JGV Papers in Press. Published November 4, 2014 as doi:10.1099/vir.0.069914-0

1

JGV paper no. VIR/2014/069914

2

Title: RIG-I from waterfowl and mammals have different abilities to induce antiviral responses

3

against influenza A viruses

4

Running title: Various RIG-Is have different antiviral abilities on IAV

5

Subject Category: Animal Viruses-Negative-strand RNA

6

Word count: 5297

7

Authors: Qiang Shao, Wenpin Xu, Qiang Guo, Li Yan, Lei Rui, Jinhua Liu, Yaofeng

8

Zhao & Zandong Li

9

Affiliation: Correspondence to Zandong Li, PhD, [email protected]

10

Qiang Shao and Wenpin Xu contributed equally to this work.

11

State key Laboratory for Agrobiotechnology, College of Biological Sciences, China

12

Agricultural University, No.2 Yuan Ming Yuan West Road, Beijing, 100193, China

13

Tel.: 86-10-62732144. Fax: 86-10-62895439

14

Abstract

15

Retinoic acid-induced gene I (RIG-I), plays a crucial role in sensing viral RNA and facilitating the

16

production of IFN-β. It varies in length and sequence among different species. The present study

17

assessed the functional differences among RIG-I proteins derived from mammals and birds. The

18

transfection of duck caspase recruitment domains (CARDs) and RIG-I (dCARDs and dRIG-I) and

19

goose CARDs and RIG-I (gCARDs and gRIG-I) into DF-1 cells increased the production of

20

IFN-β mRNA and IFN-stimulated genes and decreased influenza A virus (IAV) replication,

21

whereas, human CARDs and RIG-I (hCARDs and hRIG-I) and mouse CARDs and RIG-I

22

(mCARDs and mRIG-I) had no effect. In 293T and A549 cells, hCARDs had the strongest

23

IFN-inducing activity, followed by mCARDs, dCARDs, and gCARDs. The IFN-inducing activity

24

of hRIG-I was stronger than that of mRIG-I, dRIG-I, and gRIG-I, in that order. The results also

25

showed that although the ability dCARDs to activate IFN was stronger than that of gCARDs in

26

DF-1, 293T, and A549 cells, dRIG-I had a weaker ability to activate IFN than gRIG-I in DF-1

27

with or without IAV infection. Taken together, these data suggest that RIG-I protein has different

28

amino acid sequences and functions among species. This genetic and functional diversity renders

29

it flexible, adaptable, and capable of recognizing many viruses in different species. 1

30

Key words: influenza; H1N1; H9N2; RIG-I; CARDs; IFN;

31 32

Introduction

33

Type-I IFNs and other pro-inflammatory cytokines are produced when host pattern

34

recognition receptors (PRRs), such as Toll-like receptors (TLRs) and retinoic acid-induced gene I

35

(RIG-I)-like receptors (RLRs), recognize pathogen-associated molecular patterns (PAMPs)

36

including nucleic acids or other structural components of microbes (Creagh and O’Neill, 2006;

37

Kawai and Akira, 2009). RIG-I, as a cytoplasmic sensor of viral double strand RNA (dsRNA) and

38

single strand RNA (ssRNA), can be triggered by viral RNA (vRNA). RIG-I is composed of two

39

N-terminal CARDs, a central DExD/H box helicase/ATPase, and a C-terminal regulatory domain

40

(CTD/RD) (Hausmann et al., 2008). Under normal conditions, RIG-I is in an inactive, closed

41

conformation (Takahasi et al., 2008). Viral ssRNA or short dsRNA bearing a 5′-triphosphate can

42

bind to the DExD/H box helicase domain at its carboxy-terminus, which causes RIG-I to undergo

43

a conformational change that exposes the CARDs to the cytoplasm (Saito et al., 2007). The

44

exposure of the N-terminal CARDs of RIG-I leads to the Lys63-linked ubiquitination of Lys172.

45

The ubiquitinated CARDs then allow RIG-I bind to MAVS (also called IPS-1, CARDsif, or VISA)

46

(Kawai et al., 2005; Meylan et al., 2005; Xu et al., 2005). Other components of the pathway are

47

also recruited to this multi-protein signal complex, which activates TBK1 (TANK-binding kinase

48

1) and IKK (IκB kinase) (Meylan et al., 2005; Xu et al., 2005). The activation of these kinases

49

results in the phosphorylation and activation of IRF-3 and IRF-7, which form homodimers and

50

heterodimers, translocate into the nucleus and bind to IFN-stimulated response elements (ISREs),

51

to stimulate the expression of type I IFN genes and a set of IFN-inducible genes (Honda et al.,

52

2005; Nakhaei et al., 2006).

53

RIG-I has been identified in many species, including mammals (humans, mice, pigs, and

54

rabbits), birds (ducks, geese, pigeons, and zebra finches), and fish (trout and carp). The signaling

55

pathways activated by hRIG-I and mRIG-I have been well illustrated previously (Loo and Gale Jr,

56

2011). As natural reservoir of most influenza A strains, waterfowl have a crucial role in the

57

ecology and evolution of IAV (Webster et al., 1992). Similar to hRIG-I and mRIG-I, dRIG-I and

58

gRIG-I also induce type I interferon after influenza infection and poly I:C transfection (Barber et

59

al., 2010; Sun et al., 2013). Although dRIG-I and gRIG-I share similar CARDs and a helicase 2

60

domain with hRIG-I and mRIG-I, S8 and T170, which are critical for phosphorylation, and T55

61

and K172, which are essential for interaction and polyubiqutination by tripartite motif containing

62

25 (TRIM25) in humans, are not conserved in mice, ducks, or geese. Recently, Domingo et al.

63

demonstrated that although TRIM25-mediated ubiquitination occurs at K167 and K193, the

64

activation of dRIG-I is independent of the anchored ubiquitination (Miranzo-Navarro and Magor,

65

2014). These results suggest that the RIG-I signaling pathway functions differently in waterfowl

66

and mammals (Miranzo-Navarro and Magor, 2014).

67

In this study, the ability of the CARDs of RIG-I and full-length RIG-I encoded by humans,

68

mice, ducks, and geese to induce type I IFN in chicken DF-1, human A549, and human 293T cells

69

during IAV infection from mice, swine, and avian species was examined. Results showed that the

70

ability of waterfowl RIG-I to induce the expression of antiviral genes was stronger than

71

mammalian RIG-I in DF-1 cells during IAV infection. However, the ability of mammalian RIG-I

72

to induce the expression of antiviral genes was stronger than that of waterfowl RIG-I in 293T and

73

A549 cells during IAV infection. Waterfowl RIG-I interfered with the replication of IAV more

74

efficiently than mammalian RIG-I in DF-1 cells, whereas, mammalian RIG-I was more efficient

75

than waterfowl RIG-I in A549 and MDCK cells.

76 77

2. Results

78

2.1 Protein expression of RIG-I and CARDs in chicken DF-1 cells and human 293T cells.

79

EGFP-tagged RIG-I and CARDs were generated and transfected into 293T or DF-1 cells.

80

The detection of EGFP-tagged RIG-I and CARDs using fluorescence microscopy indicated that

81

significant levels of EGFP-tagged RIG-I and CARDs accumulated in 293T and DF-1 cells (Fig.

82

1A, B). The fluorescence intensity of waterfowl-derived RIG-I and CARDs was significantly

83

higher than that of mammalian-derived RIG-I and CARDs (Fig. 1A, B), respectively. In addition,

84

the expression of EGFP-tagged CARDs and RIG-I proteins derived from humans, mice, ducks,

85

and geese was detected in DF-1 (Fig. 1C) and 293T cells (Fig. 1D) using western blotting with

86

anti-EGFP antibody. Similar to the fluorescence observation, the expression of hCARDs and

87

mCARDs protein was significantly lower than that of the dCARDs and gCARDs as well as

88

hRIG-I, mRIG-I, dRIG-I, and gRIG-I in DF-1 and 293T cells. These data suggest that

89

waterfowl-derived RIG and CARDs are expressed more efficiently than mammal-derived RIG and 3

90

CARDs in both DF-1 and 293T cells.

91 92

2.2 Waterfowl RIG-Is induce the expression of antiviral genes more strongly than

93

mammalian RIG-I in DF-1 cells during IAV infection

94

To evaluate the differences in IFN-inducing activity between mammal-derived and

95

waterfowl-derived RIG-I and CARDs variants in chicken DF-1 cells, cells were transfected with

96

EGFP-tagged RIG-I and CARDs expression plasmids (Fig. 1A, C), and then infected with

97

influenza viruses or mock-treated. The mRNA expression of IFN-β and the antiviral

98

IFN-stimulated genes Mx1 and PKR were measured using qRT-PCR at 8 h post-infection (p.i.)

99

with A/Swine/Shandong/731/2009 (H1N1) (SW731), A/Chicken/Shandong/ZB/2007 (H9N2)

100

(ZB07), or A/WSN/33 (H1N1) (WSN) viruses. Transfection with dCARDs, gCARDs, dRIG-I,

101

and gRIG-I transfection alone induced significantly more IFN-β, Mx1, and PKR synthesis

102

compared with empty vector transfection with or without IAV infection (Fig. 2A–F). However,

103

hCARDs, mCARDs, hRIG-I, and mRIG-I did not increase the mRNA production of IFN-β, Mx1,

104

or PKR compared with empty vector transfection with or without IAV infection (Fig. 2A–F).

105

Transfection with EGFP-mRIG-I inhibited IFN-β, Mx1, and PKR mRNA synthesis significantly

106

only in WSN infection group (Fig. 2B, D, and F). Although dCARDs induced more IFN-β, Mx1,

107

and PKR mRNA production than did gCARDs, dRIG-I induced significant less IFN-β, Mx1, and

108

PKR mRNA production than did gRIG-I (Fig. 2A–F). ZB07, WSN, and SW731 infection was also

109

blocked the gRIG-I-mediated increase in Mx1 and PKR mRNA compared with the mock-treated

110

group. The inhibitory effects of WSN were the strongest (Fig. 2D, F). These observations suggest

111

that transfected waterfowl RIG-I and CARDs have a stronger ability to induce the expression of

112

antiviral genes than mammalian RIG-I and CARDs in DF-1 cells during IAV infection. Moreover,

113

dCARDs was more efficient in IFN-β mRNA production than gCARDs.

114 115

2.3 Mammalian RIG-Is induce the expression of antiviral genes more strongly than

116

waterfowl RIG-Is in human cells during IAV infection

117

To assess the differences in IFN-inducing activity between mammalian and waterfowl RIG-Is

118

in 293T cells, cells were transfected with EGFP-tagged RIG-I and CARDs expression plasmids

119

(Fig. 1B, D), and then infected with influenza viruses or mock-treated. The mRNA expression of 4

120

IFN-β, Mx, and PKR was detected using qRT-PCR at 8 h p.i. hCARDs and mCARDs induced

121

more IFN-β mRNA production compared with dCARDs and gCARDs in the presence or absence

122

of IAV infection (Fig. 3A). hCARDs had the strongest ability to induce IFN-β, followed by

123

mCARDs, dCARDs, and gCARDs (Fig. 3A). Similarly, hRIG-I had the most pronounced ability

124

to induce IFN-β, followed by mRIG-I, dRIG-I, and gRIG-I (Fig. 3B). Although gRIG-I

125

transfection alone increased IFN-β synthesis (Fig. 3B), it did not further increase the production of

126

IFN-β over empty vector transfection after IAV infection (Fig. 3B). Interestingly, gRIG-I blocked

127

IFN-β mRNA production during WSN or SW731 infection (Fig. 3B).

128

Because of the low expression of Mx and PKR in 293T cells (data not shown), we also

129

performed the transfections and infections in a human lung adenocarcinoma epithelial cell

130

line (A549 cells). Fig.4A and B shows that hRIG-I, mRIG-I, dRIG-I, gRIG-I, hCARDs, mCARDs,

131

dCARDs, and gCARDs had the similar IFN-β-inducing ability in A549 and 293T cells. Fig. 4C

132

and E show that hCARDs induced the highest mRNA levels of Mx and PKR, followed by

133

mCARDs, dCARDs, and gCARDs. Similarly, hRIG-I had the strongest ability to induce IFN-β,

134

followed by mRIG-I, dRIG-I, and gRIG-I (Fig. 4D, F). Although gCARDs facilitated IFN-β, Mx

135

and PKR mRNA production (Fig. 4A, C, and E), gRIG-I inhibited IFN-β, Mx and PKR mRNA

136

synthesis compared with empty vector transfection after IAV infection (Fig. 4B, D, and F).

137

Taken together, these data suggest that the transfected mammalian RIG-Is and CARDs had a

138

stronger ability to induce the expression of antiviral genes in 293T and A549 cells during IAV

139

infection than did waterfowl. In addition, gRIG-I could negatively regulate IFN-β signaling

140

pathway in 293T and A549 cells during IAV infection.

141 142

2.4 Waterfowl RIG-Is reduce the replication of IAV more efficiently than mammalian

143

RIG-Is in DF-1 cells

144

We next evaluated the effects of all the CARDs and RIG-Is on viral replication in DF-1 cells.

145

Cells were transfected with EGFP-tagged RIG-I and CARDs plasmids or empty vector alone, and

146

then were challenged with ZB07, WSN, or SW731 at a multiplicity of infection (MOI) of 0.01

147

(Fig. 5). At 24 and 48 h p.i., the viral titers in the cell cultures were determined as 50% tissue

148

culture infectious doses (TCID50/ml). No differences were observed between the ZB07, WSN, or

149

SW731 viral titers in hCARDs-, mCARDs-, hRIG-I-, mRIG-I-, or empty vector-transfected DF-1 5

150

cells at 24 or 48 h p.i. (Fig. 5A–F). However, significantly decreased ZB07, WSN, and SW731

151

viral titers were observed in dCARDs-, gCARDs-, dRIG-I-, and gRIG-I-transfected DF-1cells

152

compared with empty vector transfected DF-1 cells at both 24 and 48 h p.i. (Fig. 5A–F). Similar to

153

the observations regarding IFN-inducing activity, dCARDs, gCARDs decreased the ZB07, WSN,

154

and SW731 viral titers to a lower level than dRIG-I and gRIG-I in DF-1 cells (Fig. 5A–F).

155

Although dCARDs induced more IFN-β mRNA than did gCARDs did, there were no difference

156

between the ZB07 and WSN viral titers on in dCARDs- and gCARDs- transfected DF-1 cells.

157

dCARDs decreased SW731 titers to a lower level than did gCARDs (Fig. 5A). Taken together,

158

these results suggest that waterfowl RIG-Is reduce IAV replication more efficiently than

159

mammalian RIG-Is in DF-1 cells, and gCARDs have a weaker ability to inhibit replication of IAV

160

than dCARDs.

161 162

2.5 Mammalian RIG-I reduces IAV replication more efficiently than waterfowl RIG-I in

163

A549 cells

164

The effects of mammalian and waterfowl RIG-Is on viral multiplication was evaluated

165

further in A549 cells. Cells were transfected with EGFP-tagged CARDs and RIG-I expression

166

plasmids or empty vector alone, and were then challenged with ZB07, WSN, or SW731 at an MOI

167

of 0.01. The viral titers were determined as TCID50 at 24 and 48 h p.i.. hCARDs and mCARDs

168

exhibited the most significant inhibitory effects on the multiplication of ZB07, WSN, and SW731

169

at both 24 and 48 h (Fig. 6A, C, and E). In contrast, dCARDs only slightly decreased the of WSN

170

and SW731 viral titers at 24 h and ZB07 titers at 24 and 48 h (Fig. 6A, C, and E). No differences

171

were observed between the ZB07, WSN, or SW731 viral titers in A549 cells transfected with

172

gCARDs compared with empty vector at 24 or 48 h (Fig. 6A, C, and E). Consistent with hCARDs

173

and mCARDs, hRIG-I and mRIG-I suppressed the replication of the three influenza viruses at 24

174

and 48 h (Fig. 6B, D, and F). No inhibitory effects on viral titers were observed in dRIG-I- and

175

gRIG-I-transfected cells. Similar results were also observed in transfection and viral multiplication

176

experiments performed in Madin-Darby canine kidney cells (MDCK cells) (Supplementary Fig. 1).

177

These data suggest that mammalian RIG-I and CARDs reduce the replication of IAV in

178

mammalian cells more efficiently than do waterfowl RIG-I and CARDs.

179 6

180

3. Discussion

181

Here we examined the differences in the protein expression of RIG-I and CARDs proteins

182

derived from mammals and birds, as well as their IFN-inducing activity and ability to suppress the

183

replication of influenza viruses in chicken and human cells. QRT-PCR was used to analyze the

184

ability of various RIG-Is and CARDs variants to induce IFN expression in DF-1, 293T, and A549

185

cells. Waterfowl RIG-I and CARDs had a more pronounced ability to induce the expression of

186

antiviral genes than did mammalian in DF-1 cells after IAV infection. Conversely, mammalian

187

RIG-I and CARDs had a more pronounced ability to induce the expression of antiviral genes than

188

did waterfowl RIG-I and CARDs in 293T and A549 cells during IAV infection. Data revealed that

189

waterfowl RIG-I interfered with the replication of IAV more efficiently than did mammalian

190

RIG-I in DF-1 cells, whereas, mammalian RIG-I reduced replication of IAV more efficiently than

191

did waterfowl RIG-I in A549 and MDCK cells. Therefore, RIG-I from waterfowl and mammals

192

had different abilities to induce an antiviral response against IAV.

193

Previous studies revealed that duck and goose CARDs domains activate IFN-β in chicken

194

cells, suggesting that chicken cells contain all the proteins required for the activation and

195

downstream signaling of RIG-I (Miranzo-Navarro and Magor, 2014; Sun et al., 2013). Recent in

196

vitro studies demonstrated that human TRIM25 does not interact with, ubiquitinate, or activate

197

human CARDs domains that were co-transfected into chicken DF-1 cells (Miranzo-Navarro and

198

Magor, 2014). Similarly, dCARDs, gCARDs, dRIG-I, and gRIG-I activated IFN-β production in

199

chicken DF-1 cells, whereas hCARDs, mCARDs, hRIG-I, and mRIG-I did not. In the current

200

study, this suggests that only waterfowl-derived RIG-I are functional in chicken DF-1 cells, and

201

that some factors necessary for mammal-derived RIG-I activation are missing or unrecognizable

202

in chicken cells. For example, RIPLET, which ubiquitinates Lys849 and Lys851 in the CTD of

203

RIG-I (Gao et al., 2009; Oshiumi et al., 2009; Oshiumi et al., 2010), is absent from the chicken

204

and duck genomes (Magor et al., 2013; Rajsbaum et al., 2012).

205

SW731, WSN, and ZB07 have different IFN-inducing activity (Fig.2A and B; Fig. 3A and B).

206

Sequence alignment showed that three viruses have different length of non-structural

207

protein1(NS1) (SW731, 219 amino acid (aa); WSN, 230 aa; ZB07, 217 aa) (data not shown). The

208

most pronounced IFN synthesis of ZB07 infected DF-1 and 293T cells (Fig.2A and B; Fig. 3A and

209

B), may due to the weak suppression of IFN by NS1. Otherwise, mRIG-I, but not mCARD, 7

210

blocked the IFN-β, Mx, and PKR mRNA synthesis only in WSN infected DF-1 cells (Fig.2A–F)..

211

Previously study reported that only avian influenza virus (AIV) derived non-structural protein1

212

(NS1) can bind to chicken TRIM25, thereby suppressing the induction of IFN (Rajsbaum et al.,

213

2012). Mouse-adapted IAV NS1 binds to mouse RIPLET specifically in mouse cells (Rajsbaum et

214

al., 2012). NS1 or other viral protein of WSN might specifically interact with mRIG-I to suppress

215

the induction of IFN in DF-1 cells.

216

Although gRIG-I induced slightly more IFN-β production in DF-1 cells than did dRIG-I with

217

or without IAV infection, the expression of gRIG-I induced significantly more Mx and PKR

218

expression (Fig. 2B, D, and F). This suggests that gRIG-I might induce Mx and PKR in

219

IFN-independent manner. Interestingly, dCARDs share 92.98% amino acid identity with gCARDs;

220

however, dCARDs induced significantly more IFN-β production and IFN-stimulated gene

221

expression than did gCARDs in both DF-1 and 293T cells with or without IAV infection. An

222

alignment of dRIG-I and gRIG-I sequences revealed that the phosphorylation sites for inactivation

223

are not conserved between ducks (S8) and geese (G8). It was reported that the phosphorylation of

224

S8 negatively regulates hRIG-I, possibly by preventing TRIM25 binding (Gack et al., 2010;

225

Nistal-Villán et al., 2010). However, it is unclear whether the S8G mutantion in gRIG-I is

226

responsible for its pronounced IFN-inducing activity.

227

Recent in vitro studies demonstrated that dRIG-I interacts with the unanchored K63-linked

228

polyubiquitin chains generated by hTRIM25 in 293T cells (Miranzo-Navarro and Magor, 2014).

229

Sun et al. demonstrated that NDV infection in gRIG-I-transfected 293T/17 cells resulted in

230

upregulated activity of the IFN-β promoter and IRF-3 and IFIT1 mRNA levels, but decreased viral

231

titers (Sun et al., 2013). Consistent with these reports, dCARDs, gCARDs, dRIG-I, and gRIG-I

232

were promoted IFN expression in mock-treated 293T and A549 cells in the current study.

233

However, the ability of dCARDs and gCARDs to induce IFN production was significantly less

234

pronounced than that of hCARDs and mCARDs. The ability of dRIG-I and gRIG-I to induce IFN

235

production was also reduced significantly compared with hRIG-I and mRIG-I. Surprisingly,

236

gRIG-I blocked IFN-β, Mx1, and PKR mRNA production during influenza infection, which is

237

almost the opposite effect reported for gRIG-I in 293T cells after NDV infection (Sun et al., 2013).

238

One possible cause of this inhibitory effect is that gRIG-I could compete with endogenous hRIG-I

239

to bind influenza vRNA. However, gRIG-I could not activate the IFN pathway as effectively as 8

240

could hRIG-I, which resulted in competitive inhibition.

241

Although mRIG-I protein has 76.57% amino acid identity with hRIG-I, hRIG-I induced

242

significantly more IFN-β production in 293T and A549 cells compared with mRIG-I with or

243

without IAV infection. mCARDs also share 75.09% amino acid identity with hCARDs, but

244

hCARDs induced significantly more IFN production and IFN-stimulated gene expression in 293T

245

and A549 cells than did mCARDs regardless of IAV infection. In the current study, hCARDs and

246

dCARDs have conserved phosphorylation sites (S8 and S168 in duck, S8 and T170 in human), but

247

mCARDs contain N8 and V170. Moreover, T55, which plays a critical role in the interaction, is

248

displaced by serine in mCARDs (Gack et al., 2008). K172, the site that is ubiquitinated by

249

TRIM25 in humans, is not conserved in mRIG-I. This might be responsible for the fact that

250

mCARDs and mRIG-I had reduced ability to induce IFN than did hCARDs and hRIG-I (Jiang et

251

al., 2012; Zeng et al., 2010).

252

In the current study, the effects of various RIG-Is and CARDs variants on virus replication

253

were evaluated in DF-1, A549, and MDCK cells. ZB07, WSN, and SW731 titers in hCARDs-,

254

mCARDs-, hRIG-I-, and mRIG-I-transfected DF-1 cells confirmed the inability of hCARDs,

255

mCARDs, hRIG-I, and mRIG-I to function in chicken DF-1 cells. The inability of dRIG-I and

256

gRIG-I to function in MDCK cells was also demonstrated by the ZB07, WSN, and SW731 titers

257

in dRIG-I- and gRIG-I-transfected A549 and MDCK cells. This suggests that the RIG-I signaling

258

pathway has changed during evolution, possibly because of selection pressure from viruses that

259

disrupted this regulation. Both birds and mammals evolved from reptiles 200 million years ago.

260

RIG-Is derived from humans and mice were not functional in chicken DF-1 cells. Although

261

RIG-Is derived from ducks and geese are functional in human 293T cells, but they do not

262

functions as efficiently as hRIG-I. This might be due to the altered amino acid sequence of RIG-I

263

during evolution and the emergence of novel signal transducers, such as RIPLET. Nevertheless,

264

waterfowl still retain part of the original signal transduction system. Therefore, waterfowl RIG-I

265

are still partially functional in human 293T cells. Taken together, these data suggest that RIG-I can

266

only activate the IFN signaling pathway effectively in its own or very similar cellular systems

267

(dRIG-I and gRIG-I in DF-1, and hRIG-I and mRIG-I in A549, 293T, and MDCK cells). Together,

268

these results suggest that additional attention should be paid to whether transgenes from very

269

evolutionary distant species can function in alternative hosts. 9

270 271

4. Materials and methods

272

4.1 Plasmids.

273

The open reading frame (ORF) of hRIG-I was cloned from cDNA isolated from human A549

274

cells

275

TTTGGACATTTCTGCTGGATCA-3′ based on human RIG-I sequence (GenBank accession no.

276

AF038963). The mRIG-I ORF was cloned from mouse lung cDNA using the primers

277

5′-ATGACCGCGGCGCAGCGGC-3′

278

based on mRIG-I sequence (GenBank accession no. AY553221). The dRIG-I ORF was obtained

279

from

280

5′-ATGACGGCGGACGAGAAGC-3 ′ and 5 ′-CTAAAATGGTGGGTACAAGTTGGAC-3 ′ based

281

on the duck RIG-I sequence (GenBank accession no. EU363349.1). Finally, the gRIG-I ORF was

282

amplified from goose splenic cDNA using the primers 5′-ATGACGGCGGAGGAAAAGC-3′ and

283

5′-CTA AAATGGTGGGTACAAGTTGGAC-3′ based on goose RIG-I sequence (GenBank

284

accession no. JF804977). The homologous arms of pEGFP-N1 were added to the 5′ and 3′ termini

285

of human, mouse, duck, and goose full-length RIG-I and CARDs from human, mouse, duck, and

286

goose RIG-I using PCR with the primers shown in Supplementary Table 1. Human, mouse, duck,

287

and goose full-length RIG-I and CARDs were cloned into mammalian expression vector

288

pEGFP-N1 by homologous recombination in accordance with the manufacturer’s instructions

289

(NovoRec).

290

4.2 Viruses.

using

duck

the

(A.

primers

5 ′-ATGACCACCGAGCAGCGAC-3′

and

5 ′-CTA

platyrhynchos)

and

5′-CTA

TACGGACATTTCTGCAGGATCGA-3 ′

splenic

cDNA

using

the

primers

291

The viruses were amplified in chicken embryos. The titers were determined as TCID50 by

292

monitoring the cytopathic effect (CPE) of endpoint dilutions on MDCK cells in the presence of 1

293

μg/ml tosylsulfonyl phenylalanyl chloromethyl ketone (TPCK)-trypsin (Worthington Biochemical,

294

Lakewood, NJ, USA.). All experiments with virus were performed under biosafety level 2 (BSL-2)

295

conditions with investigators wearing appropriate protective equipment and complying with

296

general biosafety standards in the microbiological and biomedical laboratories of Ministry of

297

Health of the People’s Republic of China (WS 233-2002).

298

4.3 Cell culture, transfections, and infections.

299

DF-1, 293T, A549, and MDCK cells were maintained in Dulbecco’s modified Eagle’s 10

300

medium (DMEM), supplemented with 0.6 μg/ml penicillin, 60 μg/ml streptomycin and 10% fetal

301

bovine serum (FBS) (Gibco), in an atmosphere of 5% CO2 at 37°C. For infection of transfected

302

cells, 2.5 × 105 DF-1, 293T, A549, or MDCK cells were seeded into 24-well plates and

303

maintained in DMEM plus 10% FBS overnight. DF-1, 293T, and MDCK cells were transfected

304

with 800 ng of EGFP-tagged plasmids with Lipofectamine 2000 (Invitrogen). A549 cells were

305

transfected with EGFP-tagged plasmids using FuGENE® HD (Promega). Then, 24 h p.i., cells

306

were infected

307

μg/ml)TPCK-treated trypsin was used for infection at in view of the trypsin sensitivity of DF-1

308

and 293T cells, compared with 0.5 μg/ml in A549, and 2 μg/ml in MDCK cells (Lee et al., 2008).

309

4.4 Western blotting.

with SW731, WSN, or ZB07

viruses.

A lower concentration (0.1

310

293T and DF-1 cells were lysed in RIPA buffer containing 20 mM Tris base-HCl (pH 7.5),

311

150 mM NaCl, 1.0 mM EDTA (pH 8.0), 1.0 M megta (pH 8.0), 0.1% SDS, 0.5% DOC, 1% NP-40,

312

protease inhibitor cocktail (Roche), and phosphatase inhibitor cocktail (Roche). Proteins were

313

separated using SDS-PAGE under reducing conditions and analyzed using western blotting with

314

anti-EGFP (Beijing B&M Biotech Co., Ltd, Beijing, China) and anti-GAPDH (Beijing CoWin

315

Biotech, Beijing, China) antibodies followed by HRP-labeled secondary antibodies (BIO-RAD,

316

USA) (Chen et al., 2010).

317

4.5 Quantitative real-time PCR (qRT-PCR)

318

Total RNA was extracted using TRIzol (Invitrogen) and treated with DNase I (Promega) to

319

remove the contaminating DNA. Then, 1 μg of RNA was reverse-transcribed into cDNA using a

320

GoScript reverse transcription system (Promega) in a 20-μl reaction mixture. The cDNA was

321

analyzed using qRT-PCR with SYBR green Master I (Roche). The primers specific for chicken

322

GAPDH, IFN-β, Mx, and PKR have been described previously (Barber et al., 2010). The specific

323

primers for human GAPDH, IFN-β, Mx, and PKR were designed using Primer Express 3.0 and are

324

shown in Supplementary Table 2. qRT-PCR was performed under the following cycling

325

conditions: 95°C for 10 min for activation, followed by 40 cycles of 95°C for 15 sec and 60°C for

326

1 min. This was followed by one cycle of 95°C for 15 sec, 60°C for 15 sec, and 95°C for 15 sec.

327

The final step was to obtain a melting curve for the PCR products to determine the specificity of

328

amplification. All control and infected samples were analyzed in triplicate on the same plate. The

329

relative mRNA abundances were calculated using the 2 -∆∆Ct method with GAPDH as a reference 11

330

and plotted as fold change compared with mock-control samples (Livak and Schmittgen, 2001).

331

4.6 Measurement of virus growth in avian and mammalian cells.

332

DF-1, A549, and MDCK cells transfected with EGFP-tagged plasmids were infected with

333

SW731,WSN, and ZB07 at an MOI of 0.01, and cell cultures were collected at different times (24

334

and 48 h) after infection. The cell culture samples were centrifuged at 5000 × g for 1 min, and the

335

supernatants were stored at −80°C until use. The viral contents in the supernatants were titrated

336

using TCID50 in MDCK cells (Reed and Muench, 1938).

337

4.7 Statistical analysis.

338

All data analyses were performed using SPSS 16.0 software. Independent sample t-tests were

339

used to determine significant differences between different plasmid-transfected groups. P values ≤

340

0.05 were considered significant.

341 342

Acknowledgments

343

This work was supported by the National Basic Research Program (973 Program, number

344

2013CB945000). We thank Dr. Zhiyi Wan for critical reading of the manuscript. We thank LetPub

345

(www.letpub.com) for its linguistic assistance during the preparation of this manuscript.

346 347

References

348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365

Barber, M.R., Aldridge Jr, J.R., Fleming-Canepa, X., Wang, Y.-D., Webster, R.G., Magor, K.E., 2013. Identification of avian RIG-I responsive genes during influenza infection. Molecular immunology 54, 89-97. Barber, M.R., Aldridge, J.R., Webster, R.G., Magor, K.E., 2010. Association of RIG-I with innate immunity of ducks to influenza. Proceedings of the National Academy of Sciences 107, 5913-5918. Chen, C.-J., Chen, G.-W., Wang, C.-H., Huang, C.-H., Wang, Y.-C., Shih, S.-R., 2010. Differential localization and function of PB1-F2 derived from different strains of influenza A virus. Journal of virology 84, 10051-10062. Creagh, E.M., O’Neill, L.A., 2006. TLRs, NLRs and RLRs: a trinity of pathogen sensors that co-operate in innate immunity. Trends in immunology 27, 352-357. Gack, M.U., Kirchhofer, A., Shin, Y.C., Inn, K.-S., Liang, C., Cui, S., Myong, S., Ha, T., Hopfner, K.-P., Jung, J.U., 2008. Roles of RIG-I N-terminal tandem CARD and splice variant in TRIM25-mediated antiviral signal transduction. Proceedings of the National Academy of Sciences 105, 16743-16748. Gack, M.U., Nistal-Villán, E., Inn, K.-S., García-Sastre, A., Jung, J.U., 2010. Phosphorylation-mediated negative regulation of RIG-I antiviral activity. Journal of virology 84, 3220-3229. Gack, M.U., Shin, Y.C., Joo, C.-H., Urano, T., Liang, C., Sun, L., Takeuchi, O., Akira, S., Chen, Z., Inoue, S., 2007. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature 446, 916-920. 12

366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409

Gao, D., Yang, Y.-K., Wang, R.-P., Zhou, X., Diao, F.-C., Li, M.-D., Zhai, Z.-H., Jiang, Z.-F., Chen, D.-Y., 2009. REUL is a novel E3 ubiquitin ligase and stimulator of retinoic-acid-inducible gene-I. PLoS ONE 4, e5760. Hausmann, S., Marq, J.-B., Tapparel, C., Kolakofsky, D., Garcin, D., 2008. RIG-I and dsRNA-induced IFNβ activation. PLoS ONE 3, e3965. Honda, K., Yanai, H., Negishi, H., Asagiri, M., Sato, M., Mizutani, T., Shimada, N., Ohba, Y., Takaoka, A., Yoshida, N., 2005. IRF-7 is the master regulator of type-I interferon-dependent immune responses. Nature 434, 772-777. Jiang, X., Kinch, L.N., Brautigam, C.A., Chen, X., Du, F., Grishin, N.V., Chen, Z.J., 2012. Ubiquitin-induced oligomerization of the RNA sensors RIG-I and MDA5 activates antiviral innate immune response. Immunity 36, 959-973. Kawai, T., Akira, S., 2009. The roles of TLRs, RLRs and NLRs in pathogen recognition ARTICLE. International immunology 21, 317-337. Kawai, T., Takahashi, K., Sato, S., Coban, C., Kumar, H., Kato, H., Ishii, K.J., Takeuchi, O., Akira, S., 2005. IPS-1, an adaptor triggering RIG-I-and Mda5-mediated type I interferon induction. Nature immunology 6, 981-988. Kowalinski, E., Lunardi, T., McCarthy, A.A., Louber, J., Brunel, J., Grigorov, B., Gerlier, D., Cusack, S., 2011. Structural basis for the activation of innate immune pattern-recognition receptor RIG-I by viral RNA. Cell 147, 423-435. Lee, C.-W., Jung, K., Jadhao, S., Suarez, D., 2008. Evaluation of chicken-origin (DF-1) and quail-origin (QT-6) fibroblast cell lines for replication of avian influenza viruses. Journal of virological methods 153, 22-28. Livak, K.J., Schmittgen, T.D., 2001. Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the 2< sup>− ΔΔCT Method. methods 25, 402-408. Loo, Y.-M., Gale Jr, M., 2011. Immune signaling by RIG-I-like receptors. Immunity 34, 680-692. Magor, K.E., Miranzo Navarro, D., Barber, M.R., Petkau, K., Fleming-Canepa, X., Blyth, G.A., Blaine, A.H., 2013. Defense genes missing from the flight division. Developmental & Comparative Immunology 41, 377-388. Meylan, E., Curran, J., Hofmann, K., Moradpour, D., Binder, M., Bartenschlager, R., Tschopp, J., 2005. Cardif is an adaptor protein in the RIG-I antiviral pathway and is targeted by hepatitis C virus. Nature 437, 1167-1172. Miranzo-Navarro, D., Magor, K.E., 2014. Activation of Duck RIG-I by TRIM25 Is Independent of Anchored Ubiquitin. PLoS ONE 9, e86968. Nakhaei, P., Paz, S., Hiscott, J., 2006. Through Toll-like Receptor-dependent and-independent Pathways. The Interferons: Characterization and Application. Nistal-Villán, E., Gack, M.U., Martínez-Delgado, G., Maharaj, N.P., Inn, K.-S., Yang, H., Wang, R., Aggarwal, A.K., Jung, J.U., García-Sastre, A., 2010. Negative Role of RIG-I Serine 8 Phosphorylation in the Regulation of Interferon-β Production. Journal of Biological Chemistry 285, 20252-20261. Oshiumi, H., Matsumoto, M., Hatakeyama, S., Seya, T., 2009. Riplet/RNF135, a RING finger protein, ubiquitinates RIG-I to promote interferon-β induction during the early phase of viral infection. Journal of biological chemistry 284, 807-817. Oshiumi, H., Miyashita, M., Inoue, N., Okabe, M., Matsumoto, M., Seya, T., 2010. The ubiquitin ligase Riplet is essential for RIG-I-dependent innate immune responses to RNA virus infection. Cell host & microbe 8, 496-509. 13

410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437

Pichlmair, A., Schulz, O., Tan, C.P., Näslund, T.I., Liljeström, P., Weber, F., e Sousa, C.R., 2006. RIG-I-mediated antiviral responses to single-stranded RNA bearing 5'-phosphates. Science 314, 997-1001. Rajsbaum, R., Albrecht, R.A., Wang, M.K., Maharaj, N.P., Versteeg, G.A., Nistal-Villán, E., García-Sastre, A., Gack, M.U., 2012. Species-specific inhibition of RIG-I ubiquitination and IFN induction by the influenza A virus NS1 protein. PLoS pathogens 8, e1003059. Reed, L.J., Muench, H., 1938. A simple method of estimating fifty per cent endpoints. American Journal of Epidemiology 27, 493-497. Saito, T., Hirai, R., Loo, Y.-M., Owen, D., Johnson, C.L., Sinha, S.C., Akira, S., Fujita, T., Gale, M., 2007. Regulation of innate antiviral defenses through a shared repressor domain in RIG-I and LGP2. Proceedings of the National Academy of Sciences 104, 582-587. Sun, Y., Ding, N., Ding, S.S., Yu, S., Meng, C., Chen, H., Qiu, X., Zhang, S., Yu, Y., Zhan, Y., 2013. Goose RIG-I functions in innate immunity against Newcastle disease virus infections. Molecular immunology 53, 321-327. Takahasi, K., Yoneyama, M., Nishihori, T., Hirai, R., Kumeta, H., Narita, R., Gale Jr, M., Inagaki, F., Fujita, T., 2008. Nonself RNA-sensing mechanism of RIG-I helicase and activation of antiviral immune responses. Molecular cell 29, 428-440. Webster, R.G., Bean, W.J., Gorman, O.T., Chambers, T.M., Kawaoka, Y., 1992. Evolution and ecology of influenza A viruses. Microbiological reviews 56, 152-179. Xu, L.-G., Wang, Y.-Y., Han, K.-J., Li, L.-Y., Zhai, Z., Shu, H.-B., 2005. VISA is an adapter protein required for virus-triggered IFN-β signaling. Molecular cell 19, 727-740. Yoneyama, M., Kikuchi, M., Natsukawa, T., Shinobu, N., Imaizumi, T., Miyagishi, M., Taira, K., Akira, S., Fujita, T., 2004. The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nature immunology 5, 730-737. Zeng, W., Sun, L., Jiang, X., Chen, X., Hou, F., Adhikari, A., Xu, M., Chen, Z.J., 2010. Reconstitution of the RIG-I pathway reveals a signaling role of unanchored polyubiquitin chains in innate immunity. Cell 141, 315-330.

438 439

Figure legends

440

Figure 1. Expression levels of RIG-Is and CARDs in DF-1 and 293T cells.

441

(A) DF-1 cells were transfected with EGFP-tagged RIG-I and CARDs expression plasmids or

442

empty vector alone. After 24 h, the expression of all the RIG-Is and CARDs was confirmed using

443

fluorescence microscopy. (a) pEGFP-N1; (b) EGFP-hCARDs; (c) EGFP-mCARDs; (d)

444

EGFP-dCARDs; (e) EGFP-gCARDs; (f) EGFP-hRIG-I; (g) EGFP-mRIG-I; (h) EGFP-dRIG-I,

445

and (i) EGFP-gRIG-I. Scale bars = 100 μm.

446

(B) 293T cells were transfected with EGFP-tagged RIG-I and CARDs expression plasmids or

447

empty vector alone. After 24 h, the expression of RIG-Is and CARDs was confirmed using

448

fluorescence microscopy. (a) pEGFP-N1; (b) EGFP-hCARDs; (c) EGFP-mCARDs; (d) 14

449

EGFP-dCARDs; (e) EGFP-gCARDs; (f) EGFP-hRIG-I; (g) EGFP-mRIG-I; (h) EGFP-dRIG-I,

450

and (i) EGFP-gRIG-I. Scale bars = 100 μm.

451

(C) DF-1 cells were transfected with EGFP-tagged RIG-I and CARD expression plasmids or

452

empty vector alone. After 24 h, cell lysates were separated by SDS-PAGE and probed with

453

anti-EGFP and anti-GAPDH antibodies. Results are representative of two independent

454

experiments.

455

(D) 293T cells were transfected with EGFP-tagged RIG-I and CARD expression plasmids or

456

empty vector alone. After 24 h, cell lysates were separated by SDS-PAGE and probed with

457

anti-EGFP and anti-GAPDH antibodies. Results are representative of two independent

458

experiments.

459

The relative fleorescence intensity was analysed with Image-pro-plus. Data represent mean ± SEM

460

from three wells per group. *P ≤ 0.05 vs. dCARDs, #P ≤ 0.05 vs. dRIG-I, $P ≤ 0.05 vs. gRIG-I.

461

Results are representative of two independent experiments.

462 463

Figure 2. Waterfowl RIG-Is and CARDs induce the innate immune genes more strongly than do

464

mammalian RIG-Is and CARDs in DF-1 cells during IAV infection.

465

(A, B) RIG-I/CARDs-transfected DF-1 cells responded to IAV infection (MOI, 1) by increasing

466

the expression of chicken IFN-β relative to empty vector-transfected and mock treated cells. DF-1

467

cells were transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all

468

plasmid-transfected DF-1 cells were infected with SW731, WSN, or ZB07 viruses, or mock

469

treated. RNA was extracted from cells for qRT-PCR at 8 h p.i..

470

(C, D) RIG-I/CARDs-transfected DF-1 cells respond to IAV infection (MOI, 1) by increasing the

471

expression of chicken Mx relative to empty vector-transfected and mock-treated cells. DF-1 cells

472

were transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all

473

plasmid-transfected DF-1 cells were infected with SW731, WSN, or ZB07 viruses, or mock

474

treated. RNA was extracted from cells for qRT-PCR at 8 h p.i..

475

(E, F) RIG-I/CARDs-transfected DF-1 cells respond to IAV infection (MOI, 1) by increasing the

476

expression of chicken PKR relative to empty vector-transfected and mock-treated cells. DF-1 cells

477

were transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all

478

plasmid-transfected DF-1 cells were infected with SW731, WSN, or ZB07 viruses, or mock 15

479

treated. RNA was extracted from cells for qRT-PCR at 8 h p.i..

480

Data represent mean ± SEM from three wells per group. *P ≤ 0.05 vs. vector; **P ≤ 0.005 vs.

481

vector. Results are representative of two independent experiments.

482 483

Figure 3. Mammalian RIG-Is and CARDs induced type I IFN more strongly than waterfowl in

484

293T cells during IAV infection.

485

(A, B) RIG-I/CARD-transfected 293T cells responded to IAV infection (MOI, 1) with increased

486

expression of IFN-β relative to empty vector-transfected and mock-treated cells. 293T cells were

487

transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all plasmid-transfected

488

293T cells were infected with SW731, WSN, or ZB07 viruses, or mock treated. RNA was

489

extracted from cells for qRT-PCR at 8 h p.i..

490

Data represent mean ± SEM from three wells per group. αP ≤ 0.05 vs. mCARDs, βP ≤ 0.05 vs.

491

dCARDs, γP ≤ 0.05 vs. gCARDs, δP ≤ 0.05 vs. vector, aP ≤ 0.05 vs. mCARDs, bP ≤ 0.05 vs.

492

dCARDs, cP ≤ 0.05 vs. gCARDs, d P ≤ 0.05 vs. vector. Results are representative of two

493

independent experiments.

494

Figure 4. Mammalian RIG-Is and CARDs induced type I IFN more strongly than waterfowl in

495

A549 cells during IAV infection.

496

(A, B) RIG-I/CARDs-transfected A549 cells responded to IAV infection (MOI, 1) with increased

497

expression of IFN-β relative to empty vector-transfected and mock treated cells. A549 cells were

498

transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all plasmid-transfected

499

A549 cells were infected with SW731, WSN, or ZB07 viruses, or mock treated. RNA was

500

extracted from cells for qRT-PCR at 8 h p.i..

501

(C, D) RIG-I/CARDs-transfected A549 cells respond to IAV infection (MOI, 1) with increased

502

expression of Mx relative to empty vector-transfected and mock-treated cells. A549 cells were

503

transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all plasmid-transfected

504

A549 cells were infected with SW731, WSN, or ZB07 viruses, or mock treated. RNA was

505

extracted from cells for qRT-PCR at 8 h p.i..

506

(E, F) RIG-I/CARDs-transfected A549 cells respond to IAV infection (MOI, 1) with increased

507

expression of PKR relative to empty vector-transfected and mock-treated cells. A549 cells were

508

transfected with EGFP-tagged plasmids or empty vector. Then, 24 h later, all plasmid-transfected 16

509

A549 cells were infected with SW731, WSN, or ZB07 viruses, or mock treated. RNA was

510

extracted from cells for qRT-PCR at 8 h p.i..

511

Data represent mean ± SEM from three wells per group. αP ≤ 0.05 vs. mCARDs, βP ≤ 0.05 vs.

512

dCARDs, γP ≤ 0.05 vs. gCARDs, δP ≤ 0.05 vs. vector, aP ≤ 0.05 vs. mCARDs, bP ≤ 0.05 vs.

513

dCARDs, cP ≤ 0.05 vs. gCARDs, d P ≤ 0.05 vs. vector. Results are representative of two

514

independent experiments.

515 516

Figure 5. Waterfowl RIG-Is and CARDs reduce IAV replication more efficiently than mammalian

517

RIG-Is in DF-1 cells.

518

(A, B) DF-1 cells were transfected with EGFP-tagged plasmids or control vector, and, 24 h later,

519

infected with SW731 at an MOI of 0.01. At 24 and 48 h p.i., the viral titers in the cell cultures

520

were determined as TCID 50.

521

(C, D) DF-1 cells were transfected with EGFP-tagged plasmids or control vector. Then, 24 h later,

522

they were infected with WSN at an MOI of 0.01. At 24 and 48 h p.i., the viral titers in the cell

523

cultures were determined as TCID 50.

524

(E, F) DF-1 cells were transfected with EGFP-tagged plasmids or control vector. Then, 24 h later,

525

they were infected with ZB07 at an MOI of 0.01. At 24 and 48 h p.i., the viral titers in the cell

526

cultures were determined as TCID 50.

527

Data represent mean ± SEM from three wells per group. *P ≤ 0.05 vs. vector, **P ≤ 0.005 vs.

528

vector. Results are representative of two independent experiments.

529 530

Figure 6. Mammalian RIG-I interfere with replication of IAV more efficiently than waterfowl

531

RIG-I in A549 cells.

532

(A, B) A549 cells were transfected with EGFP-tagged plasmids or control vector. Then, 24 h later,

533

they were infected with SW731 at an MOI of 0.01. At 24 and 48 h p.i., the viral titers in the cell

534

cultures were determined as TCID50.

535

(C, D) A549 cells were transfected with EGFP-tagged plasmids or control vector. Then, 24 h later,

536

they were infected with WSN at an MOI of 0.01. At 24 and 48 h p.i., the viral titers in the cell

537

cultures were determined as TCID50.

538

(E, F) A549 cells were transfected with EGFP-tagged plasmids or control vector. Then, 24 h later, 17

539

they were infected with ZB07 at an MOI of 0.01. At 24 and 48 h p.i., the viral titers in the cell

540

cultures were determined as TCID50.

541

Results were analyzed with the independent sample t test (n=3). Data represent mean ± SEM from

542

three wells per group. *P ≤ 0.05 vs. vector, **P ≤ 0.005 vs. vector, #P ≤ 0.05 vs. dCARDs, $P ≤

543

0.05 vs. dRIG-I. Results are representative of two independent experiments.

544

18

RIG-I from waterfowl and mammals differ in their abilities to induce antiviral responses against influenza A viruses.

The retinoic acid-induced gene I (RIG-I) plays a crucial role in sensing viral RNA and IFN-β production. RIG-I varies in length and sequence between d...
18MB Sizes 0 Downloads 5 Views