249G, G309D, and T339M) are located on

324

the surface of the PB2 protein. This finding supports the concept that these residues likely

325

interact with viral or host proteins, leading to the replicative advantage of the 36285 virus.

326

PB2-E249G and T339M have an interesting common feature in that they belong to areas

327

identified as ‘cap-binding' sites (34-37). Influenza virus has developed a mechanism called

328

‘cap-snatching,’ to steal the cap from the host cellular RNA in order to synthesize viral

20

329

protein. The polymerase PB2 subunit binds to the 5’ cap of host pre-mRNAs, which are

330

cleaved after 10–13 nucleotides by the PA subunit (38-40). An X-ray structure of the

331

cap-binding domain shows that the amino acid at position 339 of PB2 is located at the edge

332

of the cap-binding pocket (37). The precise cap-binding site, however, remains controversial;

333

one study identified the PB2 amino acid residues responsible for RNA cap-binding as

334

PB2-363F and PB2-404F (35), but another study identified the PB2 amino acid residues

335

242–282 and 538–577 as important (36), and yet another study suggested that amino acid

336

residues 318–483 were involved (37). The effect of the amino acid 339T in vivo and in vitro

337

on biologic events is also inconsistent. We recently found that 339T has a major impact on

338

the polymerase activity of H5N1 virus jointly with 147T and 588T (28). However, another

339

paper reported that K339T reduced PB2 cap binding activity and influenza polymerase

340

activity in vitro, and also attenuated virulence in mice (41). This latter paper suggested that

341

the K339T substitution in PB2 reduced the cap binding affinity because the side chain of

342

threonine is shorter than that of lysine and threonine is uncharged (41). In the current study,

343

we found that PB2 amino acid T339M helped to facilitate virus growth in A549 cells. Both

344

threonine and methionine are uncharged but the side chain of methionine is longer than that

345

of threonine. Although the function of the three amino acid mutations identified in this study

346

remains unclear, our findings may help improve our understanding of PB2 functions.

347

Human influenza A virus acquired the PB2 amino acid mutation E249G upon mouse

21

348

adaptation (42), suggesting that the amino acid at position 249 probably is involved in some

349

mechanism for adaptation to mammals. Sequence analysis showed that the prevalence of

350

PB2 amino acid 249G in human seasonal H3N2 viruses was almost 50%. Almost all of the

351

seasonal human H3N2 viruses isolated since 2006 have glycine at position 249, implying

352

that some positive pressure for adaptation to humans led to glycine selection over aspartic

353

acid. These data reinforce the notion that PB2 amino acid 249G plays a role in adaptation to

354

humans or mammals.

355

The MLD50 values of 36285-RG virus and Wb/AH82-RG virus in mice differed by one log unit

356

(Fig. 4), whereas the titers of the two viruses in A549 cells differed by more than 3 log units

357

(Fig. 3). The HA and NA genes are involved, to some extent, in the difference in the growth

358

ability of Wb/AH82-RG and 36285-RG in A549 cells. The pathogenicity in mice could have

359

been influenced by HA, because the dominant types of receptor to influenza viruses that are

360

distributed on bronchi and lungs differ with the host species (43-46). Humans primarily have

361

sialic acid linked to galactose by an α2,6-linkage (Sia-α2,6Gal) (43-45), whereas mice

362

primarily have Sia-α2,3Gal-type linkages (46). One study showed that on the surface of

363

A549 cells there are large amounts of Sia-α2,6Gal and a small amount of Sia-α2,3Gal (47).

364

Perhaps, the difference between the receptor specificities of the viruses and the receptor

365

types displayed on the lung cells of mice may have influenced the pathogenicity in mice.

366

In conclusion, here we found that the PB2 amino acid substitutions E249G, G309D, and

22

367

T339M enhance the replicative ability of H5N1 virus in A549 cells. Our study suggests that

368

these PB2 substitutions could assist H5N1 viruses in adapting to human lung cells. Although

369

the contribution of the PB2 C-terminal domain to virus host range is now well established,

370

that of the middle portion of the PB2 segment remains largely unknown. The full structure of

371

PB2 is needed to better understand its functions and role in host adaptation.

372 373

23 374

Acknowledgements

375

We are grateful to Susan Watson for editing the manuscript. We thank Kiyoko

376

Iwatsuki-Horimoto and Maki Kiso for technical assistance with mouse experiments. We also

377

thank Shufang Fan, Masato Hatta, Gabriele Neumann, Amie J. Eisfeld, and Takeo Gorai for

378

fruitful discussion. This work was supported by the Japan Initiative for Global Research

379

Network on Infectious Diseases from the Ministry of Education, Culture, Sports, Science and

380

Technology, Japan; by grants-in-aid from the Ministry of Health, Labour and Welfare, Japan;

381

and by the National Institute of Allergy and Infectious Diseases (NIAID), the National

382

Institutes of Health (NIH).

383 384 385 386

The authors do not have any competing interests.

24 387

References

388

1) K. Subbarao, A. Klimov, J. Katz, H. Regnery, H. Hall, C. Bender, J. Huang, M.

389

Hemphill, T. Rowe, M. Shaw, X. Xu, K. Fukuda, N. Cox. 1998. Characterization of an

390

avian influenza A (H5N1) virus isolated from a child with a fatal respiratory illness.

391

Science. 279 : 393-396

392

2) Xiyan Xu, Kanta Subbarao, Nancy J. Cox, and Yuanji Guo. 1999. Characterization of

393

the pathogenic influenza A/Goose/Guangdong/1/96 (H5N1) virus: similarity of its

394

hemagglutinin gene to those of H5N1 viruses from the 1997 outbreaks in Hong Kong.

395

Virology. 15 : 15-19

396

3) H. Chen, G. J. D. Smith, S. Y. Zhang, K. Qin, J. Wang, K. S. Li, R. G. Webster, J. S. M.

397

Peiris & Y. Guan. 2005. Avian flu: H5N1 virus outbreak in migratory waterfowl. Nature.

398

14: 191-192

399

4) The Writing Committee of the World Health Organization (WHO) Consultation on

400

Human Influenza A/H5. 2005. Avian Influenza A (H5N1) Infection in Humans. N Engl J

401

Med 353:1374-1385

402

5) The Writing Committee of the World Health Organization (WHO) Consultation on

403

Human Influenza A/H5. 2008. Update on Avian Influenza A (H5N1) Virus Infection in

404

Humans. N Engl J Med 358:261-273

405 406

6) Wang H, Feng Z, Shu Y, Yu H, Zhou L, Zu R, Huai Y, Dong J, Bao C, Wen L, Wang H, Yang P, Zhao W, Dong L, Zhou M, Liao Q, Yang H, Wang M, Lu X, Shi Z, Wang W,

25

407

Gu L, Zhu F, Li Q, Yin W, Yang W, Li D, Uyeki TM, Wang Y. 2008. Probable limited

408

person-to-person transmission of highly pathogenic avian influenza A (H5N1) virus in

409

China. Lancet. 371:1427-1434

410

7) Ungchusak K, Auewarakul P, Dowell SF, Kitphati R, Auwanit W, Puthavathana P,

411

Uiprasertkul M, Boonnak K, Pittayawonganon C, Cox NJ, Zaki SR, Thawatsupha P,

412

Chittaganpitch M, Khontong R, Simmerman JM, Chunsutthiwat S. 2005. Probable

413

person-to-person transmission of avian influenza A (H5N1). N Engl J Med. 352:333-340

414

8) Aggarwal S, Dewhurst S, Takimoto T, Kim B. 2011. Biochemical impact of the host

415

adaptation-associated PB2 E627K mutation on the temperature-dependent RNA

416

synthesis kinetics of influenza A virus polymerase complex. J Biol Chem.

417

286:34504-34513

418

9) Steel J, Lowen AC, Mubareka S, Palese P. 2009. Transmission of influenza virus in a

419

mammalian host is increased by PB2 amino acids 627K or 627E/701N. PLoS Pathog. 5,

420

e1000252

421

10) Hatta M, Hatta Y, Kim JH, Watanabe S, Shinya K, Nguyen T, Lien PS, Le QM,

422

Kawaoka Y. 2007. Growth of H5N1 influenza A viruses in the upper respiratory tracts of

423

mice. PLoS Pathog. 3:1374-1379

424 425

11) Yamada S, Hatta M, Staker BL, Watanabe S, Imai M, Shinya K, Sakai-Tagawa Y, Ito M, Ozawa M, Watanabe T, Sakabe S, Li C, Kim JH, Myler PJ, Phan I, Raymond A,

26

426

Smith E, Stacy R, Nidom CA, Lank SM, Wiseman RW, Bimber BN, O'Connor DH,

427

Neumann G, Stewart LJ, Kawaoka Y. 2010. Biological and structural characterization

428

of a host-adapting amino acid in influenza virus. PLoS Pathog. 6, e1001034

429

12) Tarendeau F, Boudet J, Guilligay D, Mas PJ, Bougault CM, Boulo S, Baudin F,

430

Ruigrok RW, Daigle N, Ellenberg J, Cusack S, Simorre JP, Hart DJ. 2007. Structure

431

and nuclear import function of the C-terminal domain of influenza virus polymerase PB2

432

subunit. Nature Structural & Molecular Biology 14, 229–233

433

13) Li Z, Chen H, Jiao P, Deng G, Tian G, Li Y, Hoffmann E, Webster RG, Matsuoka Y,

434

Yu K. 2005. Molecular basis of replication of duck H5N1 influenza viruses in a

435

mammalian mouse model. J Virol. 79:12058-12064

436

14) Bussey KA, Bousse TL, Desmet EA, Kim B, Takimoto T. 2010. PB2 residue 271

437

plays a key role in enhanced polymerase activity of influenza A viruses in mammalian

438

host cells. J Virol. 84:4395-4406

439

15) Zhou B, Li Y, Halpin R, Hine E, Spiro DJ, Wentworth DE. 2011. PB2 residue 158 is a

440

pathogenic determinant of pandemic H1N1 and H5 influenza a viruses in mice. J Virol.

441

85:357-365

442

16) Hu J, Hu Z, Mo Y, Wu Q, Cui Z, Duan Z, Huang J, Chen H, Chen Y, Gu M, Wang X,

443

Hu S, Liu H, Liu W, Liu X, Liu X. 2013. The PA and HA gene-mediated high viral load

444

and intense innate immune response in the brain contribute to the high pathogenicity of

27

445 446

H5N1 avian influenza virus in mallard ducks. J Virol. 87:11063-11075 17) Dankar SK, Miranda E, Forbes NE, Pelchat M, Tavassoli A, Selman M, Ping J, Jia J,

447

Brown EG. 2013. Influenza A/Hong Kong/156/1997(H5N1) virus NS1 gene mutations

448

F103L and M106I both increase IFN antagonism, virulence and cytoplasmic localization

449

but differ in binding to RIG-I and CPSF30. Virol J. 10:243

450 451 452

18) Shinya K, Ebina M, Yamada S, Ono M, Kasai N, Kawaoka Y. 2006. Avian flu: Influenza virus receptors in the human airway. Nature. 440, 435-436 19) Kim JH, Hatta M, Watanabe S, Neumann G, Watanabe T, Kawaoka Y. 2010. Role of

453

host-specific amino acids in the pathogenicity of avian H5N1 influenza viruses in mice. J

454

Gen Virol. 91:1284-1289

455

20) Mänz B, Brunotte L, Reuther P, Schwemmle M. 2012. Adaptive mutations in NEP

456

compensate for defective H5N1 RNA replication in cultured human cells. Nat Commun.

457

3:802

458

21) Xu C, Hu WB, Xu K, He YX, Wang TY, Chen Z, Li TX, Liu JH, Buchy P, Sun B. 2012.

459

Amino acids 473V and 598P of PB1 from an avian-origin influenza A virus contribute to

460

polymerase activity, especially in mammalian cells. J Gen Virol. 93:531-540

461

22) Song MS, Pascua PN, Lee JH, Baek YH, Lee OJ, Kim CJ, Kim H, Webby RJ,

462

Webster RG, Choi YK. 2009. The polymerase acidic protein gene of influenza a virus

463

contributes to pathogenicity in a mouse model. J Virol. 83:12325-12335

28

464

23) Mehle A, Dugan VG, Taubenberger JK, Doudna JA. 2012. Reassortment and

465

mutation of the avian influenza virus polymerase PA subunit overcome species barriers.

466

J Virol. 86:1750-1757

467

24) Mänz B, Dornfeld D, Götz V, Zell R, Zimmermann P, Haller O, Kochs G, Schwemmle

468

M. 2013. Pandemic influenza A viruses escape from restriction by human MxA through

469

adaptive mutations in the nucleoprotein. PLoS Pathog. 9:e1003279

470

25) Ping J, Keleta L, Forbes NE, Dankar S, Stecho W, Tyler S, Zhou Y, Babiuk L,

471

Weingartl H, Halpin RA, Boyne A, Bera J, Hostetler J, Fedorova NB, Proudfoot K,

472

Katzel DA, Stockwell TB, Ghedin E, Spiro DJ, Brown EG. 2011. Genomic and protein

473

structural maps of adaptive evolution of human influenza A virus to increased virulence

474

in the mouse. PLoS One. 6:e21740

475

26) Mänz B, Schwemmle M, Brunotte L. 2013. Adaptation of avian influenza A virus

476

polymerase in mammals to overcome the host species barrier. J Virol. 87:7200-7209

477

27) Jiao P, Tian G, Li Y, Deng G, Jiang Y, Liu C, Liu W, Bu Z, Kawaoka Y, Chen H. 2008.

478

A single-amino-acid substitution in the NS1 protein changes the pathogenicity of H5N1

479

avian influenza viruses in mice. J Virol. 82:1146-1154

480

28) Fan S, Hatta M, Kim JH, Halfmann P, Imai M, Macken CA, Le MQ, Nguyen T,

481

Neumann G, Kawaoka Y. 2014. Novel residues in avian influenza virus PB2 protein

482

affect virulence in mammalian hosts. Nature Communications. 5:5021

29

483

29) Heiny AT, Miotto O, Srinivasan KN, Khan AM, Zhang GL, Brusic V, Tan TW, August

484

JT. 2007. Evolutionarily conserved protein sequences of influenza A viruses, avian and

485

human, as vaccine targets. PLoS One.2:e1190

486

30) Miotto O, Heiny A, Tan TW, August JT, Brusic V. 2008. Identification of

487

human-to-human transmissibility factors in PB2 proteins of influenza A by large-scale

488

mutual information analysis. BMC Bioinformatics. 9(Suppl 1): S18

489

31) Miotto O, Heiny AT, Albrecht R, García-Sastre A, Tan TW, August JT, Brusic V. 2010.

490

Complete-proteome mapping of human influenza A adaptive mutations: implications for

491

human transmissibility of zoonotic strains. PLoS One. 5:e9025.

492

32) Neumann G, Watanabe T, Ito H, Watanabe S, Goto H, Gao P, Hughes M, Perez DR,

493

Donis R, Hoffmann E, Hobom G, Kawaoka Y. 1996. Generation of influenza A viruses

494

entirely from cloned cDNAs. Proc Natl Acad Sci U S A. 96:9345-9350

495 496

497 498

33) Pflug A, Guilligay D, Reich S, Cusack S. 2014. Structure of influenza A polymerase bound to the viral RNA promoter. Nature. 516:355-360. 34) Fechter P, Brownlee GG. 2005. Recognition of mRNA cap structures by viral and cellular proteins. J Gen Virol. 86:1239-1249

499

35) Fechter P, Mingay L, Sharps J, Chambers A, Fodor E, Brownlee GG. 2003. Two

500

aromatic residues in the PB2 subunit of influenza A RNA polymerase are crucial for cap

501

binding. J Biol Chem. 278:20381-20388

30

502

36) Honda A, Mizumoto K, Ishihama A. 1999. Two separate sequences of PB2 subunit

503

constitute the RNA cap-binding site of influenza virus RNA polymerase. Genes Cells.

504

4:475-485

505

37) Guilligay D, Tarendeau F, Resa-Infante P, Coloma R, Crepin T, Sehr P, Lewis J,

506

Ruigrok RW, Ortin J, Hart DJ, Cusack S. 2008. The structural basis for cap binding by

507

influenza virus polymerase subunit PB2. Nat Struct Mol Biol. 15:500-506

508

38) Dias A, Bouvier D, Crépin T, McCarthy AA, Hart DJ, Baudin F, Cusack S, Ruigrok

509

RW. 2009. The cap-snatching endonuclease of influenza virus polymerase resides in the

510

PA subunit. Nature. 458:914-918

511

39) Yuan P, Bartlam M, Lou Z, Chen S, Zhou J, He X, Lv Z, Ge R, Li X, Deng T, Fodor E,

512

Rao Z, Liu Y. 2009. Crystal structure of an avian influenza polymerase PA(N) reveals an

513

endonuclease active site. Nature. 458:909-913

514

40) Tomassini J, Selnick H, Davies ME, Armstrong ME, Baldwin J, Bourgeois M,

515

Hastings J, Hazuda D, Lewis J, McClements W. 1994. Inhibition of cap

516

(m7GpppXm)-dependent

517

2,4-dioxobutanoic acid compounds. Antimicrob Agents Chemother. 38:2827-2837

endonuclease

of

influenza

virus

by

4-substituted

518

41) Liu Y, Qin K, Meng G, Zhang J, Zhou J, Zhao G, Luo M, Zheng X. 2013. Structural

519

and functional characterization of K339T substitution identified in the PB2 subunit

520

cap-binding pocket of influenza A virus. J Biol Chem. 288:11013-11023

31

521

42) Ping J, Keleta L, Forbes NE, Dankar S, Stecho W, Tyler S, Zhou Y, Babiuk L,

522

Weingartl H, Halpin RA, Boyne A, Bera J, Hostetler J, Fedorova NB, Proudfoot K,

523

Katzel DA, Stockwell TB, Ghedin E, Spiro DJ, Brown EG. 2011. Genomic and protein

524

structural maps of adaptive evolution of human influenza A virus to increased virulence

525

in the mouse. PLoS One. 6:e21740

526

43) Rogers GN, Paulson JC. 1983. Receptor determinants of human and animal influenza

527

virus isolates: differences in receptor specificity of the H3 hemagglutinin based on

528

species of origin. Virology. 127:361-373

529

44) Connor RJ, Kawaoka Y, Webster RG, Paulson JC. 1994. Receptor specificity in

530

human, avian, and equine H2 and H3 influenza virus isolates. Virology. 205:17-23

531

45) Ibricevic A, Pekosz A, Walter MJ, Newby C, Battaile JT, Brown EG, Holtzman MJ,

532

Brody SL. 2006. Influenza virus receptor specificity and cell tropism in mouse and

533

human airway epithelial cells. J Virol. 80:7469-7480

534

46) Stevens J, Blixt O, Tumpey TM, Taubenberger JK, Paulson JC, Wilson IA. 2006.

535

Structure and receptor specificity of the hemagglutinin from an H5N1 influenza virus.

536

Science. 312:404-410

537

47) Takeyuki K. 2009. A barrier function made up of the receptor preference of the virus and

32

538

the paucity of receptors in human airways against avian influenza virus infection.

539

Kawasaki Medical Journal. 35:307-318

540 541

33 542

Figure legends

543

Fig. 1: Comparison of the growth properties of H5N1 viruses isolated from humans

544

and birds.

545

(A) A549 cells were infected with H5N1 human viruses isolated in Vietnam in 2010 at a

546

multiplicity of infection (m.o.i.) of 0.0002 and cultured at 37℃. Virus release into cell culture

547

supernatant was titrated by plaque assays with MDCK cells at the indicated time points. The

548

standard deviation represents 3 independent experiments.

549

(B) A549 cells were infected with the H5N1 human viruses 36285-wild type and 36285-RG

550

at an m.o.i. of 0.0002 and cultured at 37℃. Error bars represent the standard deviation of

551

three independent experiments.

552

(C) A549 cells were infected with H5N1 human virus 36285-RG and avian viruses at an m.o.i.

553

of 0.0002 and cultured at 37℃. Supernatants were titrated by plaque assays with MDCK

554

cells at the indicated time points. The standard deviation represents 3 independent

555

experiments. *The titers of the 36285-RG virus are significantly different from those of the

556

Wb/AH82-RG virus (p

Identification of PB2 mutations responsible for the efficient replication of H5N1 influenza viruses in human lung epithelial cells.

Highly pathogenic H5N1 avian influenza viruses have caused outbreaks among poultry worldwide, resulting in sporadic infections in humans with approxim...
1002KB Sizes 2 Downloads 5 Views