JVI Accepts, published online ahead of print on 14 May 2014 J. Virol. doi:10.1128/JVI.03178-13 Copyright © 2014, American Society for Microbiology. All Rights Reserved.

1

Single domain antibodies targeting neuraminidase protect against an H5N1 influenza

2

virus challenge

3

Francisco Miguel Cardoso1,2 , Lorena Itatí Ibañez1,2$, Silvie Van den Hoecke1,2, Sarah De

4

Baets1,2, Anouk Smet1,2, Kenny Roose1,2, Bert Schepens1,2, Francis J. Descamps1,2#, Walter

5

Fiers1,2, Serge Muyldermans3,6, Ann Depicker4,5, Xavier Saelens1,2

6

1

7

2

8

Belgium

9

3

10

4

11

5

12

6

13

Brussels, Belgium

14

$

15

#

VIB Inflammation Research Center, Technologiepark 927, Ghent, Belgium Department for Biomedical Molecular Biology, Ghent University, Technologiepark 927, Ghent,

Structural Biology Research Center, VIB, Pleinlaan 2, B-1050 Brussels, Belgium Department of Plant Systems Biology, VIB, Technologiepark 927, Ghent, Belgium Department of Biotechnology and Bioinformatics, Technologiepark 927, Ghent, Belgium Laboratory of Cellular and Molecular Immunology, Vrije Universiteit Brussel, Pleinlaan 2, 1050

Current address: ICT-Milstein – CONICET, Ciudad Autónoma de Buenos Aires, Argentina Current address: Ablynx NV, Ghent, Belgium

16 17

Address correspondence to: Ann Depicker, [email protected], or Xavier Saelens,

18

[email protected]

19

Running Title: NA single domain antibodies protect against H5N1

20 1

21

Word Count for abstract: 220

22

Word Count for text: 9790

2

23

ABSTRACT

24

Influenza virus neuraminidase (NA) is an interesting target of small-molecule antiviral drugs. We

25

isolated a set of H5N1 NA-specific single domain antibodies (N1-VHHm) and evaluated their in

26

vitro and in vivo antiviral potential. Two of them inhibited NA activity and in vitro replication of

27

clade 1 and 2 H5N1 viruses. We then generated bivalent derivatives of N1-VHHm by two

28

methods. First we made N1-VHHb by genetically joining two N1-VHHm moieties by a flexible

29

linker. Second, bivalent N1-VHH-Fc proteins were obtained by genetic fusion of the N1-VHHm

30

moiety the crystallizable region of mouse IgG2a (Fc). The in vitro antiviral potency against H5N1

31

of both bivalent N1-VHHb formats was 30- to 240-fold higher than their monovalent

32

counterparts, with IC50 values in the low nanomolar range. Moreover, single-dose prophylactic

33

treatment with bivalent N1-VHHb or N1-VHH-Fc protected BALB/c mice against a lethal

34

challenge with H5N1 virus, including an Oseltamivir-resistant H5N1 variant. Surprisingly, also an

35

N1-VHH-Fc fusion without in vitro NA inhibitory or antiviral activity protected mice against H5N1

36

challenge. Virus escape selection experiments indicated that one amino acid residue close to

37

the catalytic site is required for N1-VHHm binding. We conclude that single domain antibodies

38

directed against influenza NA protect against H5N1 virus infection, and when engineered with a

39

conventional Fc domain, they can do so in the absence of detectable NA inhibitory activity.

40

IMPORTANCE

41

Highly pathogenic H5N1 viruses are a zoonotic threat. Outbreaks of bird flu caused by these

42

viruses occur in many parts of the world, are associated with tremendous economic loss and

43

these viruses can cause very severe disease in human. In such cases, small molecule inhibitors

44

of the viral neuraminidase are among the few treatment options for patients. However treatment

45

with such drugs often results in the emergence of resistant viruses. Here we show that single

46

domain antibody fragments that are specific for NA can bind and inhibit H5N1 viruses in vitro

3

47

and can protect laboratory mice against challenge with an H5N1 virus, including an oseltamivir-

48

resistant virus. In addition, plant-produced VHH fused to a conventional Fc domain, can protect

49

in vivo even in the absence of NA inhibitory activity. Thus, NA of influenza virus can be

50

effectively targeted by single domain antibody fragments, which are amenable to further

51

engineering.

4

52

INTRODUCTION

53

Zoonotic influenza A virus infections are a persistent threat due to their pandemic potential. In

54

particular, highly pathogenic avian influenza viruses (HPAIV) of the H5N1, H7N1 and H7N7

55

subtypes occasionally cross the species barrier between domesticated birds and human. These

56

viruses could become transmissible between humans through reassortment with circulating

57

swine or human influenza viruses or by gradually accumulating mutations (1, 2). In the last

58

decade, zoonotic outbreaks had a major effect on public health. HPAIV H5N1 (3), the swine flu

59

(H1N1) outbreak in 2009 (4), and more recently, human infections with H7N9 in South Asia (5)

60

illustrate our poor preparedness for pandemic influenza (6). HPAIV H5N1 infection in humans

61

has a confirmed case fatality rate of approximately 60%. The high pathogenicity of HPAIV H5N1

62

in humans can be attributed to a high replication rate and a broad cellular tropism that can lead

63

to systemic virus spread. In addition, deregulated induction of proinflammatory cytokines and

64

chemokines (“cytokine storm”) is associated with severe HPAIV H5N1 infections, and can result

65

in a disproportionate immunological response (7).

66

Influenza neuraminidase (NA) is a homotetrameric type II membrane glycoprotein with sialidase

67

activity. The NA catalytic site is located at the top of each monomer, opposite to the tetramer

68

interface. NA plays an essential role in the spread of influenza viruses by cleaving sialic acids

69

from the host cell receptors and from virions. NA activity also contributes to virus entry by

70

cleaving decoy receptors present in mucins that line the layer of respiratory epithelial cells (8).

71

Immunologically, NA is the second major humoral antigenic determinant (after HA) and is

72

subject to antigenic drift and occasional shift. In addition, experimental influenza vaccines

73

supplemented with NA have improved efficacy (9-11). NA is also a co-determinant of influenza

74

A virus (IAV) pathogenicity (12-14) and is involved in limiting IAV superinfections and

75

reassortment (15). Decreased NA activity has been correlated with H5N1 adaptation to human

5

76

airway epithelium (16) and antibodies against NA contribute to protection against H5N1 virus

77

challenge in a mouse model (17).

78

Hemagglutinin (HA), the other major antigen, and NA cooperate in a tightly controlled way. For

79

example, the fitness of mutant IAV virus lacking NA activity can be rescued by selection of HA

80

mutants with decreased affinity for receptors containing sialic acid (18-20). These data

81

demonstrate the importance of NA during IAV infection, so targeting NA is a rational strategy.

82

Indeed, three licensed influenza antivirals, Oseltamivir, Zanamivir and Peramivir, target NA.

83

Influenza viruses that are resistant to Oseltamivir frequently emerge in humans. In addition, NA-

84

specific antibodies protect mice, and serum anti-NA antibodies are associated with resistance to

85

influenza A virus challenge in humans (21).

86

Effective prevention and treatment strategies are needed to control H5N1 infections and single-

87

domain antibody-based technology have been described as promising antivirals (22). Naturally

88

occurring single heavy chain antibodies have been found in sharks and camelids (23, 24).

89

Unlike conventional antibodies that are typically composed of a light and a heavy polypeptide

90

chain that together determine the epitope specificity, these natural single-chain antibodies are

91

composed of heavy chains only, and hence their epitope specificity is confined in a single

92

domain (so called VHH, Variable domain of Heavy chain). Single-domain antibody fragments

93

derived from camelids are highly specific and robust, easy to produce and flexible, and hence

94

they are used as novel therapeutics for many diseases (25, 26). VHHs usually have extended

95

antibody loop structures and can interact with cognate antigens with affinities matching the best

96

affinities of conventional antibodies (27-29). In addition, VHHs that potently inhibit enzymes

97

have been described (30, 31), and they have been used to stabilize membrane proteins in order

98

to determine crystal structure (32). The VHH tends to target concave-shaped epitopes, in

99

contrast to the planar epitopes that are typically recognized by “classical” light and heavy chain

100

antibodies (33). In addition, several examples of VHHs with potent antiviral activity have been

6

101

reported (34-36). We previously reported on an HA-specific H5N1 neutralizing VHH (35) and

102

H5N2 neutralizing VHH were reported more recently (37). Moreover, a VHH directed against the

103

influenza M2 protein inhibited replication of amantadine-sensitive and amantadine-resistant

104

viruses and protected mice against lethal influenza challenge (38). Recently, Harmsen et al.

105

described a panel of NA-binding VHHs, some of which displayed cross-subtype NA-binding, but

106

the antiviral potential of these VHHs was not evaluated (39).

107

Here, we report the isolation and characterization of a set of H5N1 NA-binding VHHs (N1-

108

VHHm) of which some have potent NA inhibitory activity and in vitro and in vivo antiviral

109

activities. We also describe the production of bivalent in tandem N1-VHHm formats in

110

Escherichia coli and N1-VHHm proteins fused to a mouse IgG2a Fc domain in seeds of

111

transgenic Arabidopsis plants. Further, we compared the in vitro antiviral activity against H5N1

112

clade 1 (Oseltamivir-sensitive and -resistant strains) and clade 2 viruses, and we evaluated the

113

N1-VHHm based prophylactic protection of mice against a potentially lethal infection with H5N1

114

virus.

7

115

MATERIALS AND METHODS

116

Influenza viruses

117

H5N1 IAV strains NIBRG-14 and NIBRG-23 were obtained from the UK National Institute for

118

Biological Standards and Control, a center of the Health Protection Agency. NIBRG-14 and

119

NIBRG-23 are 2:6 reverse genetics–derived reassortants with NA and HA (lacking the polybasic

120

cleavage site) segments derived from A/Vietnam/1194/2004 (H5N1) and A\turkey\Turkey\2005

121

(H5N1), respectively, and the other six segments from A/PR/8/34 (H1N1) viruses. The H5N1

122

H274Y virus described here is a 1:1:6 reverse genetics-derived reassortant, with NA derived

123

from A/crested eagle/Belgium/01/2004 (40) carrying the H274Y mutation introduced by site-

124

specific mutagenesis, HA from NIBRG-14, and the remaining six genome segments from

125

A/PR/8/34 (41). This virus was rescued by transfection of co-cultured HEK-293T and MDCK

126

cells. The supernatant from these cells was used for end point dilution to obtain a clonal H5N1

127

H274Y virus sample that was subsequently amplified on MDCK cells, pelleted from the cell

128

supernatant, and mouse adapted by serial passage in BALB/c mice. All HA segments of these

129

H5N1 viruses lack the coding information for the polybasic cleavage site. Following adaptation

130

to BALB/c mice, the HA and NA-coding regions of the mouse-adapted NIBRG-14 (NIBRG-14

131

ma) and H5N1 H274Y (H5N1 H274Y ma) were sequenced and found to be identical to those of

132

the parental viruses. pH1N1 (kindly provided by Dr. Bernard Brochier, Scientific Institute of

133

Public Health, Brussels, Belgium) is derived from a clinical isolate of the pH1N1 virus of 2009

134

and was adapted to mice by serial passages (42). A/New Caledonia/20/99 was kindly provided

135

by Dr. Alan Hay (MRC National Institute for Medical Research, Mill Hill, London, UK). The

136

median tissue culture infectious dose (TCID50) and median lethal dose (LD50) of NIBRG-14 ma

137

and H5N1 H274Y ma viruses were calculated by the method of Reed and Muench (43). All the

138

H5N1 and pH1N1 experiments described above were performed in BSL-2+ rooms.

8

139

Baculovirus-based production of recombinant soluble tetrameric neuraminidase

140

A soluble and enzymatically active tetrameric NA (N1rec) derived from A/crested

141

eagle/Belgium/01/2004 (H5N1) was produced by proprietary technology (WO 2002/074795).

142

Briefly, the N1rec expression cassette consisted of the secretion signal of IAV HA (ssHA, 16

143

residues), a tetramerizing variant of the leucine zipper domain of the Saccharomyces cerevisiae

144

GCN4 transcription factor (tGCN4, 32 residues) (44, 45), and the extracellular part of NA

145

derived from A/crested eagle/Belgium/01/2004 (H5N1) (amino acid residues 53–449) (40) (Fig.

146

1A). This N1rec expression cassette was cloned into the pAcMP2 baculovirus transfer vector to

147

produce pAcMP2n1rec. Recombinant baculovirus was generated by co-transfection of

148

pAcMP2n1rec and linearized DNA of Autografa californica Nuclear Polyhedrosis Virus (AcNPV)

149

(BaculoGold, BD Biosciences) into Sf9 insect cells (derived from Spodoptera frugiperda), using

150

the ESCORT IV Transfection Reagent (Sigma-Aldrich) to generate N1rec encoding recombinant

151

baculovirus. Sf9 cells infected with AcNPVN1rec were incubated at 28°C in rolling bottles in TC-

152

100 medium (Invitrogen), supplemented with 10% FCS and penicillin + streptomycin. Four days

153

after infection, the supernatant was harvested and clarified by centrifugation (1 h at 50,000 g).

154

Recombinant neuraminidase purification

155

All purification steps were carried out at 4°C. Two parts (v:v) of n-butanol were added to three

156

parts of cleared AcNPVN1rec-infected Sf9 cell supernatant and the mixture was shaken

157

vigorously for 5 min. The aqueous phase, containing soluble N1rec, was separated from the

158

organic phase, diluted 2.5 times in 5 mM KH2PO4 pH 6.6 and passed through a 0.22-µm filter.

159

The diluted aqueous phase was subsequently applied to an XK26\70 column (GE Healthcare)

160

packed with HA Ultrogel Hydroxyapatite Chromatography Sorbent (Pall), and eluted with a

161

gradient of 5-400 mM KH2PO4 pH 6.6 in 4% butanol. Eluted fractions with high NA activity were

162

pooled and loaded onto a Blue Sepharose (Sigma-Aldrich) column, equilibrated with 50 mM

9

163

MES pH 6.6 containing 5% glycerol and 8 mM CaCl2. Bound proteins were eluted with 50 mM

164

MES pH 6.6 containing 5% glycerol, 8 mM CaCl2 and 1.5 M NaCl. Fractions with NA activity

165

were loaded on a HiLoad 16/60 Superdex 200 (GE Healthcare) gel filtration column pre-

166

equilibrated with 50 mM MES pH 6.6 with 5 % glycerol and 8 mM CaCl2, 150 mM NaCl. Peak

167

fractions containing NA activity were pooled, aliquoted, and stored at -80ºC until used. All the

168

chromatography steps were performed using an Akta purification station (GE Healthcare).

169

Neuraminidase activity assays

170

NA activity was quantified by measuring the rate of cleavage of the fluorogenic substrate 4-

171

MUNANA (2’-4-Methylumbelliferyl-α-D-N-acetylneuraminic acid, sodium salt hydrate (Sigma-

172

Aldrich) into 4-methylumbelliferone. The NA activity reaction was performed in 200 mM NaAc, 2

173

mM CaCl2 with 1 % butanol and 1 mM 4-MUNANA and measured in a kinetic mode, with

174

excitation at 365 nm and emission at 450 nm in an Optima Fluorostar. A standard curve of

175

increasing concentrations of soluble 4-methylumbelliferone was included to correlate the

176

fluorescence intensity with the molar amount of 4-methylumbelliferone. One NA activity unit is

177

defined as the activity needed to generate 1 nmol of 4-methylumbelliferone per min.

178

Fetuin (5 µg/ml; Sigma-Aldrich) was coated on Nunc 96 well plates overnight at 4°C. Excess

179

fetuin was washed away with PBS, and IAV dilutions (diluted in PBS with 1 mM CaCl2 and 0.5

180

mM MgCl2), with or without added single domain antibodies, were added and incubated 1 h at

181

37°C. The amount of desialylated fetuin was measured by colorimetry to determine binding of

182

horseradish peroxidase (HRP) coupled peanut agglutinin (PNA, Sigma-Aldrich). The plates

183

were washed three times with PBS + 0.1% Tween 20 and then incubated with 50 µl PNA-HRP

184

(2.5 µg/ml in PBS + 0.05% Tween 20) at room temperature. Then the plates were washed three

185

times with PBS, after which 50 µl of TMB substrate (Pharmigen BD) was added and absorbance

186

was measured at 450 nm with a reference at 650 nm. In the Accelerated Viral Inhibition Assay

187

(AVINA) (46), IAV dilutions containing the indicated N1-VHHm concentrations were transferred

10

188

to a black 96-well plate and 75 μl of 20 μM MUNANA was added and incubated 1 h at 37°C.

189

Then, 100 μl stop solution (0.1 M glycine, pH 10.7, 25% ethanol) was added to each well, and

190

fluorescence was determined. For the AVINA and fetuin substrate assays, we used the

191

following amounts of IAV: 7 x 105 pfu of NIBRG-14 ma, 1 x 104 pfu of H5N1 H274Y ma, and 3 x

192

104 plaque forming units (pfu) of pH1N1.

193

Camelid immunization, phage library construction, and panning

194

An alpaca (Vicugna pacos) was injected subcutaneously with 125 µg of N1rec in the presence

195

of incomplete Freund’s adjuvant. Injections were once a week for five weeks. On day 39, anti-

196

coagulated blood was collected, lymphocytes were isolated using a UNI-SEP density gradient

197

separation kit (NOVAmed), and total RNA was extracted. cDNA was prepared using oligo (dT)

198

primers,

199

(GTCCTGGCTGCTCTTCTACAAGG)

200

(GGTACGTGCTGTTGAACTGTTCC). PstI and NotI restriction sites were inserted into the

201

amplified sequences using the primers A6E (GATGTGCAGCTGCAGGAGTCTGGRGGAGG)

202

and 38 (GGACTAGTGCGGCCGCTGGAGACGGTGACCTGGGT). The resulting PCR product

203

(550 bp) and the pHEN4 vector were digested with PstI and NotI, ligated (47), and transformed

204

into electrocompetent TG1 E. coli cells. Transformants were grown in 2xTY medium,

205

(supplemented with 100 µg/ml ampicillin and 1% glucose) and infected with helper phage

206

M13K07. The resulting VHH phage display library was panned four times using solid-phase

207

coated N1rec.

208

Production and purification of monovalent neuraminidase-binding VHH (N1-VHHm)

209

The VHH coding sequence of the selected N1rec-binding phages (n = 13) was amplified by

210

PCR with the primers A6E and 38. The N1rec-binding VHH genes obtained by PCR and the

and

the

VH

and

VHH

genes

were and

amplified

with

primer

primer

CALL001 CALL002

11

211

pHEN6c vector were digested with PstI and BstEII, ligated, and transformed into E. coli WK6

212

strain. In the pHEN6c vector, the coding sequence for the N1rec-binding VHH candidates were

213

cloned downstream of an N-terminal PelB secretion signal sequence and a carboxyterminal

214

hexahistidine tag was introduced (Fig. 1B)(48). For production and purification of soluble VHH,

215

pHEN6c-transformed WK6 cells were grown in TB medium supplemented with 100 µg/ml

216

ampicillin, 2 mM CaCl2 and 0.1% glucose. After 16 hours, VHH production was induced with 1

217

mM of isopropyl β-D-1-thiogalactopyranoside, and the periplasmic fraction was extracted by

218

osmotic shock using 0.2 M Tris pH 8.0, 0.5 mM EDTA and 0.5 M sucrose. Periplasmic extracts

219

were centrifuged at 8000 rpm and 4°C and the supernatant was applied to a His Select Nickel

220

Affinity gel (Sigma-Aldrich). After loading, the column was washed with PBS, and VHHs were

221

eluted with 0.5 M imidazole and dialyzed against PBS at 4°C. Monovalent N1rec-binding VHH

222

(N1-VHHm) was concentrated with a Vivaspin 5000-kDa cutoff filter (Vivascience). Total protein

223

concentration was determined with the BCA protein Assay kit (Thermo Scientific) and endotoxin

224

levels with a commercial Limulus Amebocyte Lysate-based assay (Toxinsensor Chromogenic

225

LAL Endotoxin Assay Kit, Genscript Corp.). We obtained 13 recombinant N1 NA-specific VHH

226

proteins, which were named N1-1-VHHm to N1-13-VHHm. An irrelevant monovalent VHH

227

against Arabidopsis thaliana albumin (SA-VHHm) was used as a negative control (49). The N1-

228

VHHm sequences were aligned using the MegAlign software (DNAstar) and the VHH CDR3

229

electronegativity surface potential was predicted with the open source software ESyPred3D and

230

visualized with DeepView (50)(Fig. 1C and 1D).

231

ELISA assays

232

N1rec or M2e-tGCN4 (44) at 10 µg/ml was used to coat O/N in 96-well Maxisorp (Nunc) plates

233

using 100 µl per well. Blocking was performed overnight with 5% skim milk in PBS. Dilution

234

series of the 13 N1-VHHm candidates (from 10 µM - 0.1 nM) were incubated for 2 h at 37°C,

12

235

and excess VHH was removed by three washes with PBS containing 0.05 % Tween 20. Binding

236

of the N1-VHHm to N1rec tGCN4 or NA globular head moieties was assessed by a mouse anti-

237

His-tag antibody fused with horse radish peroxidase (HRP) (1:5000). After three washes with

238

PBS, 100 µl of Tetramethylbenzidine (TMB) substrate (Pharmigen BD) was added, and

239

absorbance was measured at 450 nm.

240

N1-VHHm surface plasmon resonance analysis

241

The affinity of monovalent N1-3-VHHm, N1-5-VHHm and N1-7-VHHm for N1rec was

242

determined by surface plasmon resonance (SPR) on a Biacore 3000 platform (GE Healthcare).

243

N1rec antigen was first immobilized (2000 resonance units) into a Series S sensor chip CM5

244

(GE Healthcare) by chemical -NH2 coupling, in 10 mM NaAc pH 4.5. N1rec loaded channels

245

were regenerated with 0.02% SDS. Each monovalent N1-VHHm was diluted in HBS buffer (0.01

246

M HEPES, 0.15 M NaCl, 0.005% Tween 20, pH 6.4) to obtain a gradient ranging from 1.95 to

247

1000 nM. The different N1-VHHs were injected over the N1rec coated chip to record their

248

binding kinetics. The Biacore T00 Evaluation Software was used to calculate the association

249

rate constants (Kon), dissociation rate constants (Koff) and the equilibrium dissociation constants

250

(KD = Koff/Kon).

251

Construction of bacteria-produced bivalent N1-VHH (N1-VHHb)

252

The coding information for N1-3-VHHm and N1-5-VHHm was amplified with primers MH

253

(CATGCCATGGGAGCTTTGGGAGCTTTGGAGCTGGGGGTCTTCGCTGTGGTGCGCTGAGG

254

AGACGGTGACCTGGGT)

255

(CATGCCATGATCCGCGGCCCAGCCGGCCATGGCTGATGTGCAGCTGGTGGAGTCT)

256

introduce an NcoI restriction enzyme site at both ends of the amplified fragment. The MH primer

257

also introduced the hinge sequence of llama γ2c (AHHSEDPSSKAPKAPMA) (51) to separate

258

the two N1-VHHm domains. PCR amplicons of N1-3-VHHm and N1-5-VHHm were cloned into

and

A4short to

13

259

pHEN6c plasmid containing the n1-3-vhh or the n1-5-vhh gene in order to generate homo-

260

bivalent N1-3-VHHb and N1-5-VHHb expression constructs, which also contained a carboxy-

261

terminal hexahistidine tag (Fig. 1B). These bivalent constructs were expressed and purified from

262

the periplasmic fraction of E. coli as outlined above. We used an irrelevant bivalent in tandem

263

fused VHH specific for bacterial beta lactamase (BL-VHHb) as a negative control (52) (kindly

264

provided by Dr. Trong Nguyen-Duc, VUB, Brussels). Endotoxin levels of the bivalent VHH

265

preparations were determined as above.

266

Production of bivalent N1-VHHb fused to the mouse IgG2 Fc fragment (N1-VHH-Fc) in

267

seeds of stably transformed Arabidopsis thaliana lines

268

The three different n1-vhh-fc sequences coding for N1-3VHHm, N15-VHHm and N1-7-VHHm

269

fused at the aminoterminal region with the 2S2 signal peptide and at the carboxyterminal region

270

with the Fc region of mouse IgG2a and the KDEL retention signal were synthesized

271

commercially (GenScript), and cloned into a PphasGW binary T-DNA vector, resulting in the

272

PphasGWn1-vhh-fc vector (Fig. 1B). The Coronavirus GP4 subunit protein fused to the same

273

mouse IgG2a Fc moiety (GP4-Fc, kindly provided by Robin Piron, VIB and Ghent University)

274

was used as an irrelevant control in the antiviral assays. Agrobacterium C58C1RifR (pMP90)

275

was transformed with PphasGWn1-3-vhh-fc, PphasGWn1-5-vhh-fc, PphasGWn1-7-vhh-fc or

276

PphasGWGP4-fc constructs.

277

Agrobacterium

278

PphasGWn1-5-vhh-fc, PphasGWn1-7-vhh-fc and PphasGWGP4-fc binary T-DNA vectors and

279

the resulting strains were used for Arabidopsis thaliana floral dip transformation (53, 54). T1 to

280

T3 generations were produced and screened for best N1-VHH-Fc expressers. For protein

281

purification, 1 g of transformed A. thaliana seeds was crushed in 25 ml of 50 mM Tris-HCl, pH

282

8.0, 200 mM NaCl, 5 mM EDTA, 0.1% (v/v) Tween 20 and Complete protease inhibitor tablets

C58C1RifR (pMP90)

was

transformed

with

the

PphasGWn1-3-vhh-fc,

14

283

(Roche). The seed extracts were clarified by centrifugation and applied to a protein G

284

sepharose column (GE Healthcare). N1-VHH-Fc proteins were eluted in 0.1 M glycine pH 3.0

285

and immediately neutralized with 1 M Tris-HCl, pH 9.0. The recovery was between 1 and 10 mg

286

of N1- VHH-Fc. Endotoxin levels of the VHH-Fc preparations were determined as above.

287

Plaque assays

288

Monolayers of MDCK cells were grown in DMEM supplemented with 10% fetal calf serum, 1%

289

penicillin + streptomycin and 1% glutamine and non-essential amino acids at 37°C in 5% CO2.

290

MDCK TMPRSS2 cells (kindly provided by Dr. Wolfgang Garten, Philipps-Universität Marburg,

291

Germany) were cultured in DMEM supplemented with the above as well as geneticin (0.3

292

mg/ml) and puromycin (2 µg/ml). Expression of TMPRSS2 was induced with doxycycline (0.5

293

µg/ml) (55). At 70% confluence, the MDCK cells were infected at a multiplicity of infection (moi)

294

of 10 with NIBRG-14 ma, NIBRG-23 or H5N1 H274Y ma. Samples containing monovalent N1-

295

3/5/7-VHHm, bivalent N1-3/5-VHHb, bivalent N1-3/5/7/GP4-VHH-Fc, Oseltamivir or control

296

medium were mixed with 0.8% Avicel RC-591™ as overlay (56). After 2 to 4 days of incubation,

297

the cells were fixed with 4% paraformaldehyde in PBS for 30 min. The cells were permeabilized

298

with 20 mM glycine containing 0.5% Triton X-100 (v/v). After blocking with BSA, the cells were

299

incubated for 2 h with polyclonal α-NIBRG-14 (1:1000) or anti-M2e monoclonal antibody

300

(1:5000). After washing, α-mouse whole IgG-HRP was used to visualize plaques with the

301

substrate TrueBlue™ Peroxidase Substrate (KPL).

302

Mouse experiments

303

All mouse experiments were conducted according to national (Belgian Law 14/08/1986 and

304

22/12/2003, Belgian Royal Decree 06/04/2010) and European legislation (EU Directives

305

2010/63/EU, 86/609/EEG) on animal regulations. All experiments on mice were approved by the

306

ethics committee of Ghent University (permit number LA1400091) and efforts were made to

15

307

avoid or diminish suffering of the animals. Specific-pathogen–free female BALB/c mice, 7–9

308

weeks old, were purchased from Charles River (Germany) and used in all experiments. Mice

309

were housed in ventilated cages with high-efficiency particulate air filters in temperature-

310

controlled, air-conditioned facilities and had food and water ad libitum. Mice were anaesthetized

311

by intraperitoneal injection of xylazine (10 μg/g) and ketamine (100 μg/g) before intranasal

312

administration of N1-VHHm, N1-VHHb or N1-VHH-Fc. The N1-VHHm in any format were diluted

313

in endotoxin-free phosphate-buffered saline (PBS) with 1% (w/v) bovine serum albumin, and

314

applied intranasally in doses ranging from 0.25 to 5 mg/kg, as indicated in the figure legends.

315

Mice were challenged 4 or 24 h later, as indicated, with 6 LD50 (H5N1 H274Y ma) or 4 LD50

316

(NIBRG-14 ma) of virus (50 μl, divided equally between the nostrils). A 30% loss in body weight

317

was the end point at which sick mice were euthanized. On day 4 after challenge, lung

318

homogenates were prepared in PBS, cleared by centrifugation at 4°C, and used for virus

319

titration. Monolayers of MDCK cells were infected with 50 μl of serial 1:10 dilutions of the lung

320

homogenates in a 96-well plate in serum-free Dulbecco's modified Eagle medium (Invitrogen)

321

supplemented with L-glutamine, non-essential amino acids, penicillin and streptomycin. After 1

322

h, the inoculum was replaced by medium containing 2 μg/ml of L-(tosylamido-2-phenyl) ethyl

323

chloromethyl ketone–treated trypsin (Sigma-Aldrich). End-point virus titers were determined by

324

hemagglutination of chicken red blood cells and expressed as TCID50 per milliliter. IAV RNA

325

levels were determined by quantitative RT-PCR. RNA was isolated from 150 μl of cleared lung

326

homogenate using the Nucleospin RNA virus kit (Machery-Nagel). The relative amount of

327

NIBRG-14 ma genomic RNA was determined by preparing viral cDNA and performing

328

quantitative PCR with M-genomic segment primers (CGAAAGGAACAGCAGAGT and

329

CCAGCTCTATGCTGACAAAATG) and probe (GGATGCTG) (probe no. 89; Universal

330

ProbeLibrary, Roche) and the LightCycler 480 Real-Time PCR System (Roche).

16

331

Selection and characterization of H5N1 N1-VHHm escape mutant viruses

332

NIBRG-14 ma virus was serially diluted and used to infect MDCK cells in the presence of N1-3-

333

VHHm (4.2 µM) or N1-H5-VHHm (3.7 µM), a dose corresponding to ten-fold the IC50 (Table 2).

334

Escape mutants were screened by monitoring cytopathic effects. After 9 passages in the

335

presence of N1-3-VHHm or N1-5-VHHm, escape viruses were plaque purified by growth on

336

MDCK cells overlaid with 0.6% low-melting agarose in the presence of N1-3-VHHm or N1-5-

337

VHHm. Escape virus isolates were amplified on MDCK cells, still in the presence of a ten-fold

338

IC50 of N1-3-VHHm or N1-5-VHHm. Total RNA was extracted from the supernatant and used to

339

clone the cDNA corresponding to the NA viral RNA segment (57). Deduced amino acid

340

substitutions in the NA sequence of escape viruses were modeled with PyMol (Delano

341

Scientific,

342

A/Vietnam/1194/2004 (PDB code 2HTY). N1-VHH binding assays were performed as described

343

(35). Briefly, 4-8 HA unit of wild type or plaque purified candidate escape viruses were used to

344

coat ELISA plates. After blocking, wells were incubated with different concentrations of N1-3-

345

VHHm, N1-3-VHHb, N1-5-VHHm or N1-5-VHHb. Binding of N1-VHHs was revealed by

346

incubation with an anti-His tag monoclonal antibody, followed by a secondary sheep anti-mouse

347

IgG antibody conjugated to HRP and visualized by addition of TMB substrate and absorbance

348

was measured at 450/655 nm.

349

Statistical analysis

350

Graphpad software (Graphpad Prism, version 6) was used for statistical analysis. For

351

comparing two groups in the fetuin and AVINA assays, t test was used. Differences between

352

mouse groups were tested using two-way ANOVA. When this test demonstrated a significant

353

difference between groups (P < 0.05), t tests were used to compare two groups on each day.

354

Kaplan-Meier survival curves were plotted and differences in survival were analyzed by the

355

Mantel-Cox log-rank test.

http://www.pymol.org)

using

the

H5N1

NA

structure

derived

from

17

356

RESULTS

357

Production of recombinant tetrameric neuraminidase

358

A Baculovirus Expression Vector System was used to produce recombinant soluble tetrameric

359

and enzymatically active H5N1 NA derived from A/crested eagle/Belgium/01/2004 (N1rec). The

360

N1rec coding sequence was cloned into a pAcMP2 expression vector (pAcMP2n1rec), under

361

the control of the AcNPV basic protein promoter that is active in the late phase of the

362

Baculovirus infection cycle (Fig. 1A). Infection of Spodoptera frugiperda (Sf9) cells with the

363

recombinant Baculovirus AcNPVN1rec resulted in sialidase activity in the cell supernatant,

364

suggesting that the N1rec product was soluble and enzymatically active NA (Fig. 1A). N1rec

365

protein was purified from the supernatant of infected Sf9 cells by three successive

366

chromatography steps, including size exclusion chromatography to collect tetrameric N1rec (see

367

Materials and Methods). Using MUNANA as a substrate, we calculated that purified N1rec had

368

a specific activity of 56,851 NA units/mg.

369

Immunization and VHH phage library construction

370

N1rec was used as an antigen for the generation and selection of NA-specific VHH. From an

371

immunized alpaca, we obtained a VHH phage library of 2 X 108 independent transformants,

372

57% of which had a cDNA insert of the correct size (data not shown). Seventy-eight positive

373

clones were selected, 13 of which had unique VHH sequences. The complementarity

374

determining region 3 (CDR3) had the most variable amino acid sequence among the 13 N1rec-

375

binding VHHs (N1-VHHm) (Fig. 1C). In silico structure prediction of the monomeric N1-VHHm

376

candidates showed that the CDR3 of N1-1-VHHm, N1-3-VHHm, N1-4-VHHm, N1-5-VHHm and

377

N1-6-VHHm had an electronegative protruding paratope (Fig. 1D).

378

Production and characterization of soluble monomeric N1-VHHm

18

379

The binding of the N1-VHHm proteins to N1rec was assessed by ELISA and compared with

380

binding to the recombinant M2e-tGCN4 protein, which like N1rec contains the same

381

tetramerizing leucine zipper (Table 1 and Fig. 1E). Except for N1-11-VHHm and N1-13-VHHm,

382

none of the N1-VHHm bound substantially to recombinant M2e-tGCN4 protein. We next used

383

the MUNANA substrate to assess the ability of the N1-VHHm candidates to inhibit the

384

enzymatic activity of N1rec. Four candidates, N1-1-VHHm, N1-3-VHHm, N1-5-VHHm and N1-6-

385

VHHm, inhibited the N1rec catalytic activity (Table 1). N1-3-VHHm and N1-5-VHHm inhibited

386

N1rec in the sub-micromolar range (Table 2), and together with N1-7-VHHm, they were

387

analyzed by SPR to determine their affinity for immobilized N1rec. Both N1-3-VHHm and N1-5-

388

VHHm had a high affinity for N1rec, with an equilibrium dissociation constant (KD) in the low

389

nanomolar (N1-5-VHHm) to high picomolar (N1-3-VHHm) range. N1-7-VHHm showed

390

approximately 10- to 100-fold lower affinity for N1rec than N1-5-VHHm and N1-3-VHHm,

391

respectively (Table 1). These binding affinities of N1-3-VHHm and N1-5-VHHm for N1rec

392

resemble the affinities reported for a monomeric VHH that inhibits the activity of lysozyme (33).

393

Competitive SPR analysis also showed that prior binding of N1-3-VHHm to N1rec abolished

394

subsequent binding of N1-1-VHHm, N1-5-VHHm and N1-6-VHHm. Conversely, N1-7-VHHm

395

binding to N1rec was not affected by prior incubation of N1rec with N1-3-VHHm or N1-5-VHHm

396

(Fig. 1F). This suggests that N1-3-VHHm and N1-5-VHHm bind an overlapping epitope within

397

N1rec, or that binding of N1-3-VHHm sterically hinders binding of N1-5-VHHm, while N1-7-

398

VHHm binds a different epitope.

399

We next analyzed the antiviral potential of N1-3-VHHm and N1-5-VHHm using the H5N1 strain

400

NIBRG-14 ma. The amino acid sequences of the NA of NIBRG-14 ma (derived from

401

A/Vietnam/1194/2004) and N1rec (derived from A/crested eagle/Belgium/01/2004) are highly

402

homologous, with only 13 amino acid differences, none of which map in or close to the NA

403

catalytic site. In the presence of N1-3-VHHm or N1-5-VHHm, the size of NIBRG-14 ma plaques

19

404

was reduced in a concentration dependent manner, and with IC50 values in low nanomolar

405

range (Table 2). In contrast, N1-7-VHHm did not affect the size or number of NIBRG-14 ma

406

plaques (Table 2). These results indicate that the in vitro antiviral potential of NA-specific VHH

407

proteins depends on their NA-inhibitory activity.

408

Bivalent N1-VHHb formats have enhanced NA inhibitory and antiviral potential

409

It has been reported that generation of multivalent formats of VHHs increases their functional

410

affinity by introducing avidity (34). For example, the introduction of avidity drastically decreased

411

the dissociation constant (Koff) of the VHH molecules directed against lysozyme, and

412

significantly improved their enzyme inhibitory activity over their monovalent format (51). To

413

increase the avidity of the N1-VHHm, we compared two different bivalent formats. In one

414

approach, we used the llama IgG2c hinge (17 amino acid residues) as a flexible linker to fuse

415

two identical N1-VHHm gene moieties in tandem, resulting in the bivalent n1-3-vhhb and n1-5-

416

vhhb expression elements (Fig. 1B and 2A). Despite repeated attempts, we could not generate

417

an N1-7-VHHb expression plasmid. The corresponding proteins were expressed and purified

418

from E. coli and their in vitro N1-VHHb NA-inhibitory and antiviral activities were compared with

419

those of their monovalent counterparts. N1rec inhibition by N1-3-VHHb and N1-5-VHHb was

420

2000- and 500-fold stronger than the corresponding N1-3-VHHm and N1-5-VHHm formats,

421

respectively (Table 2). Surprisingly, in a plaque assay using MDCK cells infected with NIBRG-

422

14 ma virus, both N1-3-VHHb and N1-5-VHHb seemingly had antiviral activities that were

423

comparable to the activities of their monovalent counterparts (data not shown). In this assay,

424

exogenous trypsin is used to facilitate maturation of HA in newly produced virions to allow

425

multicycle replication of the recombinant NIBRG-14 ma virus. We noticed that the llama IgG2c

426

hinge linker used to generate bivalent VHH formats was sensitive to trypsin cleavage (Fig. 2A).

427

To circumvent the use of exogenously added trypsin, we took advantage of TMPRSS2 MDCK

428

cells, which are transformed with the doxycycline-inducible serine protease TMPRSS2 and

20

429

allow multicycle replication of IAV in the absence of exogenous trypsin (55). Using monolayers

430

of TMPRSS2 MDCK cells for infection with NIBRG-14 ma, the antiviral effect of N1-3-VHHb and

431

N1-5-VHHb was increased 240- and 58- fold, respectively, compared with their monovalent

432

formats (Table 2). These results obtained with N1-VHHb indicate that there is a significantly

433

enhanced NA-inhibitory and antiviral activity of both bivalent N1-VHHb compared with the

434

monovalent N1-VHHm formats.

435

We next engineered N1-VHH variants fused to a mouse IgG2a-derived Fc domain as an

436

alternative bivalent format (N1-Vhh-Fc). To produce these more complex proteins, we used a

437

robust plant seed-based expression system with reported high yield of recombinant antibodies

438

(58, 59). For this, the n1-3-vhh, n1-5-vhh and the n1-7-vhh genes were fused to the sequence

439

encoding the hinge and Fc tail of mouse IgG2a (n1-vhh-fc). These n1-vhh-fc constructs were

440

cloned into the binary vector Pphas as a T-DNA expression cassette (Fig. 1B and 2B).

441

Subsequently, transgenic Arabidopsis thaliana plants were generated by Agrobacterium-

442

mediated floral dip transformation. Seed extracts of segregating T3 Arabidopsis transformants

443

were screened by ELISA using antibodies against the Fc part for the best expressers of the

444

recombinant dimeric N1-VHH-Fc proteins. N1-3-VHH-Fc and N1-5-VHH-Fc accumulated to high

445

levels in several transformants, and levels of 16 % of total soluble seed protein were found. On

446

the contrary, N1-7-VHH-Fc did not accumulate well and the best expressors contained about

447

1% VHH-Fc of total soluble seed protein (data not shown). A good correlation was observed

448

between the ELISA result and NA inhibitory activity in crude seed extracts from different T3

449

transformants (data not shown). N1-3-VHH-Fc, N1-5-VHH-Fc and N1-7-VHH-Fc were purified

450

from seed extracts by protein G affinity chromatography. Reducing SDS-PAGE and Coomassie

451

staining showed that the N1-VHH-Fc monomers migrated at approximately 42 kDa (Fig. 2B).

452

N1-3-VHH-Fc and N1-5-VHH-Fc had stronger N1rec inhibition capacity and in vitro antiviral

453

activity than their monovalent N1-3-VHHm and N1-5-VHHm counterparts (Table 2). Bivalent N1-

21

454

7-VHH-Fc bound to N1rec in ELISA assays (data not shown), but it failed to inhibit N1rec and

455

the H5N1 viruses in vitro. Taken together, these results indicate that the in vitro NA inhibition

456

and antiviral activity against NIBRG-14 ma virus is enhanced at least 30-fold by using bivalent

457

N1-VHHb and N1-VHH-Fc formats, compared with the monovalent N1-VHHm domain.

458

In vitro antiviral activity of N1-VHH against H5N1 clade 2.2, H5N1 Oseltamivir-resistant

459

and pandemic H1N1 2009 virus

460

Like NA from NIBRG-14, N1rec NA (derived from A/crested eagle/Belgium/01/2004) belongs to

461

the clade 1 of the H5N1 NAs. To assess the potential cross reactivity of the NA activity inhibition

462

(NAI) of the isolated VHHs in the three available formats (N1-VHHm, N1-VHHb and N1-VHH-

463

Fc), we focused first on NIBRG-23 NA (derived from A/turkey/Turkey/01/2005), a clade 2.2

464

H5N1 virus (60). Although there are only a few laboratory-confirmed human infections with

465

H5N1, it appears that clade 2 H5N1 viruses are responsible for most of these zoonotic

466

infections with HPAIV (61). N1-3-VHHm and N1-5-VHHm reduced in vitro growth of NIBRG-23

467

virus with an IC50 in the low micromolar range (Table 2). N1-3-VHHb, N1-5-VHHb, N1-3-VHH-Fc

468

and N1-5-VHH-Fc displayed about 150-fold higher in vitro antiviral activity compared to the

469

corresponding monovalent N1-VHHm against this clade 2 virus, as judged by a plaque size

470

reduction assay. Based on these findings, we conclude that the two NA-inhibitory VHHs inhibit

471

in vitro replication of representative H5N1 IAV from clades 1 and 2 with a comparable efficiency,

472

suggesting that they target a common epitope shared by the NA of these viruses.

473

Oseltamivir-resistant IAV frequently emerges and spreads in the human population. Several

474

mutations have been reported to contribute to Oseltamivir-resistance, but among them the

475

H274Y mutation (N2 numbering) is the most common found in Oseltamivir-resistant viruses

476

(62). We used reverse genetics to generate a clade 1 H5N1 virus harboring the H274Y mutation

477

in NA (derived from A/crested eagle/Belgium/01/2004), while the HA segment was derived from

22

478

NIBRG-14 and the remaining six segments from PR/8, resulting in H5N1 H274Y virus. Next, we

479

determined if N1-3-VHHm, N1-5-VHHm, N1-3-VHHb, N1-5-VHHHb, N1-3-VHH-Fc and N1-5-

480

VHH-Fc can inhibit this virus. As so far all N1-3-VHH formats performed alike N1-5-VHH formats

481

in vitro, we were surprised that only N1-3-VHH (monovalent and both bivalent formats), but not

482

N1-5-VHH in any format, reduced the growth of H5N1 H274Y (Table 2). Compared to NIBRG-

483

14 ma, the H5N1 H274Y IC50 values were three- to seven-fold higher, but still in the low

484

nanomolar range. Even though the competitive SPR analysis suggested that the N1-3-VHHm

485

and N1-5-VHHm epitopes in NA overlap (Fig. 1F), the contact residues necessary for their

486

binding are probably not identical. We conclude that the three evaluated formats of N1-3-VHH

487

can inhibit growth of H5N1 H274Y viruses in vitro, and that both bivalent formats have

488

enhanced in vitro antiviral activity. In addition, the H274Y mutation seems to be sufficient to

489

abolish the in vitro antiviral effect of the N1-5-VHH formats.

490

Next, we tested the antiviral potential of all N1-3-VHH and N1-5-VHH formats against a

491

pandemic H1N1 2009 virus isolate (pH1N1). We used fetuin and MUNANA (AVINA assay) as

492

two alternative substrates for virion-associated NA activity. Using NIBRG-14 ma, and based on

493

MUNANA hydrolysis, monovalent N1-3-VHHm and N1-5-VHHm had significant inhibitory

494

activity, compared with the negative controls SA-VHHm (P < 0.05) and PBS (P < 0.01), although

495

this tendency was not significant in the fetuin assay (Fig. 3A). We attribute these differences

496

between the two assays to the type of substrate used: small molecule MUNANA versus fetuin,

497

which is a large protein carrying multiple sialylated glycans. In the case of fetuin, HA plays an

498

important role in recruiting NA to the substrate, which may alter the accessibility of the N1-VHHs

499

to the NA target. We found that the bivalent molecules N1-3-VHHb, N1-5-VHHb, N1-3-VHH-Fc

500

and N1-5-VHH-Fc significantly inhibited NA activity for both substrates (P < 0.01, Fig. 3A).

501

In line with our previous results using H5N1 H274Y IAV, only N1-3-VHHb and N1-3-VHH-Fc

502

showed NAI activity using both substrates. (P < 0.05, or P < 0.01) (Fig. 3B). N1-5-VHHb and

23

503

N1-5-VHH-Fc showed inhibitory potential against pH1N1 in both assays. (P < 0.05) (Fig. 3C).

504

We could not detect binding of the N1-VHHs to A/NewCaledonia/20/99 or inhibition of NA

505

activity of this virus (Fig. 3D and data not shown). These AVINA and fetuin results for NIBRG-14

506

ma and H5N1 H274Y NA are in accordance with the N1rec inhibition and the plaque size

507

reduction assays mentioned before (Table 2). In addition, both bivalent N1-5-VHHb formats

508

inhibited pH1N1 virion-associated NA activity, suggesting a degree of intra subtype inhibitory

509

effect of the N1-VHHs.

510

N1-3-VHHm and N1-5-VHHm treatment reduce early morbidity, but not viral lung titer, in

511

H5N1-challenged mice

512

We next evaluated the in vivo antiviral effect of N1-3-VHHm and N1-5-VHHm. The endotoxin

513

levels of the VHH preparations that were used in vivo were low and comparable for the NA-

514

specific VHH proteins and their irrelevant control counterparts (Table 3). In the first experiment

515

we administered intranasally 100 µg of N1-3-VHHm, N1-5-VHHm or N1-7-VHHm to BALB/c

516

mice 4 hours before challenge with 4 LD50 of NIBRG-14 ma. Two groups were included as

517

positive controls (i) a group of mice that received 30 µg of H5-VHHb, a bivalent NIBRG-14 ma

518

HA neutralizing VHH (36), and (ii) a group of mice that received daily oral administration of

519

Oseltamivir at a high dose (45 mg/kg/day). Mice treated with N1-3-VHHm or N1-5-VHHm

520

showed significantly reduced morbidity at 72 and 96 h after infection, compared with the groups

521

treated with N1-7-VHHm and PBS (P < 0.001) (Fig. 4A). Four days after infection, the mice were

522

sacrificed to determine lung virus titers by endpoint dilution in a TCID50 assay. All treated groups

523

of mice (N1-3-VHHm, N1-5-VHHm, N1-7-VHHm, SA-VHHm, PBS and Oseltamivir), except the

524

H5-VHHb group, had a high lung virus load ranging from 104.75 to 106.48 TCID50/ml (data not

525

shown). We also measured the amount of viral RNA in the lung homogenates using a genome

526

strand-specific RT-qPCR method. Except for the samples derived from the H5-VHHb treated

527

mice (P < 0.001), all the N1-VHH treated groups showed high viral RNA levels 96 h after

24

528

infection, comparable to those in the PBS-treated group (Fig. 4B). NA activity in the lung

529

homogenates of the treated groups, measured by AVINA, indicated that there was no difference

530

between the N1-VHH and PBS treated groups, whereas NA activity was strongly reduced in the

531

H5-VHHb and Oseltamivir groups (Fig. 4C). These results are in line with the RT-qPCR results

532

and confirmed the presence of a high viral titer in the lung homogenates of the N1-VHH groups,

533

despite low morbidity. We speculate that the N1-VHH treatment did not have a major impact on

534

virus entry in this in vivo challenge model, but rather reduced virus spread within the lung

535

compartment. A recent systems analysis of influenza A virus-associated pathology in the

536

mouse, revealed a positive correlation between virus spread in the lung tissue and high

537

pathogenicity irrespective of the number of infectious virus particles that were produced (63). In

538

addition, shearing forces that are generated during the preparation of lung homogenates may

539

have released virions from the cell membranes, of which the release was hindered in the intact

540

lung by NA-inhibitory nanobodies. We conclude that intranasal administration of N1-3-VHHm

541

and N1-5-VHHm prevents body weight loss after challenge with NIBRG-14 ma virus during the

542

early stage of H5N1 IAV infection, without apparent decrease in viral titer in the lung

543

homogenates.

544

Protection by bivalent N1-VHHb against H5N1 challenge

545

The in vitro results indicated that the potency of the bivalent formats of the inhibitory N1-VHHs

546

against the tested H5N1 viruses was 30- to 240-fold higher than their monovalent counterparts

547

(Table 2). We therefore assessed if this heightened antiviral effect would also be reflected in an

548

in vivo challenge experiment. We first determined the protective potential during the early

549

stages of H5N1 infection. Four hours before challenge with 4 LD50 of NIBRG-14 ma, groups of

550

BALB/c mice were intranasally given 60 µg of N1-3-VHHb, N1-5-VHHb, or BL-VHHb (a bivalent

551

VHH directed against the irrelevant bacterial target β-lactamase), and 84 µg of N1-3-VHH-Fc,

552

N1-5-VHH-Fc, or GP4-Fc (a plant-produced Coronavirus GP4-Fc fusion protein). Treatment with

25

553

N1-3-VHHb or N1-3-VHH-Fc significantly improved morbidity 72 and 96 h after infection (hpi)

554

compared with the negative control groups GP4-Fc and PBS (P < 0.001) (Fig. 4D). The

555

protection against weight loss was comparable to that observed with the positive controls (H5-

556

VHHb and Oseltamivir, the latter given daily at 45 mg/kg/day by gavage). Treatment with the

557

bivalent formats of N1-5-VHH resulted in reduction of morbidity that was significant at 60, 72

558

and 96 hpi compared with the negative controls (P < 0.001) (Fig. 4D). Determination of the lung

559

virus load on day four after challenge revealed that all challenged groups, except the H5-VHHb-

560

treated mice (no virus detectable), had high and comparable lung virus loads (data not shown).

561

In addition, the NA activity in the lung homogenates of the groups treated with N1-VHH bivalent

562

formats was assessed by an AVINA assay. The lung homogenates treated with bivalent N1-

563

VHHs (N1-3-VHHb, N1-5-VHHb, N1-3-VHH-Fc and N1-5-VHH-Fc) showed 2-5 fold decrease in

564

NA activity when compared to the BL-VHHb (P < 0.001), GP4-Fc (P < 0.05, or P < 0.01) and

565

PBS treated groups (P < 0.01) (Fig. 4E). This could be attributable to residual inhibitory bivalent

566

N1-VHHb in the lung preparation, possibly due to an increased half life of the bivalent N1-VHHs,

567

relative to N1-VHHm, for which no significant reduction in lung associated NA activity was

568

demonstrated (Fig. 4C). Taken together, treatments with bivalent formats of N1-3-VHH and N1-

569

5-VHH improve protection during the first four days following NIBRG-14 ma challenge and

570

reduce the lung associated NA activity, a surrogate for viral load, in H5N1-infected mouse lungs.

571

Dose-dependent protection against H5N1 challenge

572

To test if prophylactic administration of N1-3-VHHb, N1-3-VHH-Fc, N1-5-VHHb or N1-5-VHH-Fc

573

can protect mice against a lethal (4 LD50) challenge with NIBRG-14 ma, we repeated the

574

protection studies and followed the mice for two weeks. We focused on the most potent N1-

575

VHHb bivalent formats and assessed their protective efficacy in a dose-response experiment.

576

Groups of four BALB/c mice were treated intranasally with 60, 12, 2.5 or 0.5 µg of N1-3-VHHb

577

or N1-5-VHHb. In parallel, another group was treated daily by oral administration of Oseltamivir

26

578

(45 mg/kg/day). Negative control groups were treated with 60 µg of BL-VHHb or with PBS

579

before challenge. Mice that had received 60 µg of N1-3-VHHb, N1-5-VHHb or BL-VHHb before

580

the challenge also received a second intranasal dose of 60 µg of the same bivalent VHH on day

581

6 after the challenge. All mice from the PBS and BL-VHHb treatment groups died 9-10 days

582

after infection (Fig. 5C and D). In contrast, treatment with Oseltamivir and intranasal high dose

583

N1-3-VHHb or N1-5-VHHb (60 µg) protected the mice against lethality. In agreement with this,

584

Oseltamivir and N1-5-VHHb (60 µg) reduced post-challenge body weight loss compared to

585

control treated mice (Fig. 5A and B). Although administration of the highest amount of N1-3-

586

VHHb prevented death, it did not affect body weight loss. A single intranasal dose of 12 µg or

587

2.5 µg of N1-3-VHHb or N1-5-VHHb provided partial protection against mortality but failed to

588

reduce morbidity (Fig. 5A-D).

589

We next evaluated protection against a potentially lethal challenge with NIBRG-14 ma by prior

590

single dose intranasal administration of the plant-produced N1-3-VHH-Fc and N1-5-VHH-Fc

591

formats. N1-3-VHH-Fc (80 µg and 17 µg) as well as Oseltamivir prevented death following

592

challenge with NIBRG-14 ma (Fig. 6C), whereas treatment with 3.5 µg of N1-3-VHH-Fc

593

prevented death of some animals. However, challenge was associated with significant body

594

weight loss for all doses and treatments (Fig. 6A and 6B).

595

When comparing the protective efficacy of the two bivalent formats of N1-3-VHH and N1-5-

596

VHH, it appeared that the Fc moiety in the N1-VHH-Fc formats provides an extra protective

597

effect against morbidity (compare 60 µg N1-3-VHHb with 84 µg N1-3-VHH-Fc in figures 5A and

598

6B) as well as against mortality following NIBRG-14 ma challenge (Fig. 5 and 6). We conclude

599

that a single dose of both bivalent N1-3/5-VHH formats can protect mice against a potentially

600

lethal challenge with an H5N1 virus, and that Fc-fused VHHs provide better protection,

601

suggesting that Fc Receptor binding molecules contribute to protection.

27

602

N1-3-VHHb and N1-3-VHH-Fc protect against challenge with Oseltamivir-resistant H5N1

603

Our in vitro analysis demonstrated that monovalent and bivalent formats of N1-3-VHH, but not

604

N1-5-VHH, reduced the growth of Oseltamivir-resistant H5N1 H274Y virus (Table 2). To

605

evaluate these findings in vivo, groups of six BALB/c mice received 30 µg of N1-3-VHHb, N1-5-

606

VHHb or BL-VHHb by intranasal administration 24 h before challenge with 4 LD50 of NIBRG-14

607

ma or 6 LD50 of H5N1 H274Y ma. In parallel, a group was treated by daily oral administration of

608

Oseltamivir (1 mg/kg/day), a dose that has been reported to fully protect laboratory mice against

609

challenge with highly pathogenic H5N1 (64). In line with our previous results, mice treated with

610

N1-3-VHHb, N1-5-VHHb or Oseltamivir suffered from substantial but transient weight loss, but

611

they survived challenge with NIBRG-14 ma virus, whereas mice receiving PBS or BL-VHHb

612

died before day 10 after challenge (Fig. 7). This result demonstrates that a single intranasal

613

administration of N1-3-VHHb or N1-5-VHHb is sufficient to protect against a subsequent lethal

614

challenge (24 h later) with an H5N1 virus. Following challenge with H5N1 H274Y virus, the N1-

615

3-VHHb treated mice were severely sick, but all mice survived (Fig. 7A and 7C). All other

616

groups, including the Oseltamivir treated mice, died after challenge with the H5N1 H274Y virus

617

by day 10 after infection (Fig. 7A and 7C). Taken together, bivalent N1-3-VHHb treatment of

618

mice before challenge with an Oseltamivir-resistant H5N1 virus does not control morbidity

619

adequately but results in full recovery and survival. Furthermore, bivalent N1-5-VHHb does not

620

protect mice against this challenge, suggesting that the N1-VHHb formats (i.e. tandem repeats

621

of VHH separated by a short linker) require the NA inhibitory activity for in vivo protection.

622

Finally, we compared the protective potential of intranasally instilled N1-3-VHH-Fc, N1-5-VHH-

623

Fc and N1-7-VHH-Fc in our lethal challenge model. Following challenge with 4 LD50 of NIBRG-

624

14 ma, all mice in the N1-3-VHH-Fc and Oseltamivir groups lost body weight up until day 7 after

625

challenge (Fig. 7B) and then started to recover, and all mice survived this challenge (Fig. 7D). In

626

contrast, all PBS or GP4-Fc treated mice, except one, became morbid and died by 10 days post

28

627

infection (Fig. 7B and 7D). In the N1-7-VHH-Fc group, only one out of six mice died after

628

challenge. This is very surprising because N1-7-VHHm and N1-7-VHH-Fc have no detectable

629

antiviral activity in vitro. Remarkably, in this experiment mice that had received N1-5-VHH-Fc

630

did not lose weight after challenge with NIBRG-14 ma (Fig. 7B). A challenge with 6 LD50 of

631

H5N1 H274Y virus was lethal to all mice except those that had been pretreated with N1-3-VHH-

632

Fc (all mice survived) or N1-7-VHH-Fc (four out of six mice survived), though all animals lost

633

substantial body weight (Fig. 7B and 7D). We conclude that protection against a potentially

634

lethal challenge with Oseltamivir-resistant H5N1 virus can be obtained by a single dose of N1-3-

635

VHH-Fc or with N1-7-VHH-Fc, although the latter construct is less potent. This indicates that the

636

IgG2a Fc moiety contributes to protection, even in the absence of NA inhibitory activity.

637

N1-5-VHHm escape mutants

638

To identify the amino acid residues involved in the binding of N1-3-VHHm and N1-5-VHHm to

639

H5N1 NA, escape viruses were selected and plaque purified after serial passage of NIBRG-14

640

ma virus in MDCK cells, in the presence of 10 IC50 of N1-3-VHHm or N1-5-VHHm, essentially as

641

described before (35). After nine passages, candidate escape mutants were plaque purified and

642

the NA sequences of four (N1-5-VHHm selection) and three (N1-3-VHHm selection) isolated

643

escape viruses were determined. One of the escape viruses that was selected with N1-5-VHHm

644

contained a mutation in the NA coding region that resulted in a change from isoleucine to

645

threonine at position 437 in the deduced amino acid sequence. This substitution is located on

646

the NA surface, relatively close to the active site (Fig. 8A). A conserved NA epitope that is

647

recognized by HCA-2 under denaturing conditions (65, 66) and localized within the NA catalytic

648

site, and the amino acids necessary for binding a recently reported set of mAbs that recognize

649

H1N1 and H5N1 NA (67) do not overlap with this N1-5-VHHm escape mutations (Fig. 8A).

650

Sequence alignment of a set of representative N1 influenza A viruses, including human H1N1,

651

other H5N1 viruses and an avian H6N1 virus showed that I437 is conserved in these NA

29

652

sequences (Fig. 8B). Binding studies using immobilized virus revealed that the I437T escape

653

virus was no longer bound by N1-3-VHHm, N1-3-VHHb, N1-5-VHHm and N1-5-VHHb (Fig 8C).

654

This result is in accordance with the Biacore data, which showed that N1-3-VHHm and N1-5-

655

VHHm compete for binding to recombinant H5N1 NA (Fig. 1F) and suggests that I437 is part of

656

the VHH epitope in NA or stabilizes the conformation of that epitope.

657

Interestingly, the other three plaque purified N1-5-VHHm escape viruses and all three plaque

658

purified viruses that grew in the presence of N1-3-VHHm contained wild type NA. Possibly,

659

these viruses carry compensatory mutations in other viral proteins such as HA as was described

660

before for influenza A virus passaged in vitro in the presence of small molecule inhibitors of NA

661

(68). Taken together, H5N1 viruses that overcome the inhibition by N1-3-VHHm and N1-5-

662

VHHm can be selected in vitro and one way of escape is through the acquisition of a mutation in

663

NA that results in loss of binding to the VHHs.

664

30

665

DISCUSSION

666

NA Abs usually do not virus neutralizing and are thought to exert their antiviral effects further

667

downstream in the IAV infection process. Still, NA-specific antibodies can significantly reduce

668

viral replication and disease severity (69, 70). Ab titers against N1 or N2 have been associated

669

with high NA enzymatic activity in the vaccine preparation, suggesting that the NA content and

670

activity in IAV vaccines predicts immunogenicity (71). NA of human influenza viruses is subject

671

to antigenic drift at a pace that is comparable to that of HA (72). This suggests that NA is

672

subject to intense immune selection pressure, which is surprising given that antibodies directed

673

against NA do not neutralize the virus in vitro and that during natural infections immune

674

responses against HA are more pronounced than those against NA, presumably because HA

675

outnumbers NA on influenza virions. Biologically, HA and NA cooperate and antigenic variation

676

in NA is in part driven by antigenic escape selection pressure on HA (73). Nevertheless,

677

antibodies directed against NA remain underexplored as potential treatment against influenza.

678

Our N1-VHHs are not the first inhibitory mAbs described against H5N1 NA. Recently, the

679

IgG2bκ mAb 2B9 directed against clade 1 A/Vietnam/1203/04 NA showed interclade and

680

Oseltamivir-resistant mutant NAI in vitro (74). In NAI assays using MUNANA and NA-Star

681

substrates, the 2B9 mAb IC50 ranged from 1 to 8 µg/ml against H5N1 clade 1 and clade 2.1

682

virions, and against an Oseltamivir-resistant double mutant H274Y and N294S. In another

683

recent report on IgG2b mAbs targeting multiple N1 NA, intra-subtype mAbs showed a relatively

684

high H1N1 NAI potency (IC50 of 23.6 to 68.1 ng/ml), which was even lower when

685

A/Vietnam/1203/04 H5N1 NAI potency was assessed (1.41-2.42 µg/ml) (67). We used two

686

inhibitory N1-VHH formats to assess NA inhibitory and antiviral potential. Results from our

687

N1rec inhibition assays showed that N1-3-VHHm and N1-5-VHHm IC50 was 425 nM (6.3 µg/ml)

688

and 374 nM (5.6 µg/ml), respectively, and the introduction of bivalency (N1-VHHb and N1-VHH-

689

Fc) resulted in 2-3 log-fold further increase in their inhibitory potential (5.2-125 ng/ml) (Table 2).

31

690

The above mentioned conventional mAbs, which are bivalent, have an H5N1 NAI activity that is

691

similar to our N1-VHHm. We hypothesize that the much smaller size of VHH molecules

692

compared with conventional antibodies compensates for the lack of avidity that is present in

693

conventional antibodies to explain this.

694

Harmsen et al. reported the first VHH targeting neuraminidase (39). In contrast with our work,

695

where we used recombinant NA protein to immunize an alpaca, these authors obtained VHH by

696

immunizing llamas using whole IAV. This strategy yielded mainly HA and NP binders, and

697

required a cross-selection strategy with homologous NA, followed by further selection with

698

recombinant NA to obtain NA binders. This approach resulted in VHH binding to homologous

699

and non-homologous IAV, but only some of the NA cross-binding VHHs showed NA inhibition

700

activity (fetuin substrate). It is also important to point out that the recombinant NA activities used

701

by Harmsen et al. were 316 to 13.5 times lower than our N1rec specific activity (57

702

µmol/mg/min), while their VHH NA inhibition IC50 values ranged from 124 ng/ml to 2 ng/ml, using

703

homologous NA. The apparently lower IC50 values of the Harmsen VHHs (3 to 200 fold lower

704

than for our N1-VHHm) could therefore be explained by a 1-2 log lower NA specific subtype

705

activity. All together, these results suggest that a similar NAI potency range was obtained with

706

our monovalent N1-VHHm. In vitro, monovalent N1-3-VHHm and N1-5-VHHm had an IC50 that

707

was almost identical to that of Oseltamivir for N1rec and for the two Oseltamivir-sensitive H5N1

708

viruses we assayed (table 2).

709

In the MDCK TMPRSS2-based assay, which allowed us to assess multiple rounds of viral

710

replication in the absence of exogenous trypsin, the antiviral effect of bivalent VHH formats (N1-

711

VHHb and N1-VHH-Fc) was clearly enhanced compared with monovalent N1-VHHm molecules

712

(Table 2). For NIBRG-14 ma, the N1-VHH concentration necessary to reduce the plaque size by

713

50% (plaque assay IC50) for N1-3-VHHb and N1-5-VHHb was 7.6 nM (243 ng/ml) or 14.5 nM

714

(464 ng/ml) respectively, which are 240- and 58-fold lower than their monovalent formats. In

32

715

fact, we noticed a consistent 30-fold or higher inhibitory activity of bivalent N1-3-VHHb and N1-

716

3-VHH-Fc compared to monovalent VHHs in a biochemical assay using N1rec and against the

717

three H5N1 viruses tested in vitro. H5N1 H274Y was an exception in that N1-5-VHH mono or

718

bivalent formats failed to inhibit this Oseltamivir resistant virus (Table 2). These results also

719

suggest that the epitope recognized by N1-3-VHHm and N1-5-VHHm is conserved between the

720

NAs from clade 1 and 2 H5N1 viruses.

721

Intranasal administration of a relatively high dose of N1-3-VHHb or N1-5-VHHb (60 μg; 3.3

722

mg/kg) 24 h before H5N1 challenge protected mice against death, but significant body weight

723

was still lost (Fig. 5). In this challenge model, the N1-VHH-Fc formats protected better than the

724

N1-VHHb formats. A single dose of 17 μg (1.1 mg/kg, N1-3-VHH-Fc) or 3.5 μg (0.23 mg/kg, N1-

725

5-VHH-Fc) resulted in 100% survival of mice, and reduced morbidity compared to control

726

treated animals (Fig. 6 and 7). When comparing the molar ratios of N1-VHH-Fc (N1-3-VHH-Fc

727

and N1-5-VHH-Fc) with N1-VHHb (N1-3-VHHb or N1-5-VHHb), we found that approximately 10

728

to 50 times lower amounts of the N1-VHH-Fc formats were necessary to give 100% survival to

729

H5N1 challenged mice. For comparison, also in a mouse model, conventional mAb 2B9 resulted

730

in only partial (50%) protection against H5N1 A/Vietnam/1194/2004 challenge (74). Even so,

731

this required intravenous administration of 500 μg (33.3 mg/kg) of 2B9 1 h before 10 LD50 of

732

HPAIV H5N1, with daily boosts for 5 days. Furthermore, when we compared the 2B9 treatment

733

with our single-dose treatment with N1-VHHb or N1-VHH-Fc, 10- or 150-fold lower amount of

734

mAb, respectively, was required to rescue 100% of H5N1 challenged mice without any boost.

735

When comparing with a study by Wan et al., the lowest prophylactic single-dose of a

736

conventional IgG2a mAb necessary to fully rescue mice from a 10 LD50 H5N1 challenge was 15

737

mg/kg, whereas 0.23 mg/kg of our N1-5-VHH-Fc protected all mice from death after a 4 LD50

738

challenge

739

A/Vietnam/1194/2004) is almost identical to that used in the Wan report (A/Vietnam/1203/2004).

with

NIBRG-14

ma

(67).

The

target

NA

of

NIBRG-14

(derived

from

33

740

In both reports on NA mAbs discussed above (67, 74), the route of administration (intravenous

741

or intraperitoneal) could have been an important factor and it might help explain the apparent

742

superior protection of our N1-VHH formats, which we administered intranasally, since VHHs

743

withstand nebulization. Moreover, it has been shown that anti-RSV Nanobodies give extended

744

protection against viral infection in vivo when administered via a nebulizer. It may be expected

745

that optimization of the delivery will further help to improve N1-VHH efficacy. N1-VHHm

746

treatments did not significantly reduce the virus titers or viral genomic loads in lung

747

homogenates 4 days after infection (Fig. 4C). The weak, direct antiviral activity of our NA-

748

specific VHHs in the in vivo model is in line with the paradigm that NA Abs or NA-inhibitory

749

drugs do not prevent infection.

750

The observation that N1-5-VHHm, N1-5-VHHb and N1-5-VHH-Fc did not have any antiviral

751

activity against the Oseltamivir-resistant H5N1 H274Y virus in vitro was surprising since in the

752

SPR assay N1-3-VHHm and N1-5-VHHm bound to overlapping epitopes in NA or sterically

753

hindered each other’s binding (Fig. 1F). This discrepancy was also observed in vitro for N1-7-

754

VHHm, whereas N1-7-VHH-Fc did protect mice against challenge with H5N1 H274Y virus (Fig.

755

7). The latter observation indicates that protection by NA-based antibodies does not require the

756

NAI activity of these antibodies, provided that an Fc domain is present to contribute Fc

757

Receptor-dependent effector functions. The finding that also N1-7-VHH-Fc had in vivo antiviral

758

activity despite lack of NAI activity is in line with this (Fig. 7B and 7D). Mouse IgG2a is one of

759

the most potent Ab isotypes to induce antibody-dependent cellular cytotoxicity or complement-

760

dependent cellular cytotoxicity, due to its high affinity for FcγRIII and –IV and especially for

761

FcγRI, which binds exclusively to IgG2a. IgG2a has an intermediate affinity for FcγRIV that is

762

one order of magnitude higher than its affinity for the inhibitor receptor FcγRIIB (75).

763

The VHH CDR3 is a major region responsible for the contact of VHH with antigen (25, 76). All

764

NA inhibitory N1-VHHs reported here (N1-1-VHHm, N1-3-VHHm, N1-5-VHHm and N1-6-VHHm)

34

765

have a predicted electronegatively charged protruding paratope that corresponds to their CDR3

766

(Fig. 1C and 1D). The CDR3 sequences of N1-3-VHHm (HGXXHGHGXHX, 10 residues) and

767

N1-5-VHHm (HXHXHHXHXH, 12 residues), where "H" is a hydrophobic residue and "X" is an

768

aspartic or glutamic acid, might be involved in the direct contact with their epitope or epitopes.

769

When comparing N1-3-VHHm and N1-5-VHHm CDR3 sequences with the most potent NA

770

inhibitory VHHs reported by Harmsen et al., we found similarities in their CDR3 sequences. For

771

example, the CDR3 sequences of one N8-binding VHH (UHXHHHGHGUHHXUPHXH, 18

772

residues) and one N2-binding VHH (HHXHGHHGGHGHUHXH, 16 residues), where “U” is a

773

polar uncharged residue, shared similarities with the N1-3-VHHm and N1-5-VHHm CDR3s.

774

The available crystal structures of mAbs in complex with NA are all for the group 2 NA,

775

specifically N2 and N9 (77-80). These structures showed that these mAb epitopes rarely include

776

conserved amino acid residues from the catalytic site; rather, they target amino acids in loops

777

surrounding the NA catalytic site, also described as the major antigenic regions of NA. It seems

778

that the N1-5-VHHm epitope is partially shared with other N1 NA molecules, as the pandemic

779

H1N1 NA activity inhibition suggests (Fig. 3).

780

Plant-produced Abs have been considered a promising and effective alternative for Ab

781

production by recombinant yeast or mammalian cell culture systems. Our N1-VHH-Fc is the first

782

reported plant-produced Ab against an influenza antigen. Their scalability, robustness and

783

speed of production are major advantages, even though the glycosylation patterns of the plant-

784

produced Abs (rich in high mannose residues) are different from those of the Abs produced in

785

mammalian expression platforms (81). The use of plants for Ab production also presents an

786

interesting approach for developing countries: long storage life, relative ease of purification or

787

delivery, and use of established agricultural techniques are very attractive. To our knowledge,

788

only three single domain Abs produced in plants have been reported as stably transformed lines

789

(82-84), but their yields were very low. Results from our lab demonstrated that the Fc moiety

35

790

significantly improved VHH accumulation in seeds (compared to the monomeric VHH alone)

791

with expression levels of 10-20% of total seed proteins (54, 85).

792

HPAIV H5N1 remains a public health threat. The experimental adaptation of H5N1 virus for

793

airborne transmission in ferrets, with mutations in HA and PB2, warns of the possibility of

794

human-to-human transmission of zoonotic HPAIV (1, 2). Should the HPAIV H5N1 become

795

transmissible among humans, NA targeting could be part of an effective strategy to mitigate

796

disease caused by such a virus. In addition, an interesting prophylactic and therapeutic

797

approach that deserves further validation is the combination of VHH targeting IAV NA, HA or M2

798

proteins in cocktails or biparatopic bivalent VHH constructs that could diminish the antigenic drift

799

in IAV, and lead to much higher selective pressure against IAV VHH.

36

800

Acknowledgements

801

The authors gratefully appreciate the excellent technical support of Tine Ysenbaert, Frederick

802

Vervallle and Jonah Nolf. We thank Dr. Thierry Van den Bergh from the Veterinary and

803

Agrochemical Research Centre in Brussels (Belgium) for providing cDNA of Neuraminidase of

804

A/crested eagle/Belgium/01/2004 (H5N1) influenza virus, Dr. Wolgang Garten from Philipps-

805

Universität Marburg, Germany for MDCK TMPRSS2 cells, Dr. John Wood from the UK National

806

Institute for Biological Standards and Control for providing NIBRG-14 and -23 virus, Dr. Alan

807

Hay (MRC National Institute for Medical Research, Mill Hill, London, UK) for providing A/New

808

Caledonia/20/99 virus, and St. Jude Children's Research Hospital for providing the A/PR/8/34

809

based eight-plasmid system for generating recombinant influenza A viruses. FMC was funded

810

by a pre-doctoral grant from the Biotechnology Institute of Flanders (VIB) and supported by

811

IUAP BELVIR project p7/45.

812

37

813

REFERENCES

814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855

1.

2.

3.

4. 5.

6. 7.

8. 9. 10.

Herfst, S., E. J. Schrauwen, M. Linster, S. Chutinimitkul, E. de Wit, V. J. Munster, E. M. Sorrell, T. M. Bestebroer, D. F. Burke, D. J. Smith, G. F. Rimmelzwaan, A. D. Osterhaus, and R. A. Fouchier. 2012. Airborne transmission of influenza A/H5N1 virus between ferrets. Science 336:1534-1541. Imai, M., T. Watanabe, M. Hatta, S. C. Das, M. Ozawa, K. Shinya, G. Zhong, A. Hanson, H. Katsura, S. Watanabe, C. Li, E. Kawakami, S. Yamada, M. Kiso, Y. Suzuki, E. A. Maher, G. Neumann, and Y. Kawaoka. 2012. Experimental adaptation of an influenza H5 HA confers respiratory droplet transmission to a reassortant H5 HA/H1N1 virus in ferrets. Nature 486:420-428. Tran, T. H., T. L. Nguyen, T. D. Nguyen, T. S. Luong, P. M. Pham, V. Nguyen v, T. S. Pham, C. D. Vo, T. Q. Le, T. T. Ngo, B. K. Dao, P. P. Le, T. T. Nguyen, T. L. Hoang, V. T. Cao, T. G. Le, D. T. Nguyen, H. N. Le, K. T. Nguyen, H. S. Le, V. T. Le, D. Christiane, T. T. Tran, J. Menno de, C. Schultsz, P. Cheng, W. Lim, P. Horby, J. Farrar, and T. World Health Organization International Avian Influenza Investigative. 2004. Avian influenza A (H5N1) in 10 patients in Vietnam. The New England journal of medicine 350:1179-1188. Trifonov, V., H. Khiabanian, and R. Rabadan. 2009. Geographic dependence, surveillance, and origins of the 2009 influenza A (H1N1) virus. The New England journal of medicine 361:115-119. Gao, H. N., H. Z. Lu, B. Cao, B. Du, H. Shang, J. H. Gan, S. H. Lu, Y. D. Yang, Q. Fang, Y. Z. Shen, X. M. Xi, Q. Gu, X. M. Zhou, H. P. Qu, Z. Yan, F. M. Li, W. Zhao, Z. C. Gao, G. F. Wang, L. X. Ruan, W. H. Wang, J. Ye, H. F. Cao, X. W. Li, W. H. Zhang, X. C. Fang, J. He, W. F. Liang, J. Xie, M. Zeng, X. Z. Wu, J. Li, Q. Xia, Z. C. Jin, Q. Chen, C. Tang, Z. Y. Zhang, B. M. Hou, Z. X. Feng, J. F. Sheng, N. S. Zhong, and L. J. Li. 2013. Clinical findings in 111 cases of influenza A (H7N9) virus infection. The New England journal of medicine 368:2277-2285. Roose, K., W. Fiers, and X. Saelens. 2009. Pandemic preparedness: toward a universal influenza vaccine. Drug news & perspectives 22:80-92. de Jong, M. D., C. P. Simmons, T. T. Thanh, V. M. Hien, G. J. Smith, T. N. Chau, D. M. Hoang, N. V. Chau, T. H. Khanh, V. C. Dong, P. T. Qui, B. V. Cam, Q. Ha do, Y. Guan, J. S. Peiris, N. T. Chinh, T. T. Hien, and J. Farrar. 2006. Fatal outcome of human influenza A (H5N1) is associated with high viral load and hypercytokinemia. Nature medicine 12:1203-1207. Matrosovich, M. N., T. Y. Matrosovich, T. Gray, N. A. Roberts, and H. D. Klenk. 2004. Neuraminidase is important for the initiation of influenza virus infection in human airway epithelium. Journal of virology 78:12665-12667. Johansson, B. E., J. T. Matthews, and E. D. Kilbourne. 1998. Supplementation of conventional influenza A vaccine with purified viral neuraminidase results in a balanced and broadened immune response. Vaccine 16:1009-1015. Kilbourne, E. D., B. A. Pokorny, B. Johansson, I. Brett, Y. Milev, and J. T. Matthews. 2004. Protection of mice with recombinant influenza virus neuraminidase. The Journal of infectious diseases 189:459-461.

38

856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900

11. 12. 13. 14.

15. 16. 17.

18. 19.

20.

21. 22. 23. 24.

Marcelin, G., M. R. Sandbulte, and R. J. Webby. 2012. Contribution of antibody production against neuraminidase to the protection afforded by influenza vaccines. Reviews in medical virology 22:267-279. Goto, H., and Y. Kawaoka. 1998. A novel mechanism for the acquisition of virulence by a human influenza A virus. Proceedings of the National Academy of Sciences of the United States of America 95:10224-10228. Li, S., J. Schulman, S. Itamura, and P. Palese. 1993. Glycosylation of neuraminidase determines the neurovirulence of influenza A/WSN/33 virus. Journal of virology 67:6667-6673. Matsuoka, Y., D. E. Swayne, C. Thomas, M. A. Rameix-Welti, N. Naffakh, C. Warnes, M. Altholtz, R. Donis, and K. Subbarao. 2009. Neuraminidase stalk length and additional glycosylation of the hemagglutinin influence the virulence of influenza H5N1 viruses for mice. Journal of virology 83:4704-4708. Huang, I. C., W. Li, J. Sui, W. Marasco, H. Choe, and M. Farzan. 2008. Influenza A virus neuraminidase limits viral superinfection. Journal of virology 82:4834-4843. Ilyushina, N. A., N. V. Bovin, and R. G. Webster. 2012. Decreased neuraminidase activity is important for the adaptation of H5N1 influenza virus to human airway epithelium. Journal of virology 86:4724-4733. Takahashi, Y., H. Hasegawa, Y. Hara, M. Ato, A. Ninomiya, H. Takagi, T. Odagiri, T. Sata, M. Tashiro, and K. Kobayashi. 2009. Protective immunity afforded by inactivated H5N1 (NIBRG-14) vaccine requires antibodies against both hemagglutinin and neuraminidase in mice. The Journal of infectious diseases 199:1629-1637. Gubareva, L. V., M. S. Nedyalkova, D. V. Novikov, K. G. Murti, E. Hoffmann, and F. G. Hayden. 2002. A release-competent influenza A virus mutant lacking the coding capacity for the neuraminidase active site. The Journal of general virology 83:2683-2692. Mitnaul, L. J., M. N. Matrosovich, M. R. Castrucci, A. B. Tuzikov, N. V. Bovin, D. Kobasa, and Y. Kawaoka. 2000. Balanced hemagglutinin and neuraminidase activities are critical for efficient replication of influenza A virus. Journal of virology 74:60156020. Nedyalkova, M. S., F. G. Hayden, R. G. Webster, and L. V. Gubareva. 2002. Accumulation of defective neuraminidase (NA) genes by influenza A viruses in the presence of NA inhibitors as a marker of reduced dependence on NA. The Journal of infectious diseases 185:591-598. Murphy, B. R., J. A. Kasel, and R. M. Chanock. 1972. Association of serum antineuraminidase antibody with resistance to influenza in man. The New England journal of medicine 286:1329-1332. Vanlandschoot, P., C. Stortelers, E. Beirnaert, L. I. Ibanez, B. Schepens, E. Depla, and X. Saelens. 2011. Nanobodies(R): new ammunition to battle viruses. Antiviral research 92:389-407. Greenberg, A. S., D. Avila, M. Hughes, A. Hughes, E. C. McKinney, and M. F. Flajnik. 1995. A new antigen receptor gene family that undergoes rearrangement and extensive somatic diversification in sharks. Nature 374:168-173. Hamers-Casterman, C., T. Atarhouch, S. Muyldermans, G. Robinson, C. Hamers, E. B. Songa, N. Bendahman, and R. Hamers. 1993. Naturally occurring antibodies devoid of light chains. Nature 363:446-448.

39

901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946

25.

26. 27.

28. 29. 30.

31. 32.

33.

34.

35.

36.

Coppieters, K., T. Dreier, K. Silence, H. de Haard, M. Lauwereys, P. Casteels, E. Beirnaert, H. Jonckheere, C. Van de Wiele, L. Staelens, J. Hostens, H. Revets, E. Remaut, D. Elewaut, and P. Rottiers. 2006. Formatted anti-tumor necrosis factor alpha VHH proteins derived from camelids show superior potency and targeting to inflamed joints in a murine model of collagen-induced arthritis. Arthritis and rheumatism 54:18561866. Saerens, D., G. H. Ghassabeh, and S. Muyldermans. 2008. Single-domain antibodies as building blocks for novel therapeutics. Current opinion in pharmacology 8:600-608. Decanniere, K., A. Desmyter, M. Lauwereys, M. A. Ghahroudi, S. Muyldermans, and L. Wyns. 1999. A single-domain antibody fragment in complex with RNase A: noncanonical loop structures and nanomolar affinity using two CDR loops. Structure 7:361370. Decanniere, K., S. Muyldermans, and L. Wyns. 2000. Canonical antigen-binding loop structures in immunoglobulins: more structures, more canonical classes? Journal of molecular biology 300:83-91. Ward, E. S., D. Gussow, A. D. Griffiths, P. T. Jones, and G. Winter. 1989. Binding activities of a repertoire of single immunoglobulin variable domains secreted from Escherichia coli. Nature 341:544-546. Desmyter, A., S. Spinelli, F. Payan, M. Lauwereys, L. Wyns, S. Muyldermans, and C. Cambillau. 2002. Three camelid VHH domains in complex with porcine pancreatic alpha-amylase. Inhibition and versatility of binding topology. The Journal of biological chemistry 277:23645-23650. Lauwereys, M., M. Arbabi Ghahroudi, A. Desmyter, J. Kinne, W. Holzer, E. De Genst, L. Wyns, and S. Muyldermans. 1998. Potent enzyme inhibitors derived from dromedary heavy-chain antibodies. The EMBO journal 17:3512-3520. Rasmussen, S. G., H. J. Choi, J. J. Fung, E. Pardon, P. Casarosa, P. S. Chae, B. T. Devree, D. M. Rosenbaum, F. S. Thian, T. S. Kobilka, A. Schnapp, I. Konetzki, R. K. Sunahara, S. H. Gellman, A. Pautsch, J. Steyaert, W. I. Weis, and B. K. Kobilka. 2011. Structure of a nanobody-stabilized active state of the beta(2) adrenoceptor. Nature 469:175-180. De Genst, E., K. Silence, K. Decanniere, K. Conrath, R. Loris, J. Kinne, S. Muyldermans, and L. Wyns. 2006. Molecular basis for the preferential cleft recognition by dromedary heavy-chain antibodies. Proceedings of the National Academy of Sciences of the United States of America 103:4586-4591. Hultberg, A., N. J. Temperton, V. Rosseels, M. Koenders, M. Gonzalez-Pajuelo, B. Schepens, L. I. Ibanez, P. Vanlandschoot, J. Schillemans, M. Saunders, R. A. Weiss, X. Saelens, J. A. Melero, C. T. Verrips, S. Van Gucht, and H. J. de Haard. 2011. Llama-derived single domain antibodies to build multivalent, superpotent and broadened neutralizing anti-viral molecules. PloS one 6:e17665. Ibanez, L. I., M. De Filette, A. Hultberg, T. Verrips, N. Temperton, R. A. Weiss, W. Vandevelde, B. Schepens, P. Vanlandschoot, and X. Saelens. 2011. Nanobodies with in vitro neutralizing activity protect mice against H5N1 influenza virus infection. The Journal of infectious diseases 203:1063-1072. Schepens, B., L. I. Ibanez, S. De Baets, A. Hultberg, P. Bogaert, P. De Bleser, F. Vervalle, T. Verrips, J. Melero, W. Vandevelde, P. Vanlandschoot, and X. Saelens. 2011. Nanobodies(R) specific for respiratory syncytial virus fusion protein protect 40

947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992

37.

38. 39. 40.

41. 42. 43. 44.

45. 46.

47. 48. 49. 50.

against infection by inhibition of fusion. The Journal of infectious diseases 204:16921701. Tillib, S. V., T. I. Ivanova, L. A. Vasilev, M. V. Rutovskaya, S. A. Saakyan, I. Y. Gribova, I. L. Tutykhina, E. S. Sedova, A. A. Lysenko, M. M. Shmarov, D. Y. Logunov, B. S. Naroditsky, and A. L. Gintsburg. 2013. Formatted single-domain antibodies can protect mice against infection with influenza virus (H5N2). Antiviral research 97:245-254. Wei, G., W. Meng, H. Guo, W. Pan, J. Liu, T. Peng, L. Chen, and C. Y. Chen. 2011. Potent neutralization of influenza A virus by a single-domain antibody blocking M2 ion channel protein. PloS one 6:e28309. Harmsen M, B. J., Pritz-Verschuren S, Bartenlink W, Van der Burg H and Koch G. . 2013. Isolation of panels of llama single-domain antibody fragments binding all nine neuraminidase subtypes of influenza A virus. Antibodies 2:168-192. Van Borm, S., I. Thomas, G. Hanquet, B. Lambrecht, M. Boschmans, G. Dupont, M. Decaestecker, R. Snacken, and T. van den Berg. 2005. Highly pathogenic H5N1 influenza virus in smuggled Thai eagles, Belgium. Emerging infectious diseases 11:702705. Hoffmann, E., S. Krauss, D. Perez, R. Webby, and R. G. Webster. 2002. Eightplasmid system for rapid generation of influenza virus vaccines. Vaccine 20:3165-3170. Schotsaert, M., T. Ysenbaert, K. Neyt, L. I. Ibanez, P. Bogaert, B. Schepens, B. N. Lambrecht, W. Fiers, and X. Saelens. 2013. Natural and long-lasting cellular immune responses against influenza in the M2e-immune host. Mucosal immunology 6:276-287. H, R. L. a. M. 1938. A simple method for estimating fifty percent endpoints. American Journal of Medical Hygiene 27:493-497. De Filette, M., W. Martens, K. Roose, T. Deroo, F. Vervalle, M. Bentahir, J. Vandekerckhove, W. Fiers, and X. Saelens. 2008. An influenza A vaccine based on tetrameric ectodomain of matrix protein 2. The Journal of biological chemistry 283:11382-11387. Harbury, P. B., T. Zhang, P. S. Kim, and T. Alber. 1993. A switch between two-, three-, and four-stranded coiled coils in GCN4 leucine zipper mutants. Science 262:14011407. Hassantoufighi, A., H. Zhang, M. Sandbulte, J. Gao, J. Manischewitz, L. King, H. Golding, T. M. Straight, and M. C. Eichelberger. 2010. A practical influenza neutralization assay to simultaneously quantify hemagglutinin and neuraminidaseinhibiting antibody responses. Vaccine 28:790-797. Arbabi Ghahroudi, M., A. Desmyter, L. Wyns, R. Hamers, and S. Muyldermans. 1997. Selection and identification of single domain antibody fragments from camel heavy-chain antibodies. FEBS letters 414:521-526. Kang, A. S., T. M. Jones, and D. R. Burton. 1991. Antibody redesign by chain shuffling from random combinatorial immunoglobulin libraries. Proceedings of the National Academy of Sciences of the United States of America 88:11120-11123. De Meyer, T., D. Eeckhout, R. De Rycke, S. De Buck, S. Muyldermans, and A. Depicker. 2013. Generation of VHH antibodies against the Arabidopsis thaliana seed storage proteins. Plant molecular biology. Lambert, C., N. Leonard, X. De Bolle, and E. Depiereux. 2002. ESyPred3D: Prediction of proteins 3D structures. Bioinformatics 18:1250-1256. 41

993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038

51. 52.

53. 54. 55.

56. 57.

58.

59.

60. 61.

62.

Conrath, E. K., M. Lauwereys, L. Wyns, and S. Muyldermans. 2001. Camel singledomain antibodies as modular building units in bispecific and bivalent antibody constructs. The Journal of biological chemistry 276:7346-7350. Conrath, K. E., M. Lauwereys, M. Galleni, A. Matagne, J. M. Frere, J. Kinne, L. Wyns, and S. Muyldermans. 2001. Beta-lactamase inhibitors derived from singledomain antibody fragments elicited in the camelidae. Antimicrobial agents and chemotherapy 45:2807-2812. Clough, S. J., and A. F. Bent. 1998. Floral dip: a simplified method for Agrobacteriummediated transformation of Arabidopsis thaliana. The Plant journal : for cell and molecular biology 16:735-743. De Buck, S., V. Virdi, T. De Meyer, K. De Wilde, R. Piron, J. Nolf, E. Van Lerberge, A. De Paepe, and A. Depicker. 2012. Production of camel-like antibodies in plants. Methods in molecular biology 911:305-324. Bottcher, E., C. Freuer, T. Steinmetzer, H. D. Klenk, and W. Garten. 2009. MDCK cells that express proteases TMPRSS2 and HAT provide a cell system to propagate influenza viruses in the absence of trypsin and to study cleavage of HA and its inhibition. Vaccine 27:6324-6329. Matrosovich, M., T. Matrosovich, W. Garten, and H. D. Klenk. 2006. New lowviscosity overlay medium for viral plaque assays. Virology journal 3:63. Stech, J., O. Stech, A. Herwig, H. Altmeppen, J. Hundt, S. Gohrbandt, A. Kreibich, S. Weber, H. D. Klenk, and T. C. Mettenleiter. 2008. Rapid and reliable universal cloning of influenza A virus genes by target-primed plasmid amplification. Nucleic acids research 36:e139. De Jaeger, G., S. Scheffer, A. Jacobs, M. Zambre, O. Zobell, A. Goossens, A. Depicker, and G. Angenon. 2002. Boosting heterologous protein production in transgenic dicotyledonous seeds using Phaseolus vulgaris regulatory sequences. Nature biotechnology 20:1265-1268. Van Droogenbroeck, B., J. Cao, J. Stadlmann, F. Altmann, S. Colanesi, S. Hillmer, D. G. Robinson, E. Van Lerberge, N. Terryn, M. Van Montagu, M. Liang, A. Depicker, and G. De Jaeger. 2007. Aberrant localization and underglycosylation of highly accumulating single-chain Fv-Fc antibodies in transgenic Arabidopsis seeds. Proceedings of the National Academy of Sciences of the United States of America 104:1430-1435. Group, W. O. F. H. N. E. W. 2012. Continued evolution of highly pathogenic avian influenza A (H5N1): updated nomenclature. Influenza and other respiratory viruses 6:15. Kandun, I. N., H. Wibisono, E. R. Sedyaningsih, Yusharmen, W. Hadisoedarsuno, W. Purba, H. Santoso, C. Septiawati, E. Tresnaningsih, B. Heriyanto, D. Yuwono, S. Harun, S. Soeroso, S. Giriputra, P. J. Blair, A. Jeremijenko, H. Kosasih, S. D. Putnam, G. Samaan, M. Silitonga, K. H. Chan, L. L. Poon, W. Lim, A. Klimov, S. Lindstrom, Y. Guan, R. Donis, J. Katz, N. Cox, M. Peiris, and T. M. Uyeki. 2006. Three Indonesian clusters of H5N1 virus infection in 2005. The New England journal of medicine 355:2186-2194. Wang, M. Z., C. Y. Tai, and D. B. Mendel. 2002. Mechanism by which mutations at his274 alter sensitivity of influenza a virus n1 neuraminidase to oseltamivir carboxylate and zanamivir. Antimicrobial agents and chemotherapy 46:3809-3816. 42

1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083

63. 64.

65.

66.

67.

68.

69.

70. 71.

72. 73. 74.

Brandes, M., F. Klauschen, S. Kuchen, and R. N. Germain. 2013. A systems analysis identifies a feedforward inflammatory circuit leading to lethal influenza infection. Cell 154:197-212. Govorkova, E. A., I. A. Leneva, O. G. Goloubeva, K. Bush, and R. G. Webster. 2001. Comparison of efficacies of RWJ-270201, zanamivir, and oseltamivir against H5N1, H9N2, and other avian influenza viruses. Antimicrobial agents and chemotherapy 45:2723-2732. Doyle, T. M., B. Jaentschke, G. Van Domselaar, A. M. Hashem, A. Farnsworth, N. E. Forbes, C. Li, J. Wang, R. He, E. G. Brown, and X. Li. 2013. The universal epitope of influenza A viral neuraminidase fundamentally contributes to enzyme activity and viral replication. The Journal of biological chemistry 288:18283-18289. Gravel, C., C. Li, J. Wang, A. M. Hashem, B. Jaentschke, K. W. Xu, B. Lorbetskie, G. Gingras, Y. Aubin, G. Van Domselaar, M. Girard, R. He, and X. Li. 2010. Qualitative and quantitative analyses of virtually all subtypes of influenza A and B viral neuraminidases using antibodies targeting the universally conserved sequences. Vaccine 28:5774-5784. Wan, H., J. Gao, K. Xu, H. Chen, L. K. Couzens, K. H. Rivers, J. D. Easterbrook, K. Yang, L. Zhong, M. Rajabi, J. Ye, I. Sultana, X. F. Wan, X. Liu, D. R. Perez, J. K. Taubenberger, and M. C. Eichelberger. 2013. Molecular basis for broad neuraminidase immunity: conserved epitopes in seasonal and pandemic H1N1 as well as H5N1 influenza viruses. Journal of virology 87:9290-9300. McKimm-Breschkin, J. L., T. J. Blick, A. Sahasrabudhe, T. Tiong, D. Marshall, G. J. Hart, R. C. Bethell, and C. R. Penn. 1996. Generation and characterization of variants of NWS/G70C influenza virus after in vitro passage in 4-amino-Neu5Ac2en and 4-guanidino-Neu5Ac2en. Antimicrobial agents and chemotherapy 40:40-46. Johansson, B. E., T. M. Moran, C. A. Bona, S. W. Popple, and E. D. Kilbourne. 1987. Immunologic response to influenza virus neuraminidase is influenced by prior experience with the associated viral hemagglutinin. II. Sequential infection of mice simulates human experience. Journal of immunology 139:2010-2014. Schulman, J. L., M. Khakpour, and E. D. Kilbourne. 1968. Protective effects of specific immunity to viral neuraminidase on influenza virus infection of mice. Journal of virology 2:778-786. Cate, T. R., Y. Rayford, D. Nino, P. Winokur, R. Brady, R. Belshe, W. Chen, R. L. Atmar, and R. B. Couch. 2010. A high dosage influenza vaccine induced significantly more neuraminidase antibody than standard vaccine among elderly subjects. Vaccine 28:2076-2079. Air, G. M. 2012. Influenza neuraminidase. Influenza and other respiratory viruses 6:245256. Hensley, S. E., S. R. Das, J. S. Gibbs, A. L. Bailey, L. M. Schmidt, J. R. Bennink, and J. W. Yewdell. 2011. Influenza A virus hemagglutinin antibody escape promotes neuraminidase antigenic variation and drug resistance. PloS one 6:e15190. Shoji, Y., J. A. Chichester, G. A. Palmer, C. E. Farrance, R. Stevens, M. Stewart, L. Goldschmidt, V. Deyde, L. Gubareva, A. Klimov, V. Mett, and V. Yusibov. 2011. An influenza N1 neuraminidase-specific monoclonal antibody with broad neuraminidase inhibition activity against H5N1 HPAI viruses. Human vaccines 7 Suppl:199-204.

43

1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119

75. 76.

77. 78. 79.

80. 81. 82.

83. 84. 85.

Nimmerjahn, F., and J. V. Ravetch. 2008. Fcgamma receptors as regulators of immune responses. Nature reviews. Immunology 8:34-47. Desmyter, A., T. R. Transue, M. A. Ghahroudi, M. H. Thi, F. Poortmans, R. Hamers, S. Muyldermans, and L. Wyns. 1996. Crystal structure of a camel singledomain VH antibody fragment in complex with lysozyme. Nature structural biology 3:803-811. Colman, P. M., W. G. Laver, J. N. Varghese, A. T. Baker, P. A. Tulloch, G. M. Air, and R. G. Webster. 1987. Three-dimensional structure of a complex of antibody with influenza virus neuraminidase. Nature 326:358-363. Malby, R. L., A. J. McCoy, A. A. Kortt, P. J. Hudson, and P. M. Colman. 1998. Three-dimensional structures of single-chain Fv-neuraminidase complexes. Journal of molecular biology 279:901-910. Malby, R. L., W. R. Tulip, V. R. Harley, J. L. McKimm-Breschkin, W. G. Laver, R. G. Webster, and P. M. Colman. 1994. The structure of a complex between the NC10 antibody and influenza virus neuraminidase and comparison with the overlapping binding site of the NC41 antibody. Structure 2:733-746. Tulip, W. R., J. N. Varghese, W. G. Laver, R. G. Webster, and P. M. Colman. 1992. Refined crystal structure of the influenza virus N9 neuraminidase-NC41 Fab complex. Journal of molecular biology 227:122-148. De Muynck, B., C. Navarre, and M. Boutry. 2010. Production of antibodies in plants: status after twenty years. Plant biotechnology journal 8:529-563. Ismaili, A., M. Jalali-Javaran, M. J. Rasaee, F. Rahbarizadeh, M. ForouzandehMoghadam, and H. R. Memari. 2007. Production and characterization of anti-(mucin MUC1) single-domain antibody in tobacco (Nicotiana tabacum cultivar Xanthi). Biotechnology and applied biochemistry 47:11-19. Jobling, S. A., C. Jarman, M. M. Teh, N. Holmberg, C. Blake, and M. E. Verhoeyen. 2003. Immunomodulation of enzyme function in plants by single-domain antibody fragments. Nature biotechnology 21:77-80. Winichayakul, S., A. Pernthaner, R. Scott, R. Vlaming, and N. Roberts. 2009. Headto-tail fusions of camelid antibodies can be expressed in planta and bind in rumen fluid. Biotechnology and applied biochemistry 53:111-122. De Buck, S., J. Nolf, T. De Meyer, V. Virdi, K. De Wilde, E. Van Lerberge, B. Van Droogenbroeck, and A. Depicker. 2013. Fusion of an Fc chain to a VHH boosts the accumulation levels in Arabidopsis seeds. Plant biotechnology journal 11:1006-1016.

1120

44

1121

Figure 1. Characterization of monovalent N1-VHH candidates that bind recombinant H5N1-

1122

derived neuraminidase. A. Soluble N1rec derived from H5N1 A/crested eagle/Belgium/01/2004

1123

expressed in Sf9 cells. A diagram of the n1rec Baculovirus expression cassette (top panel);

1124

promoter, Baculovirus basic protein promoter; ssHA, secretion signal of HA; tGCN4,

1125

tetramerizing GCN4-derived leucine zipper; H5N1 NA, extracellular domain of NA (amino acid

1126

residues 53–449). The lower panel shows NA activity in the culture supernatants of different

1127

amounts of AcNPVN1rec-infected cells or mock-infected cells, using MUNANA as a substrate.

1128

RFU: relative fluorescent units. B. Schematic representation of the expression cassettes of

1129

N1rec-binding VHH domains and derived bivalent formats. Bacteria-produced monovalent N1-

1130

VHHm (n1-vhhm), with the amino-terminal periplasmic signal sequence (pelB), the n1-vhhm

1131

coding information, and a carboxy-terminal hexahistidine tag (his)(top panel). Bacteria-produced

1132

bivalent N1-VHHb with two n1-vhhm moieties fused by a flexible linker (llama IgG2c hinge) and

1133

a hexahistidine tag (his)(middle panel). For expression in transgenic Arabidopsis thaliana seeds,

1134

the N1-VHH-Fc insert was cloned into the seed-specific expression cassette of the PphasGW

1135

vector. LB (left border of T-DNA); 3’ ocs (3’ end of the octopine synthase gene); npt II (neomycin

1136

phosphotransferase II open reading frame); Pnos (nopaline synthase gene promoter); Pphas (β-

1137

phaseolin gene promoter); 5’ utr (5’ UTR of arc5-I gene); SS (signal sequence of the

1138

Arabidopsis thaliana 2S2 seed storage protein gene); n1-vhh-fc (coding sequence of the N1-

1139

VHH gene fused to mouse CH2 and CH3 IgG2a); KDEL (endoplasmic reticulum retention

1140

signal); 3’ arc (3’ flanking regulatory sequences of the arc5-I gene); RB, T-DNA right border

1141

(lower panel). C. CDR3 sequences of 13 N1rec-binding VHH candidates. The VHH CDR3 is

1142

highlighted by a red dotted box. The colored histogram denotes the homology between the

1143

sequences: red, 100%; orange, 60–80%; green, 40–60 %; light blue, 20–40 %; dark blue,
15448

-

>32896

-

>15448

-

>32896

-

>32896

123 -

0.157 (+/0.206) 0.69 (+/0.231) >15448 1.31 (+/0.605) 1.49 (+/0.746)

2708

7.6 (+/- 3.0)

240

52.2 (+/36.8)

543

14.5 (+/- 11.1)

58

>707.5 >32896 89.4 (+/7.8)

c

Fold

4295 (+/2835.4) 3492 (+/2967.6)

-

>32896

-

>32896

-

24.6 (+/31.9) 23.6 (+/3.8) >32896 24.2 (+/3.7)

174 148

-

>32896

-

324

23.08

79

251

28.14

30

>17083

-

27.03

129

>7900

-

>5260

-

>5260

-

>5260

-

>7900 586.1 (+/120.5)

-

>5260 15030 (+/10039)

-

>5260

-

>5260

-

-

>243000

-

73000

-

-

-

b

NIBRG-23

71

177

Mean inhibitory concentration (IC50) from three independent neuraminidase inhibition

assays using 2’-(4-methylumbelliferyl)-α-D-N-acetylneuraminic acid (MUNANA) and 160 ng of N1rec. b

Mean plaque inhibitory concentration (IC50) from two independent plaque assays performed

in duplicate, in which the VHH concentration reduced the plaque size by 50% compared with mock VHH. c

Fold increase inhibitory potency of the N1-VHH bivalent format compared to its monovalent

format. d

Monomeric VHH against seed storage albumin.

e

Bivalent VHH against bacterial beta lactamase.

f

Coronavirus GP4 protein fused to mouse IgG2a Fc

Table 3. Endotoxin levels in the purified VHH preparations VHH N1-3-VHHm N1-5-VHHm N1-7-VHHm SA-VHHm N1-3-VHHb N1-5-VHHb BL-VHHb N1-3-VHH-Fc N1-5-VHH-Fc N1-7-VHH-Fc GP4-Fc PBS

a

LPS content (EU/ml)a 0.094 (+/- 0.007) 0.096 (+/- 0.017) 0.095 (+/- 0.012) 0.094 (+/- 0.008) 0.092 (+/- 0.005) 0.093 (+/- 0.007) 0.107 (+/- 0.024) 0.100 (+/- 0.020) 0.103 (+/- 0.019) 0.076 (+/- 0.011) 0.100 (+/- 0.016) 0.07 (+/- 0.015)

Endotoxin levels in the purified E. coli (VHHm/b) and A. thaliana (VHH-Fc and GP4-Fc)

derived proteins was determined using Limulus amebocyte lysate-based chromogenic test and calculated relative to an endotoxin standard.

Single-domain antibodies targeting neuraminidase protect against an H5N1 influenza virus challenge.

Influenza virus neuraminidase (NA) is an interesting target of small-molecule antiviral drugs. We isolated a set of H5N1 NA-specific single-domain ant...
4MB Sizes 0 Downloads 3 Views