ArticlePDF Available

Structure of the Uncleaved Human H1 Hemagglutinin from the Extinct 1918 Influenza Virus

Authors:

Abstract and Figures

The 1918 "Spanish" influenza pandemic represents the largest recorded outbreak of any infectious disease. The crystal structure of the uncleaved precursor of the major surface antigen of the extinct 1918 virus was determined at 3.0 angstrom resolution after reassembly of the hemagglutinin gene from viral RNA fragments preserved in 1918 formalin-fixed lung tissues. A narrow avian-like receptor-binding site, two previously unobserved histidine patches, and a less exposed surface loop at the cleavage site that activates viral membrane fusion reveal structural features primarily found in avian viruses, which may have contributed to the extraordinarily high infectivity and mortality rates observed during 1918.
Content may be subject to copyright.
DOI: 10.1126/science.1093373
, 1866 (2004); 303Science
et al.James Stevens,
VirusHemagglutinin from the Extinct 1918 Influenza
Structure of the Uncleaved Human H1
This copy is for your personal, non-commercial use only.
. clicking herecolleagues, clients, or customers by
, you can order high-quality copies for yourIf you wish to distribute this article to others
. herefollowing the guidelines
can be obtained byPermission to republish or repurpose articles or portions of articles
(this information is current as of June 21, 2010 ):
The following resources related to this article are available online at www.sciencemag.org
http://www.sciencemag.org/cgi/content/full/303/5665/1866
version of this article at:
including high-resolution figures, can be found in the onlineUpdated information and services,
http://www.sciencemag.org/cgi/content/full/1093373/DC1
can be found at: Supporting Online Material
found at:
can berelated to this articleA list of selected additional articles on the Science Web sites
http://www.sciencemag.org/cgi/content/full/303/5665/1866#related-content
http://www.sciencemag.org/cgi/content/full/303/5665/1866#otherarticles
, 14 of which can be accessed for free: cites 36 articlesThis article
115 article(s) on the ISI Web of Science. cited byThis article has been
http://www.sciencemag.org/cgi/content/full/303/5665/1866#otherarticles
33 articles hosted by HighWire Press; see: cited byThis article has been
http://www.sciencemag.org/cgi/collection/biochem
Biochemistry
: subject collectionsThis article appears in the following
registered trademark of AAAS.
is aScience2004 by the American Association for the Advancement of Science; all rights reserved. The title
CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience
on June 21, 2010 www.sciencemag.orgDownloaded from
6. L. D. Belmont, T. J. Mitchison, Cell 84, 623 (1996).
7. L. Cassimeris, Curr. Opin. Cell Biol. 14, 18 (2002).
8. G. Di Paolo, B. Antonsson, D. Kassel, B. M. Riederer, G.
Grenningloh, FEBS Lett. 416, 149 (1997).
9. P. Amayed, D. Pantaloni, M. F. Carlier, J. Biol. Chem.
277, 22718 (2002).
10. P. P. Budde, A. Kumagai, W. G. Dunphy, R. Heald,
J. Cell Biol. 153, 149 (2001).
11. H. Daub, K. Gevaert, J. Vandekerckhove, A. Sobel, A.
Hall, J. Biol. Chem. 276, 1677 (2001).
12. V. S. Kraynov et al., Science 290, 333 (2000).
13. M. A. Sells, A. Pfaff, J. Chernoff, J. Cell Biol. 151, 1449
(2000).
14. T. Wittmann, G. M. Bokoch, C. M. Waterman-Storer,
J. Cell Biol. 161, 845 (2003).
15. Materials and methods are available as supporting
material on Science Online.
16. O. Griesbeck, G. S. Baird, R. E. Campbell, D. A.
Zacharias, R. Y. Tsien, J. Biol. Chem. 276, 29188
(2001).
17. M. O. Steinmetz et al., EMBO J. 19, 572 (2000).
18. L. Beretta, T. Dobransky, A. Sobel, J. Biol. Chem. 268,
20076 (1993).
19. I. A. Leighton, P. Curmi, D. G. Campbell, P. Cohen, A.
Sobel, Mol. Cell. Biochem. 127-128, 151 (1993).
20. The fast diffusion of COPY (D
COPY
13 to 18
m
2
/s as approximated by FRAP experiments) and
comparison of CFP-stathmin and YFP distribution
in live A6 cells do indicate a homogeneous distri-
bution of COPY. Supporting text is available on
Science Online.
21. H. Chneiweiss, J. Cordier, A. Sobel, J. Neurochem. 58,
282 (1992).
22. S. V. Drouva, B. Poulin, V. Manceau, A. Sobel, Endo-
crinology 139, 2235 (1998).
23. H. Imamura et al., Mol. Biol. Cell 9, 2561 (1998).
24. I. Okamoto et al., J. Biol. Chem. 274, 25525 (1999).
25. P. I. Bastiaens, I. V. Majoul, P. J. Verveer, H. D. Soling,
T. M. Jovin, EMBO J. 15, 4246 (1996).
26. U. Marklund, N. Larsson, H. M. Gradin, G. Brattsand,
M. Gullberg, EMBO J. 15, 5290 (1996).
27. T. Kuntziger, O. Gavet, A. Sobel, M. Bornens, J. Biol.
Chem. 276, 22979 (2001).
28. R. E. Carazo-Salas, E. Karsenti, Curr. Biol. 13, 1728
(2003).
29. P. T. Stukenberg, Curr. Biol. 13, R848 (2003).
30. C. M. Waterman-Storer, E. D. Salmon, J. Cell Biol.
139, 417 (1997).
31. T. Wittmann, G. M. Bokoch, C. M. Waterman-Storer,
J. Biol. Chem. 279, 6196 (2003); published electron-
ically 26 Nov 2003.
32. G. C. Brown, B. N. Kholodenko, FEBS Lett. 457, 452
(1999).
33. We thank R. Heald for the kind gift of the stathmin
mutants, T. Stauber for comments on the manuscript, T.
Zimmermann and J. Rietdorf for help with microscopy, and
V. Georget for helpful advice concerning cloning proce-
dures. We also thank I. Arnal for initial trials on the detec-
tion of stathmin-tubulin interactions.
Supporting Online Material
www.sciencemag.org/cgi/content/full/303/5665/1862/
DC1
Materials and Methods
SOM Text
Fig. S1
References
Movie S1
27 November 2003; accepted 4 February 2004
Structure of the Uncleaved
Human H1 Hemagglutinin from
the Extinct 1918 Influenza Virus
James Stevens,
1
Adam L. Corper,
1
Christopher F. Basler,
3
Jeffery K. Taubenberger,
4
Peter Palese,
3
Ian A. Wilson
1,2
*
The 1918 “Spanish” influenza pandemic represents the largest recorded outbreak
of any infectious disease. The crystal structure of the uncleaved precursor of the
major surface antigen of the extinct 1918 virus was determined at 3.0 angstrom
resolution after reassembly of the hemagglutinin gene from viral RNA fragments
preserved in 1918 formalin-fixed lung tissues. A narrow avian-like receptor-binding
site, two previously unobserved histidine patches, and a less exposed surface loop
at the cleavage site that activates viral membrane fusion reveal structural features
primarily found in avian viruses, which may have contributed to the extraordinarily
high infectivity and mortality rates observed during 1918.
Influenza is a viral infection of the respiratory
tract that affects millions of people annually.
Combined with subsequent infection from
bacterial pneumonia, influenza remains one
of the leading causes of death in the United
States, killing on average more than 50,000
people per year. However, the 1918 pandem-
ic killed over 500,000 people in the United
States and more than 20 million worldwide
(1), making it the largest and most destructive
outbreak of any infectious disease in recorded
history (2, 3). Why the 1918 virus was so
devastating is still a mystery. The pandemic
struck before viruses were known as the caus-
ative agent and, consequently, no intact virus
survived. Nevertheless, fragments of the viral
genome did survive in Alaskan victims
buried in the permafrost and in fixed and
archived autopsy material, which recently
enabled gene reassembly (47).
There are three types of influenza virus
(A, B, and C), and the 1918 virus is a
member of type A, which accounts for all
known major epidemics and pandemics.
Hemagglutinin (HA) is the surface glyco-
protein responsible for virus binding to the
host receptor, internalization of the virus,
and subsequent membrane-fusion events
within the endosomal pathway in the infect-
ed cell. HA is also the most abundant an-
tigen on the viral surface and harbors the
primary neutralizing epitopes for antibod-
ies. Fifteen avian and mammalian serotypes
of HA have been identified, but only three
have become adapted to humans in the last
century, resulting in the emergence of pan-
demic strains H1 in 1918, H2 in 1957, and
H3 in 1968 (see fig. S1 for sequences).
Recently, three small outbreaks arose from
avian subtypes (H5, H7, and H9) that man-
aged to make a direct leap to humans, but
their low transmissibility prevented major
new epidemics (810). However, the emer-
gence of future influenza virus pandemic
strains is likely (11), and their severity will
depend on the ability to contain and
combat infection by timely development of
an appropriate vaccine.
The mature HA forms homotrimers of
220 kD, with multiple glycosylation
sites. Each monomer is synthesized as a
single polypeptide precursor (HA0) that is
subsequently cleaved into HA1 and HA2
subunits (12) by a candidate trypsin-type
endoprotease, “tryptase Clara,” that has
been isolated from rat bronchiolar epitheli-
al Clara cells (13). Structural information is
available only for influenza A HAs of the
human H3 (14), swine H9 (15), and avian
H5 subtypes (15), and for an influenza C
HA esterase fusion (HEF) protein (16).
Twenty-two years after the first structural
characterization of the HA from the 1968
H3 human pandemic (14), we now present
the HA crystal structure from a second
human subtype (H1) derived from reassem-
bly of the extinct 1918 influenza virus (4 ).
The ectodomain of the HA gene (fig.
S1) from the 1918 influenza virus A/South
Carolina/1/18 (18HA0) was cloned and ex-
pressed (fig. S2) in a baculovirus expres-
sion system (17, 18). 18HA0 crystallized at
pH 5.5 (table S1) (19), and its structure was
determined by molecular replacement
(MR) to 3.0 Å resolution (table S1) (20).
18HA0 is 135 Å in length with two dis-
tinct domains (Fig. 1A). The cylindrical
trimer has a tightly intertwined “stem” do-
main at its membrane-proximal base, which
is composed of HA1 residues 11 to 51 and
276 to 329 and HA2 1 to 176 (Fig. 1A). The
dominant feature of this stalk region is the
three long parallel helices (50 amino
acids in length), one from each monomer,
1
Department of Molecular Biology,
2
Skaggs Institute
for Chemical Biology, The Scripps Research Institute,
10550 North Torrey Pines Road, La Jolla, CA 92037,
USA.
3
Department of Microbiology, Mount Sinai
School of Medicine, One Gustave L. Levy Place, Box
1124, New York, NY 10029, USA.
4
Division of Molec-
ular Pathology, Department of Cellular Pathology and
Genetics, Armed Forces Institute of Pathology, Wash-
ington, DC 20306, USA.
*To whom correspondence should be addressed. E-
mail: wilson@scripps.edu
R EPORTS
19 MARCH 2004 VOL 303 SCIENCE www.sciencemag.org1866
on June 21, 2010 www.sciencemag.orgDownloaded from
that associate to form a triple-stranded
coiled coil. This region also contains the
cleavage site where host enzymes normally
cut HA0. The membrane-distal domain
consists of a globular head, which can be
further subdivided (Fig. 1B) into the R
region, containing the receptor-binding site
and major epitopes for neutralizing anti-
bodies, and the E region, with close struc-
tural homology to the esterase domain of
influenza C HEF (16).
The superimposition of other published
HAs onto the 18HA0 monomer (Fig. 1B) by
means of their HA2 domains [root mean
square deviations (RMSDs) in table S2], in-
dicates that 18HA0 is most closely related to
the avian H5 subtype (RMSD 2.3 Å), where-
as the human H3 subtype is the most diver-
gent (RMSD 4.1 Å). The HA1 receptor (R)
region of the human H3 subtype is displaced
most (RMSD 7.4 Å) from its equivalent
18HA0 domain (Fig. 1B). This conformation-
al variability in the HA globular heads stems
primarily from a rigid-body rotation of the
HA1 receptor domains relative to the HA2
stem domains about the central threefold
axis, as described previously for H3, H5, and
H9 (21). 18HA0 is rotated to the same extent
(17°) as avian H5 relative to H3, whereas
swine H9 is intermediate (11°) (fig. S3, A
and B) (22). The individual HA1 subdomains
(R and E regions in Fig. 1B) superimpose
with low RMSDs (1.1 to 2.6 Å) (table S3).
The 18HA0 structure reveals a substan-
tially different conformation for the cleavage
site loop compared with the uncleaved H3
and cleaved H5 subtypes (Fig. 2, A to D). The
H3 cleavage loop projects out from the gly-
coprotein surface, exposing it to potential
proteases (Fig. 2A) (23), whereas the 18HA0
cleavage site abuts the HA surface (Fig. 2B)
(24). From Pro
A324
, the 18HA0 cleavage
loop extends toward the trimer interface so
that Arg
A329
now covers the electronegative
cavity that is normally occupied by the HA2
fusion peptide after cleavage activation (Fig.
2, E and F). The Arg
A329
side chain points
toward solvent as a result of repulsion from
Lys
F39
and Lys
F121
of the adjacent subunit.
Thus, Arg
A329
is substantially less exposed
than the equivalent Gln
329
(15 Å farther
from the HA surface) that was mutated
from Arg to determine the crystal structure
of an uncleaved H3 subtype (Fig. 2, A and
D) (23). In cleaved HA structures at neutral
pH, the N-terminal HA2 fusion peptide in-
serts into a negatively charged cavity and
makes up to five hydrogen bonds from its
backbone amide groups to conserved HA2
ionizable residues (Asp
B109
and Asp
B112
).
The 18HA0 cleavage loop does not pene-
trate as far into this cavity (Fig. 2D, left)
and makes only one hydrogen bond,
Ser
A325
to Asp
B112
, between the loop and
the conserved acidic residues.
These different cleavage loop conforma-
tions in the H1 and H3 structures may be
influenced in part by nearby glycosylation
sites. In 18HA0, Asn
A20
and Asn
A34
are po-
sitioned above and to the side of the cleavage
loop (upper right, Fig. 2B), creating a cavity
in which a section of the HA0 loop (Ile
B10
to
Trp
B21
) can be accommodated. Equivalent
glycosylation sites (Asn
A22
and Asn
A38
)in
uncleaved H3 are farther from the cleavage
site loop (Fig. 2A and fig. S1) and, thus, may
exert less influence on its conformation. Our
attempts to cleave the 18HA0 trimer with
tryptase [molecular weight (MW) 135 kD]
from human lung failed, even after 20
hours of incubation at 28°C, yet cleavage
with trypsin (MW 43 kD) was complete
after 20 min [at neutral pH (25)]. Newly
synthesized viral proteins are exported to
the cell surface by way of the Golgi com-
plex, where the pH becomes more acidic
during progression through the secretory
pathway (2628). During viral assembly,
the 18HA0 cleavage loop could adopt this
less exposed conformation to protect from
premature cleavage (and membrane-fusion
activation) by intracellular proteases.
From the cleavage site, the HA0 main
chain traverses the surface below the glyco-
sylation site at Asn
A20
, where it then forms
another previously unobserved miniloop
structure with Met
B17
at its tip. In H3 and H5,
the HA2 Trp
B21
indole points up toward the
distal end of the HA, allowing the main chain
to loop around toward the trimer interface.
However, in 18HA0, this miniloop alters the
Trp
B21
indole direction so that it now faces
the HA membrane proximal end (Fig. 2D,
right). Interestingly, the H5 cleaved structure
(21) is more similar to the cleaved H3 sub-
type (29) (RMSD 0.7 Å), even though its
nearby glycosylation sites map onto the H1
subtype reported here, whereas in the vicinity
of Trp
B21
(HA2), the 1918 H1 and avian H5
Fig. 1. Crystal structure of 1918 HA0 and comparison to other human, avian, and swine HAs. (A)
Overview of the 18HA0 trimer, represented as a ribbon diagram. For clarity, each monomer has
been colored differently [A (HA1), red; B (HA2), pink; C, dark gray; D, light gray; E, dark green; F, light
green]. Carbohydrates observed in the electron-density maps are colored orange and labeled with
the asparagine to which they are attached. E95 is not labeled because it is positioned immediately
behind C95. The locations of the three receptor-binding and the cleavage sites are indicated on only
one monomer. The basic patch is indicated in the light blue ellipse and consists of HA1 residues
His
C298
, His
C285
, His
C47
, Lys
C50
, and His
C275
(shown from left to right). This figure was generated
with Deepview (48) and rendered with Pov-Ray 3.5 (www.povray.org). (B) Structural comparison of
the 18HA0 monomer (red) with human H3 (green), avian H5 (orange), and swine H9 (blue) HAs.
Structures were first superimposed on the HA2 domain of 18HA0 through the following residues:
18HA0: A11 to A51, A276 to A324, and B1 to B160; H3 (PDB ID code: 2hmg): A11 to A51, A276
to A324, and B1 to B160; H5 (PDB ID code: 1jsm): A1 to A41, A276 to A324, and B1 to B160; and
H9 (PDB ID code: 1jsd): A1 to A41, A267 to A315, and B1 to B160. Figure B was generated with
VMD (47) and rendered with Tachyon (49).
R EPORTS
www.sciencemag.org SCIENCE VOL 303 19 MARCH 2004 1867
on June 21, 2010 www.sciencemag.orgDownloaded from
residues are virtually identical (Fig. 3, A and
B). Thus, local sequence differences do not
easily explain the different cleavage loop
conformations. However, a major difference
is the HA0 crystallization conditions (pH 5.5
for H1 and pH 7.5 for H5). At low pH, HA0
is reported to be stable (30), and the 18HA0
structure confirms this assertion. Once
cleaved, HA is metastable at low pH and
undergoes irreversible conformational chang-
es, a prerequisite for membrane fusion and
infection (31). What exactly triggers this
change is still not clear because only the pre-
and postacidification conformations have
been determined (14, 32).
However, the differences between H1 and
H5 structures around HA2 Trp
B21
and the
cleavage site may hint at a possible mecha-
nism for fusion that until now was not appar-
ent from other structures crystallized at pH
7.5. In H1 and H5 structures, HA2 Trp
B21
is
surrounded by three pH-sensitive histidines
(His
A18
, His
A38
, and His
B111
) that form a
largely uncharged pocket at neutral pH (Fig.
3A). These histidines are conserved in human
H1, H2, and H5 sequences (www.flu.lanl.
gov), but only at two positions in 1999 H9
sequences (His
A38
and His
B111
), and at one
position (His
A18
) in human H3. Below pH
6.0, this pocket becomes positively charged,
and may account for the conformation differ-
ences observed between 18HA0 and H5 (Fig.
3B). In 18HA0, His
A38
points toward the tip
of the novel loop (Ile
B10
to Trp
B21
) and the
His
A18
backbone hydrogen bonds to the main
chain at Trp
B21
. Such differences may re-
veal a mechanism (yet to be tested experi-
mentally) for destabilization or even expul-
sion of the cleaved fusion peptide from the
electronegative cavity in H1 HAs. In this
uncleaved structure, the glycoprotein may
not undergo its full rearrangement because
of the physical connection of the HA2 fu-
sion peptide to HA1.
The primary event in influenza infection
is the binding of the virus to the host receptor.
The HA receptor-binding site is situated in a
shallow pocket in the membrane-distal HA1
domain. The nature of the receptor sialic acid
linkage to the vicinal galactose is the primary
determinant in lung epithelial cells that dif-
ferentiates avian viruses from mammals (spe-
cies barrier). Avian viruses preferentially
bind to receptors with an 2,3 linkage,
whereas human-adapted viruses are specific
for the 2,6 linkage (3335). In particular,
residues 226 and 228 have been linked to
receptor specificity (36, 37) and are Gln
226
and Gly
228
in avian viruses but Leu
226
and
Ser
228
in human-adapted H3 viruses (Fig. 4).
On the contrary, in 18HA and other human
H1 viruses, avian-type residues predominate.
Despite these binding-sequence correlations,
human H1s, such as A/PR/8/34 and A/FM/1/
47, can bind sialic acid receptors with both
2,3 and 2,6 linkages, albeit with reduced
affinity for the latter (38, 39). The only dif-
ference between swine- and swine-avian
adapted H1 viruses is a Glu
190
3 Asp
190
mu-
tation (4), that, although subtle, leads to a
slight increase in the pocket size (upper left
Fig. 2. Structural comparison of the 18HA0 cleavage site with other HAs. HA2 domains for human H3
HA0 (PDB ID code: 1ha0) and cleaved avian H5 HA1/HA2 (PDB ID code: 1jsm) (50) were aligned with
18HA0. The cleavage sites are colored (A) green for human H3 HA0, (B) red for 18HA0 and (C) orange
for H5 HA1/HA2. R
A329
Q, Arg
A329
3 Gln
A329
.(D) Overlay of cleavage loops of H3 HA0, H1 HA0, and
avian H5 HA1/HA2. The two views differ by a rotation of 90° about the threefold vertical axis. (E)
Surface views showing the trimer interface and the position of the cleavage loop. (F) Removal of the
cleavage loop reveals the electronegative cavity that it masks. Arg
329
is colored blue and N-acetyl-
glucosamines, indicating the nearby glycosylation sites, are colored gold. (A) to (D) were generated as
in Fig. 1, and (E) and (F) were generated with MSMS (46) through the program VMD (47).
Fig. 3. Structural com-
parisons of the environ-
ment around HA2 Trp
21
in 18HA0 and H5 HA1/
HA2. The avian H5
structure (PDB ID code:
1jsm) was aligned with
the 18HA0 model for
comparison, as in Fig. 2.
In the avian structure
(A), His
A18
and His
A38
are 3.7 Å apart,
whereas in 18HA0 (B),
the same residues are
13.5 Å apart. The
Trp
B21
“flip” in 18HA0 is
stabilized by close proximity to the side chains of Trp
B14
and Ala
B36
. This figure was generated in the same
way as Fig. 1A.
R EPORTS
19 MARCH 2004 VOL 303 SCIENCE www.sciencemag.org1868
on June 21, 2010 www.sciencemag.orgDownloaded from
side in Fig. 4) that could perhaps increase
affinity for 2,6 linkages.
Comparison of 18HA0 structure with
other subtypes reveals that the receptor-
binding site is more akin to avian than to
human HAs (Fig. 4). The 18HA0 pocket is
narrower than human H3 and swine H9
HAs, consistent with previous reports that a
reduced width of the avian receptor-
binding site enhances interaction of Gln
226
with Ala
138
with 2,3-linked disaccharides
(15, 40). The question of how 18HA so
efficiently infected humans remains open
and will require crystal structures with
bound ligands. Other, as yet unidentified,
18HA properties may also facilitate infec-
tion. A noteworthy feature is a second
patch of exposed ionizable histidines on the
HA1 chain adjacent to the vestigial esterase
domain (E region in Fig. 1B). Four HA1
histidines (His
A47
, His
A275
, His
A285
, and
His
A298
) and a lysine (Lys
A50
) (Fig. 1A
and fig. S4) contribute to a very basic
patch, not observed in other HA structures
(fig. S5) (41, 42). For example, H3 sub-
types have a glycosylation site at position
285 that masks this region. Other viruses,
such as the vesicular stomatitis virus, are
reported to depend on histidine protonation
for membrane fusion (43). Thus, the pH-
sensitive electrostatic properties of this region
in 18HA may also assist in the membrane-
fusion event, giving the virus a selective advan-
tage during infection. Clearly, experimental
testing of such a proposal is required.
Four antigenic sites for H1 HAs, includ-
ing 18HA, have been identified (Ca, Cb,
Sa, and Sb) (4, 44). In 18HA0, with the
exception of Ca, all are exposed for anti-
body recognition (fig. S6). The Ca site is
proximal to the oligosaccharide at HA1
Asn
95
, which may interfere with antibody
recognition of both subregions, Ca1 and
Ca2. In other reported HA structures, only
H9 has a glycosylation site at this position
(21). Otherwise, the closest relative (94%
identity) to the 1918 HA is the 1930 swine
virus (A/swine/Iowa/15/30) that is believed
to have evolved from the 1918 virus (45).
Because the life-span of swine is short,
immunological drift is much slower, and as
a result, any differences between these
1918 and 1930 viruses are minimized. Un-
fortunately, because no virus samples exist
for comparison from before 1918, it is dif-
ficult to reconstruct the data necessary to
fully explain the pathogenicity of this virus.
However, statistics reveal that people over 65
years old in 1918 were no more at risk than
for a normal pandemic (2), which suggests
that people born before 1855 may have
acquired some resistance to a related H1 virus
or other cross-reactive subtype (44).
The publication of the first sequences
(4) of the 1918 HA did not reveal any
characteristics that were obviously respon-
sible for the extreme pathogenicity of the
1918 pandemic, such as the polybasic res-
idues that make avian viruses so lethal.
Notwithstanding, recent data suggest that,
when expressed on a mouse-adapted WSN
viral backbone (A/WSN/33, H1N1 virus),
18HA is more virulent than a control H1
(A/New Caledonia/20/99) (17). The struc-
tural analysis here reveals a viral antigen
with a number of previously unobserved
features that may have contributed to al-
tered cleavage properties and/or fusion
propensity. Such characteristics may have
endowed the virus with unusual mecha-
nisms, which have not been seen in subse-
quent infections, that enhanced host-cell
infection, particularly in those individuals
with no previous exposure to an antigeni-
cally similar virus, which could have pro-
vided some antibody protection. Finally,
previously unobserved aspects of the ex-
pression system used here provide impor-
tant methodological advances for future
production of unprocessed HAs and other
homotrimeric viral coat proteins, such as
human immunodeficiency virus1 gp41,
for which additional structural information
is urgently needed.
References and Notes
1. A. H. Reid, J. K. Taubenberger, T. G. Fanning, Microbes
Infect. 3, 81 (2001).
2. What distinguished this pandemic from all others was
the high proportion of deaths among young adults. For
a typical influenza epidemic, a plot of age versus death
rate is usually U-shaped, meaning that the very young
and old are in the high-risk groups. For the 1918 pan-
demic, the graph was “W”-shaped, with a sharp peak
corresponding surprisingly to a high death rate among
15 to 34 year olds (3). Mortality rates were severe, over
2.5%, compared with 0.1% for more modern epidemics.
Some isolated populations, such as communities of
Alaskan Eskimos, experienced mortality rates above
70%. Most of these deaths occurred in young adults,
between 15 and 34 years of age, around 20 times as
high as in previous years, with 99% of excess deaths
among people under 65 years of age.
3. A. H. Reid, J. K. Taubenberger, Lab. Invest. 79, 95 (1999).
4. A. H. Reid, T. G. Fanning, J. V. Hultin, J. K. Tauben-
berger, Proc. Natl. Acad. Sci. U.S.A. 96, 1651 (1999).
5. A. H. Reid, T. G. Fanning, T. A. Janczewski, J. K. Tauben-
berger, Proc. Natl. Acad. Sci. U.S.A. 97, 6785 (2000).
6. C. F. Basler et al., Proc. Natl. Acad. Sci. U.S.A. 98,
2746 (2001).
7. A. H. Reid, T. G. Fanning, T. A. Janczewski, S. McCall,
J. K. Taubenberger, J. Virol. 76, 10717 (2002).
8. K. Subbarao et al., Science 279, 393 (1998).
9. M. Peiris et al., Lancet 354, 916 (1999).
10. J. C. de Jong et al., Ned. Tijdschr. Geneeskd. 147,
1971 (2003).
11. R. J. Webby, R. Webster, Science 302, 1519 (2003).
12. D. C. Wiley, J. J. Skehel, Annu. Rev. Biochem. 56, 365
(1987).
13. H. Kido, K. Sakai, Y. Kishino, M. Tashiro, FEBS Lett.
322, 115 (1993).
14. I. A. Wilson, J. J. Skehel, D. C. Wiley, Nature 289, 366
(1981).
Fig. 4. Structural comparison of
HA receptor-binding sites. (left)
18HA0 receptor-binding site
showing key conserved HA1 res-
idues that determine receptor
specificity. The H3 and H5 sub-
types are shown for comparison.
(right) Corresponding solvent-
excluded surfaces [probe 1.4 Å,
calculated with the program
MSMS (46)] of the receptor-
binding sites showing the surface
cavity for binding the host re-
ceptor sialic acid (51). Clearly,
the narrower 1918 HA0 binding
site (top) resembles the avian
H5 structure (bottom), rather
than more open human H3
(middle) or swine H9 binding
site (25). The Glu
190
3 Asp
190
(E190D) mutation slightly in-
creases the width of the 1918
H1 binding site compared with
avian H5. Such a small change
may allow accommodations of
different conformations of 2,6-
versus 2,3-linked sugars. This fig-
ure was generated with VMD (47)
and rendered with Tachyon (49).
R EPORTS
www.sciencemag.org SCIENCE VOL 303 19 MARCH 2004 1869
on June 21, 2010 www.sciencemag.orgDownloaded from
15. Y. Ha, D. J. Stevens, J. J. Skehel, D. C. Wiley, Proc.
Natl. Acad. Sci. U.S.A. 98, 11181 (2001).
16. P. B. Rosenthal et al., Nature 396, 92 (1998).
17. T. M. Tumpey et al., Proc. Natl. Acad. Sci. U.S.A. 99,
13849 (2002).
18. Materials and methods are available as supporting
material on Science Online.
19. 18HA0 at a concentration of 10 to 15 mg/ml was
used to grow crystals in sitting drops with a precip-
itant solution of 1.68 M sodium dihydrogen phos-
phate, 0.32 M dipotassium hydrogen phosphate, and
0.1 M phosphate-citrate, with pH 5.5 (18).
20. Swine H9 HA [Protein Data Bank (PDB) identification
(ID) code: 1jsh] was used as the initial MR model. The
final R factor R
cryst
and free R factor R
free
values (table
S1 legend) are 27.0 and 29.6%, respectively, with only
three residues (0.7%) per monomer (Ile
B10
, Met
B17
, and
Asn
B60
) in the disallowed regions of the Ramachandran
plot. Only Met
B17
has good density and is positioned at
the tip of a loop, proximal to the membrane-fusion
loop. The crystal asymmetric unit contains one ho-
motrimer (502 HA0-encoded residues per monomer)
with an estimated solvent content of 66%, based on a
Matthews’ coefficient (V
m
), of 3.57 Å
3
/dalton. For com-
parison with previous structures, the HA0 sequence is
numbered in the same way as the H3 subtype (14) and
labeled HA1 11 to 329 and HA2 1 to 175, even though
the covalent bond still links HA1 to HA2. Thus, the H1
1918 HA0 structure begins at residue 11 because of an
insertion of 10 residues in H3. Insertions in H1 relative
to H3 are labeled by the preceding residue with a letter
(e.g., Asn
19A
). The three HA0 chains in the trimer are
labeled A, C, and E (HA1) and B, D, and F (HA2). The
electron density maps revealed only two disordered
regions around residues 78 to 81 of chains HA1 and 10
to 14 of HA2 that, although in disparate regions of the
monomer, are in close proximity to other symmetry-
related molecules in the crystal. Although B values were
high in this region, main-chain atoms could be inter-
preted from the electron density in both regions, except
at Gly
12
and Gly
13
in the HA2 chain. Occupancy and B
values were set to zero for side-chain atoms that were
uninterpretable from the electron density.
21. Y. Ha, D. J. Stevens, J. J. Skehel, D. C. Wiley, EMBO J.
21, 865 (2002).
22. Angles of rotation reported here for the H5 and H9
subtypes are less than those in (21). Our method of
analysis, as described in the legend for Fig. 1, was
different but reports a similar trend and reveals a
rotation for the 18HA0 as seen for the avian subtype.
23. J. Chen et al., Cell 95, 409 (1998).
24. Electron density of the main chains around Arg
329
was well defined and could be traced through to
residue 10 in HA2, at which point the electron den-
sity became disordered through to residue 14.
25. J. Stevens et al., data not shown.
26. The pH ranges from almost neutral on exiting the
endoplasmic reticulum to pH 5.9 within the trans-
Golgi network and as low as pH 5.4 in the secretory
vesicles (27, 28).
27. N. Demaurex, W. Furuya, S. D’Souza, J. S. Bonifacino,
S. Grinstein, J. Biol. Chem. 273, 2044 (1998).
28. M. Grabe, G. Oster, J. Gen. Physiol. 117, 329 (2001).
29. RMSD was calculated by overlapping C’s at A312 to
A324 and B1 to B22 of H3 (PDB ID code: 2hmg) with
residues A309 to A321 and B1 to B22 of H5 (PDB ID
code: 1jsm).
30. C. Bo¨ttcher, K. Ludwig, A. Herrmann, M. van Heel, H.
Stark, FEBS Lett. 463, 255 (1999).
31. J. J. Skehel et al., Proc. Natl. Acad. Sci. U.S.A. 79, 968
(1982).
32. P. A. Bullough, F. M. Hughson, J. J. Skehel, D. C. Wiley,
Nature 371, 37 (1994).
33. R. J. Connor, Y. Kawaoka, R. G. Webster, J. C. Paulson,
Virology 205, 17 (1994).
34. G. N. Rogers, B. L. D’Souza, Virology 173, 317 (1989).
35. J. J. Skehel, D. C. Wiley, Annu. Rev. Biochem. 69, 531
(2000).
36. G. N. Rogers et al., Nature 304, 76 (1983).
37. C. T. Hardy, S. A. Young, R. G. Webster, C. W. Naeve,
R. J. Owens, Virology 211, 302 (1995).
38. G. N. Rogers, J. C. Paulson, Virology 127, 361 (1983).
39. Y. Suzuki et al., Biochim. Biophys. Acta 903, 417
(1987).
40. The widths of the binding pockets were calculated by
measuring the distance from Gly
A134
C to Gln
A222
C for 18HA0 (13.9 Å), Gly
A134
C to Gln
A226
C for
human H3 (15.2 Å), Gly
A130
C to Gln
A222
C for
avian H5 (14.0 Å), and Gly
A128
C to Gln
A216
C for
swine H9 (15.2 Å).
41. Only one other sequence in the current influenza
database of human H1, H2, H3, H5, and H9 subtypes
possesses the same patch (H1; A/Alma-Ata/1417/84
virus) (42).
42. A. B. Beklemishev et al., Mol. Gen. Mikrobiol. Virusol.
1, 24 (1993).
43. F. A. Carneiro et al., J. Biol. Chem. 278, 13789 (2003).
44. G. G. Brownlee, E. Fodor, Philos. Trans. R. Soc. London
Ser. B 356, 1871 (2001).
45. R. E. Shope, J. Exp. Med. 63, 669 (1936).
46. M. F. Sanner, A. J. Olson, J. C. Spehner, Biopolymers
38, 305 (1996).
47. W. Humphrey, A. Dalke, K. Schulten, J. Mol. Graph.
14, 33 (1996).
48. N. Guex, M. C. Peitsch, Electrophoresis 18, 2714
(1997).
49. J. E. Stone, thesis, University of Missouri (1998).
50. Residues 311 to 324 and 350 to 357 from the human
H3 subtype (PDB ID code: 1ha0) and A308 to A321
and B21 to B28 from the avian H5 subtype (PDB ID
number: 1jsm) were aligned with residues A311 to
A324 and B21 to B28 of 18HA0.
51. Residues A95 to A99, A128 to A161, A179 to A200,
and A220 to A230 from 18HA0; A95 to A99, A128 to
A161, A179 to A200, and A220 to A230 from the
human H3 subtype (PDB ID code: 2hmg); A87 to A92,
A124 to A156, A175 to A196, and A216 to A226
from the avian H5 subtype (PDB ID code: 1jsm); A87
to A92, A122 to A150, A169 to A190, and A210 to
A220 from the swine H9 subtype (PDB ID code: 1jsd)
were aligned and surfaces were generated with
MSMS (46) through the program VMD (47).
52. I.A.W. is supported by NIH grants CA55896 and
AI42266 and National Institute of General Medical
Sciences (NIGMS) grant P50-GM 62411. P.P. and
C.F.B. are both supported by NIH grants. C.F.B. is an
Ellison Medical Foundation New Scholar in Global
Infectious Diseases. P.P. is an Ellison Medical Founda-
tion Senior Scholar. J.K.T. is supported by NIH grant
AI50619 and by the intramural funds of the Armed
Forces Institute of Pathology. This work was carried
out in the framework of a multidisciplinary influenza
consortium with a pending NIH grant AI058113-01.
We thank the staff of the Stanford Synchrotron
Radiation Laboratory (SSRL) Beamline 9-2 for the
beamline assistance; X. Dai and T. Horton ( The
Scripps Research Institute) for expert technical assist-
ance; and R. Stanfield, M. Elsliger, and D. Zajonc ( The
Scripps Research Institute) for helpful discussions.
This is publication 16185-MB from The Scripps Re-
search Institute. Coordinates and structure factors
have been deposited in the PDB (ID code 1RD8).
Supporting Online Material
www.sciencemag.org/cgi/content/full/1093373/DC1
Materials and Methods
Figs. S1 to S6
Tables S1 to S3
References and Notes
6 November 2003; accepted 7 January 2004
Published online 5 February 2004;
10.1126/science.1093373
Include this information when citing this paper.
Conserved Genetic Basis of a
Quantitative Plumage Trait
Involved in Mate Choice
Nicholas I. Mundy,
1
* Nichola S. Badcock,
2
Tom Hart,
2
Kim Scribner,
3
Kirstin Janssen,
4
Nicola J. Nadeau
1
A key question in evolutionary genetics is whether shared genetic mechanisms
underlie the independent evolution of similar phenotypes across phylogeneti-
cally divergent lineages. Here we show that in two classic examples of melanic
plumage polymorphisms in birds, lesser snow geese (Anser c. caerulescens) and
arctic skuas (Stercorarius parasiticus), melanism is perfectly associated with
variation in the melanocortin-1 receptor (MC1R) gene. In both species, the
degree of melanism correlates with the number of copies of variant MC1R
alleles. Phylogenetic reconstructions of variant MC1R alleles in geese and skuas
show that melanism is a derived trait that evolved in the Pleistocene.
The genetic basis of independent origins of
the same phenotype is important to models of
phenotypic evolution. There are few data,
especially for vertebrates, because the loci
underlying phenotypic evolution in natural
populations are rarely known. The lesser
snow goose (Anser c. caerulescens) and arc-
tic skua (or parasitic jaeger, Stercorarius
parasiticus) have prominent melanic plum-
age polymorphisms (Fig. 1) showing clinal
variation in the frequency of melanic morph
phenotypes across their arctic breeding rang-
es (1, 2). In both species, there is quantitative
variation in the degree of melanism among
adult individuals with the melanic phenotype
(blue snow geese and intermediate and
dark skuas) and discrete separation be-
tween these and the nonmelanic phenotypes
(white geese and pale skuas).
These polymorphisms influence mate
choice. In snow geese, mate color preference
follows parental color, leading to assortative
1
Department of Zoology, University of Cambridge,
Cambridge CB2 3EJ, UK.
2
Department of Biological
Anthropology, University of Oxford, Oxford OX2
6QS, UK.
3
Department of Fisheries and Wildlife and
Department of Zoology, Michigan State University,
East Lansing, MI 48824, USA.
4
Department of Molec-
ular Biotechnology, University of Tromsø, N-9037
Tromsø, Norway.
*To whom correspondence should be addressed. E-
mail: nim21@cam.ac.uk
R EPORTS
19 MARCH 2004 VOL 303 SCIENCE www.sciencemag.org1870
on June 21, 2010 www.sciencemag.orgDownloaded from
... These serum samples were tested by ELISA to evaluate the total binding antibodies against with year 2 sera. The purified and trimeric recombinant HA proteins were expressed from the baculovirus system using the established procedures 45 . For ELISA, rHA antigens were coated at 100 ng/well. ...
Article
Full-text available
Repeat vaccination with egg-based influenza vaccines could preferentially boost antibodies targeting the egg-adapted epitopes and reduce immunogenicity to circulating viruses. In this randomized trial (Clinicaltrials.gov: NCT03722589), sera pre- and post-vaccination with quadrivalent inactivated egg-based (IIV4), cell culture-based (ccIIV4), and recombinant (RIV4) influenza vaccines were collected from healthcare personnel (18-64 years) in 2018−19 (N = 723) and 2019−20 (N = 684) influenza seasons. We performed an exploratory analysis. Vaccine egg-adapted changes had the most impact on A(H3N2) immunogenicity. In year 1, RIV4 induced higher neutralizing and total HA head binding antibodies to cell- A(H3N2) virus than ccIIV4 and IIV4. In year 2, among the 7 repeat vaccination arms (IIV4-IIV4, IIV4-ccIIV4, IIV4-RIV4, RIV4-ccIIV4, RIV4-RIV4, ccIIV4-ccIIV4 and ccIIV4-RIV4), repeat vaccination with either RIV4 or ccIIV4 further improved antibody responses to circulating viruses with decreased neutralizing antibody egg/cell ratio. RIV4 also had higher post-vaccination A(H1N1)pdm09 and A(H3N2) HA stalk antibodies in year 1, but there was no significant difference in HA stalk antibody fold rise among vaccine groups in either year 1 or year 2. Multiple seasons of non-egg-based vaccination may be needed to redirect antibody responses from immune memory to egg-adapted epitopes and re-focus the immune responses towards epitopes on the circulating viruses to improve vaccine effectiveness.
... Massive efforts have been undertaken to develop a universal influenza vaccine by eliciting strong immune responses to highly conserved antigens, such as the stalk domain of hemagglutinin (HA), the ectodomain of the M2 ion channel (M2e), and the internal proteins nucleoprotein (NP) and matrix protein (M1) [7,8]. HA, the primary antigen of the influenza virus, is a homotrimeric glycoprotein present on the virion surface and is composed of a membrane distal globular head domain and a proximal stalk domain [9]. The globular head domain with abundant neutralizing epitopes is immunogenic but frequently undergoes antigenic drift, enabling virus evasion from herd immunity [10]. ...
Article
Full-text available
Classic chimeric hemagglutinin (cHA) was designed to induce immune responses against the conserved stalk domain of HA. However, it is unclear whether combining more than one HA head domain onto one stalk domain is immunogenic and further induce immune responses against influenza viruses. Here, we constructed numerous novel cHAs comprising two or three fused head domains from different subtypes grafted onto one stalk domain, designated as cH1-H3, cH1-H7, cH1-H3-H7, and cH1-H7-H3. The three-dimensional structures of these novel cHAs were modeled using bioinformatics simulations. Structural analysis showed that the intact neutralizing epitopes were exposed in cH1-H7 and were predicted to be immunogenic. The immunogenicity of the cHAs constructs was evaluated in mice using a chimpanzee adenoviral vector (AdC68) vaccine platform. The results demonstrated that cH1-H7 expressed by AdC68 (AdC68-cH1-H7) induced the production of high levels of binding antibodies, neutralizing antibodies, and hemagglutinin inhibition antibodies against homologous pandemic H1N1, drifted seasonal H1N1, and H7N9 virus. Moreover, vaccinated mice were fully protected from a lethal challenge with the aforementioned influenza viruses. Hence, cH1-H7 cHAs with potent immunogenicity might be a potential novel vaccine to provide protection against different subtypes of influenza virus.
... Hence, we sought to make the soluble recombinant HA protein against the Inf A/Guangdong-Maonan/SWL1536/2019 virus. The HA amino acid sequence 18-530 was used for the generation of HA protein and linked to the fold on trimerization domain at the C-terminus for trimer formation and to preserve the native-like trimeric conformation of HA [37][38][39]. The sequence was further linked with an avidin (avi) tag, cleavable tobacco etch virus (TEV) protease site, and a His6 tag at the C terminus for ease of purification ( Figure 1B). ...
Article
Full-text available
Immunogens mimicking the native-like structure of surface-exposed viral antigens are considered promising vaccine candidates. Influenza viruses are important zoonotic respiratory viruses with high pandemic potential. Recombinant soluble hemagglutinin (HA) glycoprotein-based protein subunit vaccines against Influenza have been shown to induce protective efficacy when administered intramuscularly. Here, we have expressed a recombinant soluble trimeric HA protein in Expi 293F cells and purified the protein derived from the Inf A/Guangdong-Maonan/ SWL1536/2019 virus which was found to be highly virulent in the mouse. The trimeric HA protein was found to be in the oligomeric state, highly stable, and the efficacy study in the BALB/c mouse challenge model through intradermal immunization with the prime-boost regimen conferred complete protection against a high lethal dose of homologous and mouse-adapted InfA/PR8 virus challenge. Furthermore, the immunogen induced high hemagglutinin inhibition (HI) titers and showed cross-protection against other Inf A and Inf B subtypes. The results are promising and warrant trimeric HA as a suitable vaccine candidate.
... Recombinant antigens comprising the RBD or the ectodomain of the full spike of BANAL-236, À52, and À103 viruses were designed in fusion with the nanoluciferase as follows: the foldon domain (YIPEAPRDGQAYVRKDGEWVLLSTFL) was added to the C-terminus of each ectodomain to allow the S protein to trimerize (Stevens et al, 2004), resembling the native spike state of the virion. The nanoluciferase was added to the carboxy-terminal end of each construct spaced by a 3-residues GSG linker. ...
Article
Bat sarbecovirus BANAL-236 is highly related to SARS-CoV-2 and infects human cells, albeit lacking the furin cleavage site in its spike protein. BANAL-236 replicates efficiently and pauci-symptomatically in humanized mice and in macaques, where its tropism is enteric, strongly differing from that of SARS-CoV-2. BANAL-236 infection leads to protection against superinfection by a virulent strain. We find no evidence of antibodies recognizing bat sarbecoviruses in populations in close contact with bats in which the virus was identified, indicating that such spillover infections, if they occur, are rare. Six passages in humanized mice or in human intestinal cells, mimicking putative early spillover events, select adaptive mutations without appearance of a furin cleavage site and no change in virulence. Therefore, acquisition of a furin site in the spike protein is likely a pre-spillover event that did not occur upon replication of a SARS-CoV-2-like bat virus in humans or other animals. Other hypotheses regarding the origin of the SARS-CoV-2 should therefore be evaluated, including the presence of sarbecoviruses carrying a spike with a furin cleavage site in bats.
... The N terminal contains 23 hydrophobic peptides required for membrane fusion, with two repeated allosteric peptides in the middle. About 28 amino acids are anchored on the membrane at the C terminal, followed by the tail of the intracellular region of 10 residues (Stevens et al., 2004). When HA matures, HA0 is hydrolyzed and cut into HA1 and HA2 by a protease to make the virus infectious and determine its ability to spread in infected host tissues. ...
Article
Full-text available
Background The antiviral activity and underlying mechanism of Patchouli alcohol remain unclear.Methods This study evaluated the cytotoxicity, optimal methods for drug administration, anti-influenza A activity of Patchouli alcohol. The antiviral mechanism of Patchouli alcohol was also assessed via qRT-PCR, western blot, hemagglutination inhibiting (HAI) assay, and hemolysis inhibiting assay.ResultsPatchouli alcohol was shown to have low cytotoxicity and its strongest antiviral effect was associated with premixed administration. Patchouli alcohol inhibited virus replication during the early lifecycle stages of influenza A virus infection and specifically prevented expression of the viral proteins, HA and NP. In both the HAI and hemolysis inhibiting assays, Patchouli alcohol was able to block HA2-mediated membrane fusion under low pH conditions. Patchouli alcohol had lower binding energy with HA2 than HA1.Conclusion These findings suggest that Patchouli alcohol could be a promising membrane fusion inhibitor for the treatment of influenza A infection.
... The spike protein of SARS-CoV-2 is a homo-trimer [37], just such as some other class I viral glycoproteins of enveloped RNA viruses such as HIV [82,83] and influenza [84,85], etc. Generally, sufficient neutralizing antibodies in serum or mucosal secretions induced by vaccines play a vital role in blocking viruses. ...
Article
Full-text available
Vaccination is the most cost-effective means in the fight against infectious diseases. Various kinds of vaccines have been developed since the outbreak of COVID-19, some of which have been approved for clinical application. Though vaccines available achieved partial success in protecting vaccinated subjects from infection or hospitalization, numerous efforts are still needed to end the global pandemic, especially in the case of emerging new variants. Safe and efficient vaccines are the key elements to stop the pandemic from attacking the world now; novel and evolving vaccine technologies are urged in the course of fighting (re)-emerging infectious diseases. Advances in biotechnology offered the progress of vaccinology in the past few years, and lots of innovative approaches have been applied to the vaccine design during the ongoing pandemic. In this review, we summarize the state-of-the-art vaccine strategies involved in controlling the transmission of SARS-CoV-2 and its variants. In addition, challenges and future directions for rational vaccine design are discussed.
Article
The baculovirus/insect cell expression system is a very useful tool for reagent and antigen generation in vaccinology, virology, and immunology. It allows for the production of recombinant glycoproteins, which are used as antigens in vaccination studies and as reagents in immunological assays. Here, we describe the process of recombinant glycoprotein production using the baculovirus/insect cell expression system.
Article
Full-text available
Background: Highly pathogenic avian H5 influenza viruses have spread and diversified genetically and antigenically into multiple clades and subclades. Most isolates of currently circulating H5 viruses are in clade 2.3.2.1 or 2.3.4.4. Methods: Panels of murine monoclonal antibodies (mAbs) were generated to the influenza hemagglutinin (HA) of H5 viruses from the clade 2.3.2.1 H5N1 vaccine virus A/duck/Bangladesh/19097/2013 and the clade 2.3.4.4 H5N8 vaccine virus A/gyrfalcon/Washington/41088-6/2014. Antibodies were selected and characterized for binding, neutralization, epitope recognition, cross-reactivity with other H5 viruses, and the ability to provide protection in passive transfer experiments. Results: All mAbs bound homologous HA in an ELISA format; mAbs 5C2 and 6H6 were broadly binding for other H5 HAs. Potently neutralizing mAbs were identified in each panel, and all neutralizing mAbs provided protection in passive transfer experiments in mice challenged with a homologous clade influenza virus. Cross-reacting mAb 5C2 neutralized a wide variety of clade 2.3.2.1 viruses, as well as H5 viruses from other clades, and also provided protection against heterologous H5 clade influenza virus challenge. Epitope analysis indicated that the majority of mAbs recognized epitopes in the globular head of the HA. The mAb 5C2 appeared to recognize an epitope below the globular head but above the stalk region of HA. Conclusions: The results suggested that these H5 mAbs would be useful for virus and vaccine characterization. The results confirmed the functional cross-reactivity of mAb 5C2, which appears to bind a novel epitope, and suggest the therapeutic potential for H5 infections in humans with further development.
Article
Influenza viruses belong to the Orthomyxoviridae family and cause acute respiratory distress in humans. The developed drug resistance toward existing drugs and the emergence of viral mutants that can escape vaccines mandate the search for novel antiviral drugs. Herein, the synthesis of epimeric 4′‐methyl‐4′‐phosphonomethoxy [4′‐C‐Me‐4′‐C‐(O‐CH2P═O)] pyrimidine ribonucleosides, their phosphonothioate [4′‐C‐Me‐4′‐C‐(O‐CH2P═S)] derivatives, and their evaluation against an RNA viral panel are described. Selective formation of the α‐ l‐lyxo epimer, [4′‐C‐(α)‐Me‐4′‐C‐(β)‐(O‐CH2‐P(═O)(OEt)2)] over the β‐ d‐ribo epimer [4′‐C‐(β)‐Me‐4′‐C‐(α)‐(O‐CH2‐P(═O)(OEt)2)] was explained by DFT equilibrium geometry optimizations studies. Pyrimidine nucleosides having the [4′‐C‐(α)‐Me‐4′‐C‐(β)‐(O‐CH2‐P(═O)(OEt)2)] framework showed specific activity against influenza A virus. Significant anti‐influenza virus A (H1N1 California/07/2009 isolate) was observed with the 4′‐C‐(α)‐Me‐4′‐C‐(β)‐O‐CH2‐P(═O)(OEt)2‐uridine derivative 1 (EC50 = 4.56 mM, SI50 > 56), 4‐ethoxy‐2‐oxo‐1(2H)‐pyrimidin‐1‐yl derivative 3 (EC50 = 5.44 mM, SI50 > 43) and the cytidine derivative 2 (EC50 = 0.81 mM, SI50 > 13), respectively. The corresponding thiophosphonates 4′‐C‐(α)‐Me‐4′‐C‐(β)‐(O‐CH2‐P( S)(OEt)2) and thionopyrimidine nucleosides were devoid of any antiviral activity. This study shows that the 4′‐C‐(α)‐Me‐4′‐(β)‐O‐CH2‐P(═O)(OEt)2 ribonucleoside can be further optimized to provide potent antiviral agents. Epimeric 4′‐methyl‐4′‐phosphonomethoxy [4′‐C‐Me‐4′‐C‐(O‐CH2P═O)] pyrimidine ribonucleosides and their phosphonothioate [4′‐C‐Me‐4′‐C‐(O‐CH2P═S)] derivatives were synthesized and evaluated against an RNA viral panel. The thiophosphonates and thionopyrimidine nucleosides are shown to be devoid of any antiviral activity, whereas the 4′‐C‐(α)‐Me‐4′‐(β)‐O‐CH2‐P(═O)(OEt)2 ribonucleoside can be further optimized to provide potent antiviral agents.
Article
Full-text available
The influenza A virus pandemic of 1918–1919 resulted in an estimated 20–40 million deaths worldwide. The hemagglutinin and neuraminidase sequences of the 1918 virus were previously determined. We here report the sequence of the A/Brevig Mission/1/18 (H1N1) virus nonstructural (NS) segment encoding two proteins, NS1 and nuclear export protein. Phylogenetically, these genes appear to be close to the common ancestor of subsequent human and classical swine strain NS genes. Recently, the influenza A virus NS1 protein was shown to be a type I IFN antagonist that plays an important role in viral pathogenesis. By using the recently developed technique of generating influenza A viruses entirely from cloned cDNAs, the hypothesis that the 1918 virus NS1 gene played a role in virulence was tested in a mouse model. In a BSL3+ laboratory, viruses were generated that possessed either the 1918 NS1 gene alone or the entire 1918 NS segment in a background of influenza A/WSN/33 (H1N1), a mouse-adapted virus derived from a human influenza strain first isolated in 1933. These 1918 NS viruses replicated well in tissue culture but were attenuated in mice as compared with the isogenic control viruses. This attenuation in mice may be related to the human origin of the 1918 NS1 gene. These results suggest that interaction of the NS1 protein with host-cell factors plays a significant role in viral pathogenesis.
Article
Full-text available
Sera from a very high proportion of the human adults and new-born infants studied neutralized swine influenza virus; sera from children below the age of 12 years seldom exerted such an effect. The results of neutralization experiments with human sera and the virus of swine influenza have been compared with the outcome of similar tests with the virus of human influenza, and it seems evident that the presence of antibodies neutralizing swine influenza virus cannot be deemed the result of repeated exposures to the current human type of virus. From the known history of swine influenza and the similarity of its etiologic virus to that obtained from man it seems likely that the virus of swine influenza is the surviving prototype of the agent primarily responsible for the great human pandemic of 1918, as Laidlaw has already suggested. The presence in human sera of antibodies neutralizing swine influenza virus is believed to indicate a previous immunizing exposure to, or infection with, an influenza virus of the 1918 type.
Article
Because of their wide use in molecular modeling, methods to compute molecular surfaces have received a lot of interest in recent years. However, most of the proposed algorithms compute the analytical representation of only the solvent-accessible surface. There are a few programs that compute the analytical representation of the solvent-excluded surface, but they often have problems handling singular cases of self-intersecting surfaces and tend to fail on large molecules (more than 10,000 atoms). We describe here a program called MSMS, which is shown to be fast and reliable in computing molecular surfaces. It relies on the use of the reduced surface that is briefly defined here and from which the solvent-accessible and solvent-excluded surfaces are computed. The four algorithms composing MSMS are described and their complexity is analyzed. Special attention is given to the handling of self-intersecting parts of the solvent-excluded surface called singularities. The program has been compared with Connolly's program PQMS [M. L. Connolly (1993) Journal of Molecular Graphics, Vol. 11, pp. 139–141] on a set of 709 molecules taken from the Brookhaven Data Base. MSMS was able to compute topologically correct surfaces for each molecule in the set. Moreover, the actual time spent to compute surfaces is in agreement with the theoretical complexity of the program, which is shown to be O[n log(n)] for n atoms. On a Hewlett-Packard 9000/735 workstation, MSMS takes 0.73 s to produce a triangulated solvent-excluded surface for crambin (1crn, 46 residues, 327 atoms, 4772 triangles), 4.6 s for thermolysin (3tln, 316 residues, 2437 atoms, 26462 triangles), and 104.53 s for glutamine synthetase (2gls, 5676 residues, 43632 atoms, 476665 triangles). © 1996 John Wiley & Sons, Inc.
Article
The three-dimensional structures of the complete haemagglutinin (HA) of influenza virus A/Japan/305/57 (H2N2) in its native (neutral pH) and membrane fusion-competent (low pH) form by electron cryo-microscopy at a resolution of 10 A and 14 A, respectively, have been determined. In the fusion-competent form the subunits remain closely associated preserving typical overall features of the trimeric ectodomain at neutral pH. Rearrangements of the tertiary structure in the distal and the stem parts are associated with the formation of a central cavity through the entire ectodomain. We suggest that the cavity is essential for relocation of the so-called fusion sequence of HA towards the target membrane.
Article
Comparative protein modeling is increasingly gaining interest since it is of great assistance during the rational design of mutagenesis experiments. The availability of this method, and the resulting models, has however been restricted by the availability of expensive computer hardware and software. To overcome these limitations, we have developed an environment for comparative protein modeling that consists of SWISS-MODEL, a server for automated comparative protein modeling and of the SWISS-PdbViewer, a sequence to structure workbench. The Swiss-PdbViewer not only acts as a client for SWISS-MODEL, but also provides a large selection of structure analysis and display tools. In addition, we provide the SWISS-MODEL Repository, a database containing more than 3500 automatically generated protein models. By making such tools freely available to the scientific community, we hope to increase the use of protein structures and models in the process of experiment design.
Article
The haemagglutinin (HA) glycoproteins of influenza virus membranes are responsible for binding viruses to cells by interacting with membrane receptor molecules which contain sialic acid (for review see ref. 1). This interaction is known to vary in detailed specificity for different influenza viruses (see, for example, refs 2-4) and we have attempted to identify the sialic acid binding site of the haemagglutinin by comparing the amino acid sequences of haemagglutinins with different binding specificities. We present here evidence that haemagglutinins which differ in recognizing either NeuAc alpha 2 leads to 3Gal- or NeuAc alpha 2 leads to 6Gal- linkages in glycoproteins also differ at amino acid 226 of HA1. This residue is located in a pocket on the distal tip of the molecule, an area previously proposed from considerations of the three-dimensional structure of the haemagglutinin to be involved in receptor binding.
Article
It has been previously reported that several human H1 influenza viruses isolated prior to 1956, in contrast to human H3 isolates which are quite specific for SA alpha 2,6Gal sequences, apparently recognize both SA alpha 2,3Gal and SA alpha 2,6Gal sequences (Rogers, G.N., and Paulson, J.C., Virology 127, 361-373, 1983). In this report human H1 isolates representative of two epidemic periods, from 1934 to 1957 and from 1977 to 1986, and H1 influenza isolated from pigs, ducks, and turkeys were compared for their ability to utilize sialyloligosaccharide structures containing terminal SA alpha 2,3Gal or SA alpha 2,6Gal sequences as receptor determinants. Five of the eight human isolates from the first epidemic period recognize both SA alpha 2,3Gal and SA alpha 2,6Gal linkages, in agreement with our previous results. Of the remaining three strains, all isolated towards the end of the first epidemic, two appear to prefer SA alpha 2,6Gal sequences while the third preferentially binds SA alpha 2,3Gal sequences. In contrast to the early isolates, 11 of 13 human strains isolated during the second epidemic period preferentially bind SA alpha 2,6Gal containing oligosaccharides. On the basis of changes in receptor binding associated with continued passage in the laboratory for some of these later strains, it seems likely that human H1 isolates preferentially bind SA alpha 2,6Gal sequences in nature, and that acquisition of SA alpha 2,3Gal-binding is associated with laboratory passage. Influenza H1 viruses isolated from pigs were predominantly SA alpha 2,6Gal-specific while those isolated from ducks were primarily SA alpha 2,3Gal-specific. Thus, as has been previously reported for H3 influenza isolates, receptor specificity for influenza H1 viruses appears to be influenced by the species from which they were isolated, human isolates binding preferentially to SA alpha 2,6Gal-containing oligosaccharides while those isolated from ducks prefer SA alpha 2,3Gal-containing oligosaccharides. However, unlike the SA alpha 2,6Gal-specific H3 isolates, binding to cell surface receptors by the H1 influenza viruses is not sensitive to inhibition by horse serum glycoproteins, regardless of their receptor specificity. These results suggest that, while the H1 and H3 hemagglutinins appear to be subject to similar host-derived selective pressures, there appear to be certain fundamental differences in the detailed molecular interaction of the two hemagglutinins with their sialyloligosaccharide receptor determinants.
Article
A cryptically I-active sialylglycoprotein (glycoprotein 2) isolated from bovine erythrocyte membranes as Sendai virus receptor (Suzuki, Y., Suzuki, T. and Matsumoto, M. (1983) J. Biochem. 93, 1621-1633) contains N-glycolylneuraminic acid (NeuGc) as its predominate sialic acid and exhibits poor receptor activity for a variety of influenza viruses. Enzymatic modification of asialoglycoprotein-2 to contain N-acetylneuraminic acid (NeuAc) in the NeuAc alpha 2-3Gal and NeuAc alpha 2-6Gal sequences using specific sialyltransferase resulted in the appearance of receptor activity toward human influenza viruses A and B. The biological responsiveness chicken erythrocytes treated with sialidase and then reconstituted with derivatized glycoprotein 2 showed considerable recovery to influenza virus hemagglutinin-mediated agglutination, low-pH fusion and hemolysis. Specific hemagglutination inhibition activity of derivatized glycoprotein 2 was 5-16-times higher than that of human glycophorin. A/PR/8/34 (H1N1) virus preferentially recognized derivatized glycoprotein 2 containing NeuAc alpha 2-3Gal sequence over that containing NeuAc alpha 2-6Gal while the specificity of A/Aichi/2/68 (H3N2) for the sialyl linkages was reversed. B/Lee virus recognized both sequences almost equally. The biological responsiveness to the viruses of the erythrocytes labeled with the derivatized glycoprotein 2 containing NeuGc was considerably lower than that of derivatized glycoprotein 2 containing NeuAc. The results demonstrate that the hemagglutinins of human isolates of influenza viruses A and B differ in the recognition of microdomains (NeuAc, NeuGc) of the receptors for binding and fusion activities in viral penetration and the sequence to which sialic acid (SA) is attached (SA alpha 2-3Gal, SA alpha 2-6Gal). Inner I-active neolacto-series type II sugar chains may be important in revealing the receptor activity toward the hemagglutinin of both human influenza viruses A and B.
Article
The binding of influenza virus to erythrocytes and host cells is mediated by the interaction of the viral hemagglutinin (H) with cell surface receptors containing sialic acid (SA). The specificity of this interaction for 19 human and animal influenza isolates was examined using human erythrocytes enzymatically modified to contain cell surface sialyloligosaccharides with the sequence SAα2,6Galβ1,4GlcNAc; SAα2,3Galβ1,4(3)GlcNAc; SAα2,3Galβ1,3GalNAc; or SAα2,6GalNAc. Although none of the viruses agglutinated cells containing the SAα2,6GalNAc linkage, differential agglutination of cells containing the other three sequences revealed at least three distinct receptor binding types. Several virus isolates exhibited marked receptor specificity, binding only to cells containing the SAα2,6Gal or the SAα2,3Gal linkage, while others bound equally well to cells containing either linkage. Moreover, some viruses could distinguish between two oligosaccharide receptor determinants containing the terminal SAα2,3Gal linkage when present in the SAα2,3Galβ1,4(3)GlcNAc sequence or the SAα2,3Galβ1,3GalNAc sequence binding cells containing only the former. The observed receptor specificities were not significantly influenced by the viral neuraminidases as shown by the use of the potent neuraminidase inhibitor 2-deoxy-2,3-dehydro-N-acetylneuraminic acid. Receptor specificity appeared, to some extent, to be dependent on the species from which the virus was isolated. In particular, human isolates of the H3 serotype all agglutinated cells containing the SAα2,6Gal linkage, but not cells bearing the SAα2,3Galβ1,3GalNAc sequence. In contrast, antigenically similar (H3) isolates from avian and equine species preferentially bound erythrocytes containing the SAα2,3Gal linkage. This is of particular interest in view of the identification of the avian virus H3 hemagglutinin as the progenitor of the H3 hemagglutinin present on the current human Hong Kong viruses.