Content uploaded by Heesu Kim
Author content
All content in this area was uploaded by Heesu Kim on Jan 20, 2025
Content may be subject to copyright.
Article
Structural and functional analysis of the Nipah virus
polymerase complex
Graphical abstract
Highlights
dCryo-EM structure of the NiV L-P complex determined
dDocking studies with an inhibitor clarify mechanisms of
intrinsic NiV L resistance
dPalm insert, zinc fingers, and P4 extension are critical for NiV
L activity
dIntrusion loop plays an essential role in RNA replication
Authors
Side Hu, Heesu Kim, Pan Yang, ...,
Yuemin Bian, Rachel Fearns,
Jonathan Abraham
Correspondence
rfearns@bu.edu (R.F.),
jonathan_abraham@hms.harvard.edu
(J.A.)
In brief
The cryo-EM structure and mutational
analysis of the Nipah virus polymerase
complex identify features critical for RNA
replication and transcription with the
potential to aid in the development of
antivirals.
Hu et al., 2025, Cell 188, 1–16
February 6, 2025 ª2024 The Authors. Published by Elsevier Inc.
https://doi.org/10.1016/j.cell.2024.12.021 ll
Article
Structural and functional analysis
of the Nipah virus polymerase complex
Side Hu,
1,9
Heesu Kim,
2,9
Pan Yang,
1,9
Zishuo Yu,
1
Barbara Ludeke,
2
Shawna Mobilia,
2
Junhua Pan,
3
Margaret Stratton,
4
Yuemin Bian,
5
Rachel Fearns,
2,
*and Jonathan Abraham
1,6,7,8,10,
*
1
Department of Microbiology, Blavatnik Institute, Harvard Medical School, Boston, MA, USA
2
Department of Virology, Immunology & Microbiology, Boston University Chobanian & Avedisian School of Medicine, Boston, MA, USA
3
Biomedical Research Institute and School of Life and Health Sciences, Hubei University of Technology, Wuhan, China
4
Department of Biochemistry and Molecular Biology, University of Massachusetts, Amherst, MA, USA
5
School of Medicine, Shanghai University, Shanghai, China
6
Department of Medicine, Division of Infectious Diseases, Brigham & Women’s Hospital, Boston, MA, USA
7
Center for Integrated Solutions in Infectious Diseases, Broad Institute of Harvard and MIT, Cambridge, MA, USA
8
Howard Hughes Medical Institute, Boston, MA, USA
9
These authors contributed equally
10
Lead contact
*Correspondence: rfearns@bu.edu (R.F.), jonathan_abraham@hms.harvard.edu (J.A.)
https://doi.org/10.1016/j.cell.2024.12.021
SUMMARY
Nipah virus (NiV) is a bat-borne, zoonotic RNA virus that is highly pathogenic in humans. The NiV polymerase,
which mediates viral genome replication and mRNA transcription, is a promising drug target. We determined
the cryoelectron microscopy (cryo-EM) structure of the NiV polymerase complex, comprising the large protein
(L) and phosphoprotein (P), and performed structural, biophysical, and in-depth functional analyses of the NiV
polymerase. The L protein assembles with a long P tetrameric coiled-coil that is capped by a bundle of ⍺-he-
lices that we show are likely dynamic in solution. Docking studies with a known L inhibitor clarify mechanisms
of antiviral drug resistance. In addition, we identified L protein features that are required for both transcription
and RNA replication and mutations that have a greater impact on RNA replication than on transcription. Our
findings have the potential to aid in the rational development of drugs to combat NiV infection.
INTRODUCTION
Nipah virus (NiV) is an enveloped RNA virus in the family Para-
myxoviridae of the order Mononegavirales, the non-segmented
negative-strand RNA viruses (nsNSVs).
1
In humans, NiV infection
can lead to respiratory illness or encephalitis with a fatality rate
ranging from 40% to 70%.
2
Since the identification of NiV in
Malaysia in 1998, spillover events from bats into humans have
occurred almost annually in Bangladesh, and NiV has also
caused outbreaks in India and in the Philipines.
3–5
Although prior
outbreaks have been limited in size, NiV may have pandemic po-
tential because infected individuals can be asymptomatic, and
the virus can be transmitted from person to person with
droplet-based transmission through coughing.
4
There is no vac-
cine or antiviral against NiV infection, highlighting a gap in public
health preparedness. For these reasons, NiV is on the World
Health Organization Research & Development Blueprint list of
priority diseases for which there is an urgent need for acceler-
ated research and countermeasure development.
The NiV genome is a single strand of negative-sense RNA that
is transcribed and replicated by the viral polymerase. During
transcription, the polymerase generates capped and polyadeny-
lated monocistronic mRNAs, corresponding to each of the viral
genes.
6–9
During replication it produces full-length, encapsi-
dated replicative RNAs, which are uncapped and lack a poly A
tail.
7–9
The NiV polymerase comprises the large protein (L) and
phosphoprotein (P).
10,11
By analogy with the polymerases of
other nsNSVs, the L protein comprises five domains: the RNA-
dependent RNA polymerase (RdRp) domain, the capping
domain (CAP), the connector domain (CD), the methyltransfer-
ase (MTase) domain, and the C-terminal domain (CTD).
7–9,12
The P protein serves as an adaptor that allows the polymerase
to associate both with the encapsidated ribonucleoprotein
(RNP) template and with soluble nucleoprotein (N) protein that
is required to encapsidate replicative RNA as it is synthesized.
7–9
P contains an intrinsically disordered N-terminal domain (NTD), a
central oligomerization domain (OD), and a C-terminal X domain
(XD). The CTD of P binds the RNP template, and the NTD of P
binds soluble N protein for encapsidation.
13–16
Although several nsNSV polymerase structures are available,
how the features identified within them facilitate different poly-
merase functions is generally poorly defined, and it is not fully un-
derstood how the polymerase is regulated to ‘‘switch on’’ and
‘‘switch off’’ the different enzymatic activities that are required
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 ª2024 The Authors. Published by Elsevier Inc. 1
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
for mRNA transcription versus genome replication.
17
More com-
plete analyses of the relationship between polymerase structure
and function are necessary to clarify how the polymerase transi-
tions between different enzymatic activities and produces
different RNA products. Additionally, understanding the func-
tional properties of key features will help identify those that could
be leveraged as drug targets.
Here, we determined the 2.3 A
˚cryoelectron microscopy
(cryo-EM) structure of the NiV L-P complex, which reveals how
tetrameric P interacts with L. Structure-function analyses using
single-step and multi-cycle NiV minigenome assays identify fea-
tures of the polymerase that are required for transcription and/or
genome replication, providing information that has the potential
to aid rational design of antiviral molecules against NiV.
RESULTS
Expression and purification of NiV L-P
We co-expressed full-length NiV L and P in insect cells (Fig-
ure 1A), purified L-P complexes, and analyzed samples of puri-
fied complexes by SDS-PAGE analysis (Figure S1A) and liquid
chromatography-mass spectrometry (LC-MS/MS; Data S1).
Analysis of samples by mass photometry revealed a major
peak of 559 kDa, which likely represents L bound to four copies
of P (L-P
4
complex) (Figure 1B). The L-P complex was active in
an in vitro RNA synthesis assay using an oligonucleotide RNA
template containing the NiV promoter sequence and nucleoside
triphosphates (NTPs) (Figures 1C, 1D, and S1B).
Structure of the NiV L-P complex
We used single-particle cryo-EM to determine the structure of
the NiV L-P complex to a global resolution of 2.3 A
˚
(Figures S1C–S1E; Table S1). NiV L-P is shaped like a tobacco
pipe, with the stem formed by the tetrameric P OD (Figures 1E
and 1F). We observed interpretable density for L residues 5–
1463, encompassing the RdRp and CAP domains, and for the
P tetramer (P1–P4), of which we could observe different lengths
for individual protomers (P1: residues 479–584, P2: 479–708, P3:
477–579, and P4: 479–596) (Figure 1A). Although we observed
low-resolution features in 3D volumes during the earlier steps
of data processing that are consistent with the presence of the
L CD, MTase domain, and CTD, these features were lost during
additional steps of data processing (Figure S1C). Therefore, like
in the structures of the L proteins of respiratory syncytial virus
(RSV) and human metapneumovirus (HMPV) (Pneumoviridae),
the three L C-terminal globular domains in NiV L are probably
too flexible to be visualized by cryo-EM
18–23
(Data S2). For the
RdRp and CAP domains that could be resolved, root-mean-
square deviation values based on the Caranged from 1.7 to
4.7 A
˚between NiV and other nsNSV L proteins (Figure S1F), indi-
cating structural conservation, consistent with the generally
similar transcription and replication mechanisms of nsNSVs.
In the cryo-EM map, the N-terminal end of the OD of the
NiV P tetramer is capped with a mushroom-shaped density
(Figures 1E, S1C, and S1E). Focused refinement on this region
allowed us to resolve individual ⍺-helices, allowing us to model
the region based on a crystal structure of the NiV P OD.
15
The
⍺-helices form a bundle that folds back onto the outer aspects
of the coiled-coil core (Figure 1F), a feature observed in crystal
structures of the NiV and Sendai virus (SeV) (paramyxovirus)
phosphoproteins.
15,24,25
These prior studies relied on truncated
P constructs comprising only the OD, rather than full-length
P.
15,24,25
Observing the mushroom-shaped cap structure in the
NiV L-P complex reveals that this feature is present when full-
length P associates with L without the potential conformational
biases that could have been introduced by crystallization of P
OD in isolation.
Features of the NiV L protein
The NiV L RdRp domain has a conventional right-hand ‘‘fingers-
palm-thumb’’ organization containing seven motifs involved in
catalysis (A–G), like other nsNSV RdRp domains (Figures 2A
and 2B). Motifs A–E are in the palm while motifs F and G are in
the fingers. Motif C contains a ‘‘GDNE’’ motif (residues 831–
834), which is part of the active site (Figure S2A; Data S3). In
the unliganded active site, protrusion of the b-hairpin loop in
motif F in the fingers domain positions the side chain of R551
near (within 5 A
˚) motif C residue D832 (part of the GDNE motif;
Figure 2C). R551 is universally conserved in nsNSV polymerases
(Figure S2A; Data S3). Prediction of an RNA duplex- and nucle-
otide-bound NiV L with AlphaFold 3 (AF3)
26
suggests the side
chain of R551 interacts with the phosphates of the incoming
nucleotide during RNA elongation, while the side chains of
D832 (motif C) and D722 (motif A) coordinate active site metals
(Figures 2D and S2D). Interestingly, in the cryo-EM structure of
L-P complexes for parainfluenza virus type 3 (PIV3), visualized
without RNA or incoming nucleotide, the analogous motif F argi-
nine is also positioned toward the motif C catalytic aspartate and
an associated active site magnesium (Figures S2B, S2C, and
S2E),
27
supporting AF3 predictions of the NiV L active site.
In the assembled NiV L-P complex, there are channels for NTP
entry, template entry, template exit, and nascent RNA exit
(Figures 2E and 2F). The template entrance channel, identified
by analogy to the Ebola virus (EBOV) and RSV L protein-promoter
RNA-bound structures,
18,31
involves surfaces of the RdRp and
CAP domains (Figure 2F). The putative template exit channel,
identified by analogy with other viral RdRps, is within the CAP
domain (Figure 2F).
8
The putative nascent RNA exit channel is
formed by the RdRp domain and the CAP domain on the opposite
side of L relative to the template entrance channel.
A distinctive feature in the NiV RdRp domain, when compared
with that of most other nsNSV polymerases, lies in the region
found between NiV L residues 600–713, which is located imme-
diately N-terminal to motif A in the palm domain (Figures 2A and
2B). We did not observe cryo-EM density for residues 603–709
within this loop. Primary sequence alignments (Figure 2G; Data
S3)
32,33
show that this region of henipavirus and parahenipavirus
L proteins has additional sequence, ranging from 100 to > 200
amino acid residues in length, compared with other nsNSVs,
including most other paramyxoviruses. We refer to this addi-
tional sequence as the palm insert. Other than the large size
and general feature of being rich in lysine and asparagine
residues, there is substantial divergence in the henipavirus and
parahenipavirus palm insert sequences, although it is mostly
conserved among closely related viruses when they are exam-
ined as pairs (Figures 2H and S2F–S2J).
ll
OPEN ACCESS
2Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
The boundary residues that we could observe for the palm
insert, S602 and T710, are relatively close to each other in the
structure (only 10 A
˚apart) and are located near the putative
nascent RNA exit channel (Figure 2E). The proximity of the
N- and C-terminal regions of the loop and disorder in cryo-EM
maps suggests that the insert may form an appendage that
does not contribute to the fold of the RdRp domain. To test
this hypothesis, we generated NiV L that lacks residues 604–
704 (NiV L
D-P.I
.) and co-expressed it with P in insect cells. Puri-
fied particles examined by negative stain electron microscopy
had a similar shape to particles for the wild-type (WT) NiV L-P
complex (Figures 2I and S2K–S2M). These findings confirmed
that the palm insert does not contribute to the fold of the RdRp
domain.
The CAP domain
The CAP domain of nsNSV polymerases plays roles in RNA
synthesis initiation and elongation, in addition to catalyzing
the polyribonucleotidyltransferase (PRNTase) step of cap addi-
tion.
7,8,12,34
For vesicular stomatitis virus (VSV), the CAP domain
A B
C
D
E
F
Figure 1. Structure of the NiV L-P complex
(A) Domain organization of the NiV L and P proteins. The regions of tetrameric P protein that could be resolved vary in length depending on the protomer (P1–P4),
as indicated. P NTD, OD, CTD, and XD boundaries are indicated based on established boundaries.
13
(B) Mass photometry analysis of purified NiV L-P complex. The 559 kDa peak likely represents the L-P
4
complex, the 483 kDa peak is consistent with an L-P
3
complex, the 248 kDa peak may correspond to the L protein alone, and the 1098 kDa peak may represent L-P
4
dimers. The 135 kDa and 66 kDa peaks may
represent degradation products or contaminants. This experiment was performed twice, with representative data shown.
(C) Workflow diagram of the RNA synthesis assay. The template is the leader (le) sequence of NiV, nucleotides 1–17. The GTP tracer and its incorporation in the
product are in red.
(D) RNA synthesis assay with radioactive products migrated on a denaturing polyacrylamide gel. nt, nucleotide. This experiment was performed four times with
two different L-P preparations, with representative data shown. Note that some product bands migrated as doublets, and products longer than expected were
also generated (indicated with a vertical line). These are likely generated by polymerase stuttering on the template. See also Figure S1B.
(E) Cryo-EM density map of the NiV L-P complex with L domains and P protomers colored as indicated. Two views are shown.
(F) NiV L-P complex with L domains and P protomers colored as indicated.
See also Figure S1.
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 3
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
AB C
DEF
G H
I
Figure 2. Molecular features of the NiV L polymerase core
(A) NiV L RdRp domain. The fingers, palm, thumb, and NTD domains are shown. The palm active site GDNE motif residues are shown as sticks. The location of the
palm insert is indicated.
(B) The same view as in (A), with catalytic motifs A–G colored as indicated.
(C) NiV L RdRp active site in the cryo-EM structure. Active site residues D832 (motif C), and nearby conserved residues D722 (motif A) and R551 (motif F) are
shown as sticks.
(D) AlphaFold 3 (AF3)
26
prediction of the NiV L active site in the presence of a duplex RNA, GTP, and two magnesium ions (green spheres). D722 (motif A) and D832
(motif C) are predicted to coordinate metals, and R511 (motif F) is positioned to interact with the phosphate of the incoming nucleotide. Template and nascent
RNA are shown in dark and light gray, respectively. Parts of motifs A and D are semi-transparent for clarity. See Figure S2 D for predicted local distance difference
test (pLDDT) scores.
(legend continued on next page)
ll
OPEN ACCESS
4Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
also interacts with the template RNP.
35
Features that have been
identified in other nsNSV CAP domains include zinc-finger (ZF)
motifs, a priming loop, and an intrusion loop. The ZF motifs are
present in the L proteins of many but not all nsNSVs
(Figures 3A–3C and S3A–S3C; Data S3),
27,31,36–42
but their func-
tion is not known. The priming loop is thought to be involved in
stabilizing the RNA synthesis initiation complex, and the intru-
sion loop contains the catalytic histidine-arginine (HR) motif
required for PRNTase activity (Figures 3D and 3E).
7,8,12
The prim-
ing and intrusion loops have been captured in multiple distinct
conformations for different nsNSV polymerases, each thought
to represent a different functional state (Figures S3D–S3I). In
the cryo-EM map of the NiV L-P complex, most of the priming
loop and intrusion loop are not visible, likely due to flexibility (Fig-
ure 3F), although both loops could be predicted using AF3
(Figures 3G and S3J). The state captured with NiV L-P is most
similar to EBOV L-VP35 complexes visualized in the absence
of RNA, in which most of both loops were disordered, with the
loops becoming partially ordered in the presence of promoter
RNA and fully ordered in the presence of an RNA template-
primer duplex (Figures 3H–3J).
31,36
Structure of NiV P
In the NiV L-P complex, most of the N-terminal region of P
remains disordered, as is the case for other paramyxo-
viruses
27,37,41,42
(Figure 1A). The C-terminal regions of P fold
onto L by forming several extended arms, with one of the proto-
mers (designated P2 here) providing contacts through the XD
domain, which forms a bundle of three ⍺-helices (Figures 4A
and 4B). P2-XD helices ⍺-1 and ⍺-3 interact with L (Figure 4B).
P buries 3,478 A
˚
2
of surface area on NiV L. The P2 polypeptide
chain makes polar contacts as it courses toward the XD domain
(Figure 4C), and the extended arms of P1 and P4 form an exten-
sive network of polar interactions with the RdRp domain
(Figures 4D and 4E). As part of this interaction network, P1 inter-
acts with the RdRp domain through b-strand augmentation (Fig-
ure 4D). An extension in P4 interacts with the RdRp domain on
the opposite face that binds P2 XD, and P4 residues K583,
K587, and K589 make interactions that anchor a loop onto the
RdRp domain (Figure 4E).
Dynamics of L-associated phosphoprotein
We could only resolve 239 residues of P, approximately one-
third of the molecule, as is the case in other structures of L-P
complexes. The NiV polymerase structure contains the longest
P segment visualized in paramyxovirus L-P complexes, particu-
larly when the folded-back helices of the ⍺-helical cap structure
are considered, as the two bent helices add an additional 50 A
˚
to the 100 A
˚coiled core (Data S2). We hypothesized that the
network of polar interactions P makes with L would likely be
malleable, allowing the pose of the P OD stem to vary with
respect to the polymerase core. Consistent with such motion,
the distal (N-terminal tip) of P had the lowest local resolution,
with smeared-out features of the ⍺-helical cap structure map re-
gion (Figures 1E and S1E). Three-dimensional (3D) variability
analysis
43
of the cryo-EM dataset revealed a swiveling motion
of the P OD with respect to the polymerase core (Video S1).
We also performed 100 ns molecular dynamics simulations using
the cryo-EM structure as a starting point (Figure S4;Video S2;
Data S4). Root-mean-square fluctuations (RMSF) were much
higher for each of the P protomers than they were for the poly-
merase core (RdRp/CAP), with the highest fluctuations observed
in the N-terminal portions of the ⍺-helical cap structure.
Basis for NiV resistance to broad-spectrum inhibitor
GHP-88309 is a broad-spectrum, non-nucleoside inhibitor of
paramyxovirus polymerases with activity against PIV3, measles
virus (MeV), and SeV.
44–46
While GHP-88309 is active against
many paramyxoviruses, it is not active against NiV, with a half-
maximal effective concentration (EC
50
) of 314 mM in a minige-
nome assay (Figures S5A and S5B).
44
Photoaffinity labeling
experiments, resistance mapping, and in silico docking experi-
ments revealed that GHP-88309 binds in a cavity at the intersec-
tion of the RdRp and CAP domains in MeV L.
44,46
A histidine in
the predicted drug binding site (H1165) in NiV L has been impli-
cated in resistance,
44
and it was shown that introducing the
H1165Y substitution into NiV L, which mimics the residue that
is naturally present in susceptible polymerases (Figure 5A), ren-
ders NiV L susceptible to GHP-88309 inhibition, with the EC
50
value decreasing by sixty-fold (5.3 mM) (Figure S5B).
44
To better understand the basis for drug susceptibility
conferred by the H1165Y substitution, we performed in silico
docking experiments with GHP-88309. Given that PIV3 L is sus-
ceptible, we used the PIV3 L cryo-EM structure
27
as a starting
point for docking experiments. Docking of GHP-88309 with
PIV3 L revealed a low-energy binding pose that placed GHP-
88309 at the expected site at an interface between the CAP
and RdRp domains, with the isoquinoline ring interacting with
Y942 (PIV3 L numbering) and the drug surrounded by residues
implicated in drug resistance (Figure 5B). The pose is highly
akin to that predicted by Cox et al., who used the parainfluenza
virus type 5 (PIV5) structure as a starting point to generate a
homology model of MeV L.
44
We next performed docking experiments with the NiV L cryo-
EM structure in its WT form or containing the H1165Y substitu-
tion introduced in silico (Figures 5C and 5D). A low-energy
(E) NiV L-P complex shown as a ribbon diagram with transparent surface and channels shown as spheres (calculated by CAVER Web
28
). Entrance and exit
channels are indicated. Residues S602 and T710, near the two ends of the palm insert, are near the nascent RNA exit channel.
(F) Clipped surface of the NiV L-P complex showing the RNA synthesis channels in the same view as shown in (E).
(G) Sequence alignment of the palm insert region of different nsNSV L proteins generated using ESPript 3.0.
29
The NiV L palm insert (residues 603–709) is boxed in
red. See Table S2 for virus name abbreviations and GenBank accession numbers.
(H) Phylogenetic tree of the indicated nsNSVs based on L amino acid sequences generated using Clustal Omega.
30
(I) Negative stain electron microscopy 2D classes of purified NiV L-P complexes for wild-type (WT) L or an L mutant in which much of the palm insert (residues
604–704) (L-
DP.I.
) is deleted. P.I., palm insert. See also Figures S2K–S2M.
See also Figure S2.
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 5
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
ABC
DE
FG
HI J
Figure 3. Features of the NiV L CAP domain
(A and B) CAP domain ZF1 (A) and ZF2 (B). Zinc ions are shown as gray spheres. Asterisks indicate residues selected for mutational analysis.
(C) Sequence alignment of examples of nsNSV L CAP ZFs, which are present in the L proteins of the viruses indicated by the circles, generated using ESPript
3.0.
29
Numbering is based on NiV L. Triangles indicate residues that form ZFs. Residues within the purple box are alternative residues that form ZFs in VSV and
RABV. Asterisks indicate residues selected for mutational analysis.
(D and E) Sequence alignment of the priming loop (D) or intrusion loop (E) of NiV L and the indicated nsNSV L proteins. Triangles indicate positions in the intrusion
loop that were selected for mutagenesis in addition to the HR motif.
(F) The NiV L RdRp domain is shown in surface representation with a partially clipped surface, and the CAP domain is shown as a ribbon diagram. The disord ered
portions of the priming and intrusion loops are shown as dashed lines. Boundary residues are indicated.
(G) AF3
26
prediction of the NiV L RdRp and CAP domains generated in the absence of nucleic acids. Intrusion loop HR residues (H1347 and R1348) are indicated.
Yellow spheres indicate residues selected for mutational analysis.
(H–J) Cryo-EM structures of EBOV L-VP35 without RNA (PDB: 7YER)
36
(H), with promoter RNA (PDB: 8JSL)
31
(I), and with the addition of an RNA template-primer
duplex but no RNA density observed in maps (PDB: 7YES)
36
(J). Nucleic acids are shown as orange sticks. Intrusion loop HR residues are indicated.
See also Figure S3.
ll
OPEN ACCESS
6Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
pose was identified in NiV L at a binding site that is like that of
PIV3. Comparison of the PIV3, NiV L, and NiV L H1165Y docking
models suggests that NiV L is resistant to the inhibitor because
H1165 does not make a polar contact that Y1165 can make
with the ring nitrogen of GHP-88309 (Figures 5B–5D). Given
that residues near the drug binding site are otherwise generally
conserved (Figure S5C), GHP-88309 could serve as a starting
point to identify modified inhibitors that target this site in
the NiV L.
GHP-88309 was reported to inhibit de novo initiation of RNA
synthesis at the promoter.
44
However, inspection of the data
presented indicates inhibition becomes more pronounced as
the product length is increased, indicating that it inhibits a step
after initiation of RNA synthesis.
44
Consistent with this finding,
the position of GHP-88309 in our docking experiment (Figure 5E)
in comparison to AF3-modeling of RNA-bound NiV L (Figure 5F)
suggests that the inhibitor would sterically block exit of the tem-
plate from the polymerase, which would be expected to inhibit
polymerase progression and thus efficient elongation of the
RNA product.
Functional analysis of the HR motif, P4 extension, palm
insert, and ZFs
The transcription and genome replication processes performed
by the NiV polymerase each involve a different sequence of
events (reviewed in Kleiner and Fearns).
17
It is thought that dur-
ing transcription, the polymerase starts from position 1U of the
leader (le) promoter, synthesizes a small RNA that is released,
and then re-initiates RNA synthesis at a gene start (gs) signal.
17
Having initiated RNA synthesis at the gs signal, the polymerase
co-transcriptionally caps the pre-mRNA, methylates the cap,
and then elongates the pre-mRNA until it reaches a gene end
(ge) sequence, where it is directed to polyadenylate and then
release the mature mRNA.
6
Following mRNA release, the
AB
C
D
E
Figure 4. Interactions between NiV P and L RdRp
(A) Ribbon diagram of the NiV L-P complex. Interactions between L and tetrameric P occur over four main regions, which are indicated by dashed boxes with
detailed interactions provided in (B)–(E).
(B) Interactions between the P2-XD ⍺-helices and the L RdRp domain.
(C) Interactions between the P2 linker, which connects the P2 OD with the P2-XD, and the L RdRp domain.
(D) Interactions between the P1, P2, and P4 and the L RdRp domain.
(E) Interaction between the P4 peptide extension and the L RdRp domain. L residues mutated in functional assays (Q454 and C457) are indicated with an asterisk.
See also Figure S4.
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 7
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
polymerase can then locate the gs signal for the next gene. By
continuing this sequence of events, the polymerase produces
a series of subgenomic mRNAs, each with a 50methylguanosine
cap and 30polyadenylate tail.
17
During replication, the polymer-
ase again initiates at position 1U of the le promoter but elongates
beyond the end of the le region, fails to recognize the gene junc-
tion signals, and produces a full-length, uncapped, non-polya-
denylated antigenome, which is concurrently encapsidated by
the N protein. The encapsidated antigenome in turn serves as
a template for the synthesis of encapsidated genome RNA.
17
We sought to determine if key features identified in the cryo-
EM structure of the NiV L-P complex are required for transcrip-
tion and/or RNA replication.
We examined NiV L mutants using a cell-based minigenome
system, in which NiV minigenome and NiV N, P, and L protein
expression are driven by intracellular expression of T7 RNA po-
lymerase (Figures 6A and 6B).
10
The advantage of this system,
compared with in vitro biochemical assays, is that it reconsti-
tutes each of the different steps involved in transcription and
RNA replication in the order in which they would naturally occur
in a cellular environment. We used a dicistronic minigenome in
which the first gene contains chloramphenicol acetyltransferase
(CAT) reporter gene sequence, and the second gene contains a
Renilla luciferase reporter gene. We used a single-step minige-
nome that was limited to the antigenome synthesis step of
RNA replication due to a mutation in the trailer (tr) region that ab-
lated the promoter at the 30end of the antigenome (Figures 6A
and S6A–S6I). This single-step minigenome assay allows the ef-
fects of polymerase mutations on transcription to be assessed
independently of any effect that the mutations have on RNA
replication and vice versa (Figure 6A). For all assays, the WT
and mutant L proteins contained a C-terminal strep tag to allow
monitoring of L protein expression. Comparison of untagged and
strep-tagged WT L in a minigenome luciferase assay showed
that the tag had minimal effect on polymerase activity, with lucif-
erase expression from strep-tagged L being at 85% of untagged
AB C D
EF
Figure 5. Basis for resistance of NiV L to a broad-spectrum paramyxovirus inhibitor
(A) Sequence alignment of L protein residues 1164 to 1167 in the CAP domain for the indicated viruses. The broad-spectrum inhibitor GHP-88309 is active against
the viruses indicated by the filled circles.
44
(B–D) Docking of GHP-88309 into the predicted binding site in PIV3 L (B), NiV L (C), or NiV L H1165Y mutant (modeled in silico) (D). Predicted interactions or
distances between the ring nitrogen of the inhibitor and contacting amino acids are shown as black dashed lines. Residues with asterisks are sites of L sub-
stitutions that affect inhibitor susceptibility (see Figures S5B and S5C).
(E) Clipping surface showing the position of the GHP-88309 docked in NiV L. The RdRp and CAP domains are shown as surfaces and clipped to reveal template
entry and exit channels.
(F) Clipping surface showing the position of modeled template and nascent RNA in an AF3
26
prediction of NiV L bound to RNA and nucleotide. An incoming GTP
molecule shown as sticks for orientation, but note that the clipping surface does not include the NTP entrance channel, which is out of the plane of view. Active site
metals are not shown for clarity. See Figure S5D for pLDDT scores.
See also Figure S5.
ll
OPEN ACCESS
8Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
AB
CD
EF
GH
(legend on next page)
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 9
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
L levels (data not shown). As a negative control, we generated an
L plasmid that contains a D832A substitution in the RdRp active
site GDNE motif (Figures 2C and 2D) and should yield an inactive
polymerase.
Mutations were introduced into L to substitute the HR motif
in the intrusion loop of the CAP domain (H1347A/R1348A)
(Figures 3E and 3G), disrupt the interface with the P4 extension
that contacts the RdRp domain (Q454R/C457E) (Figure 4E),
delete the palm insert in the RdRp domain (D604–704) (Fig-
ure 2G), and individually disrupt ZF1 (C1236S/C1239S) or ZF2
(C1428S/C1429S) that are in the CAP domain (Figures 3A–3C).
Western blot analysis confirmed that all mutant L proteins were
efficiently expressed, although typically a cleavage product of
195 kDa was also detected (Figures S6J, S6K, S6M, and
S6N). This product is the appropriate size to represent the C-ter-
minal portion of L following cleavage at the palm insert region,
and consistent with this, the D604–704 mutant did not show ev-
idence of this cleavage product (Figure S6J). Both ZF mutants
were also consistently less prone to this cleavage (Figure S6K).
The ZF1 mutant (C1236S/C1239S) was consistently expressed
at a lower level than WT L and the other mutants (45% of WT
levels; Figures S6K and S6N), suggesting that this ZF helps to
maintain structural integrity of the L protein. Nonetheless, suffi-
cient levels of the ZF1 mutant L protein were expressed for func-
tional analysis.
RNA products generated by the L protein mutants were exam-
ined by primer extension analysis performed using a primer cor-
responding to sequence downstream of the first gs signal (Fig-
ure 6A). This primer would be able to detect antigenome RNA
that had been elongated beyond the le region into the first
gene, and mRNA initiated at the first gs signal. Analysis of RNA
initiated at the 30end of the le region (antigenome) and gs signal
(mRNA) showed that the L H1347A/R1348A, L D604-704, and
two ZF mutants generated no detectable RNAs, whereas the L
Q454R/C457E mutant initiated antigenome and mRNA at
31% and 34% of WT levels, respectively (Figures 6C and
6D). This decrease in transcription was also shown by northern
blot analysis of CAT mRNA (Figures 6E and 6F) and measure-
ment of luciferase activity (Figure S6P). Unfortunately, levels of
antigenome RNA generated from a single-step minigenome
could not be measured reliably by northern blotting due to a
co-migrating background band on some blots. While the primer
extension analysis shown in Figures 6C and 6D measures the
products of the early steps of replication, it does not measure
full-length replication products. To assess the effects of the mu-
tations on full-length replicative RNAs, we also examined the po-
lymerase mutants using a multi-cycle minigenome that had an
intact tr region so that the polymerase could perform multiple cy-
cles of RNA replication (Figure 6B). This analysis confirmed that
the L H1347A/R1348A, L D604–704, and ZF mutants were
completely defective in RNA replication and that the L Q454R/
C457E mutant was only partially active (Figures 6G and 6H).
Thus, each of the mutations inhibited transcription and RNA
replication, indicating that the L protein features that were
mutated are fundamental to polymerase activity.
Functional analysis of the intrusion loop
By analogy with rhabdoviruses, the HR motif is required for the
PRNTase activity of mRNA cap addition.
47–49
Thus, it was not
surprising that substitutions in the HR motif inhibited mRNA syn-
thesis (Figures 6C–6F). However, the L H1347A/R1348A mutant
was also completely defective in RNA replication (Figures 6C,
6D, 6G, and 6H). This result indicates that the L H1347A/
R1348A mutation either affects polymerase structure or inhibits
a core activity of the polymerase that is required for both tran-
scription and replication (in addition to inhibiting capping), or
the H1347A/R1348A mutation has another distinct effect on
RNA replication in addition to inhibiting capping. The HR motif
lies in the intrusion loop (Figures 3E–3G), which does not have
a well-defined role. To gain further insight into intrusion loop
Figure 6. Functional analysis of the L HR motif, P4 extension, palm insert, and zinc-finger motifs using a minigenome system
(A and B) Diagrams illustrating the structure of the minigenomes and the products that are generated during RNA replication and mRNA transcription. White and
black boxes indicate gene start and gene end signals, respectively. Green arrows indicate promoters. (A) shows a minigenome limited to a single-step of RNA
replication by a mutation in the 50trailer (tr) region (indicated with a lightning symbol). This mutation prevents the antigenome that is synthesized by the NiV
polymerase from acting as a template for further rounds of minigenome synthesis. The red arrow illustrates the positioning of the primer used for primer extension
analysis, and the red dotted line illustrates the cDNA that would be generated by primer extension. (B) shows a minigenome with an intact tr region that serves as a
template for multiple cycles of RNA replication, resulting in amplified levels of encapsidated minigenome RNA.
(C) Primer extension analysis of positive-sense RNA generated by the NiV polymerase in the single-step minigenome system, using the primer indicated in (A). The
upper and lower panels show a phosphorimage scan of the same gel, with the intervening region excised. Note that the cDNA bands for the gs-initiated products
sometimes migrated as doublet or triplet bands. This could be a consequence of the reverse transcriptase encountering the methylated cap on the mRNA.
(D) Quantification of the levels of 30le (gray bars) and gs (white bars) initiation products from replicates of the experiment shown in (C). The bars show the mean
and standard deviation from three independent experiments for each mutant, except for the H1347A/R1348A (HR) mutant (n=6) and the Q454R/C457E (P4)
mutant (n=5).
(E) Northern blot analysis of RNAs generated in the single-step minigenome system. The upper panel shows negative-sense minigenome template RNA
generated by T7 RNA polymerase. The lower panel shows the corresponding positive-sense antigenome RNA and CAT mRNA generated by the NiV polymerase.
Note that the antigenome RNA co-migrates with a background band.
(F) Quantification of levels of mRNA from replicates of the northern blots shown in (E). The bars show the mean and standard deviation (SD) from three inde-
pendent experiments for each mutant.
(G) Northern blot analysis of RNAs generated in the multi-cycle minigenome system shown in (B). The upper panel shows negative-sense minigenome template
RNA generated by T7 RNA polymerase and amplified by NiV polymerase. The lower panel shows the corresponding positive-sense antigenome RNA and CAT
mRNA generated by the NiV polymerase.
(H) Quantification of levels of antigenome (gray bars) and minigenome (hatched bars) from replicates of the northern blots shown in (G). The bars show the mean
and SD from three independent experiments for each mutant.
See also Figure S6
ll
OPEN ACCESS
10 Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
function, we generated five additional L mutants. Each of these
residues was individually substituted with alanine, and the
mutant L proteins were tested in the same set of assays as
described above.
Western blot analysis confirmed that each of the L mutants
was efficiently expressed relative to WT L protein (Figures S6L
and S6O). Primer extension analysis of the RNA generated by
the L mutants showed a range of phenotypes. Substitution of
T1341 and Q1355 had minimal effect on RNA initiated from the
30end of the le or the first gs signal; by contrast, substitution of
F1358 completely inhibited both antigenome and mRNA initia-
tion (Figures 7A and 7B). Substitution of N1344 and L1345 in-
hibited antigenome initiation from the 30end of the le region to
less than 20% of WT levels but had only a modest inhibitory
effect on mRNA initiation from the gs signal (Figures 7A and
7B; note that in Figure 7A the upper and lower panels are derived
from the same gel, allowing direct comparison of RNA levels
within the same reaction). Northern blot analysis of mRNAs
generated in the single-step minigenome assay confirmed that
the T1341A, N1344A, L1345A, and Q1355A L mutants were all
capable of mRNA transcription, albeit at moderately reduced
levels compared with WT polymerase (Figures 7C and 7D).
Luciferase assays showed a similar pattern of gene expression
by each of the mutants as was determined by measuring
mRNA levels with their activities being ranked as follows:
T1341A > Q1355A > N1344A > L1354A > H1347A/R1348A and
F1358A (Figure S6Q). Curiously, the levels of luciferase activity
for the T1341A, Q1355A, N1344A, and L1354A mutants relative
to WT L were slightly lower than the relative levels of mRNA
(compare Figure S6Q with Figures 7B and 7D). It is possible
that these intrusion loop mutations affect the nature of the cap
that is added and/or inhibit cap methylation, thus impairing
mRNA translation, although further research is required to test
these hypotheses.
Assessment of the L mutants in the multi-cycle minigenome
system confirmed that N1344A, L1345A, and F1358A mutants
were highly defective in RNA replication (Figures 7E and 7F).
These mutant L proteins also failed to generate detectable
mRNA in the multi-cycle minigenome system (Figures 7E
and 7F). This is likely because the inhibitory effect of the mu-
tations on RNA replication would have significantly reduced
the amount of properly encapsidated minigenome template
AB
CD
EF
Figure 7. Functional analysis of the intru-
sion loop using a NiV minigenome system
(A) Primer extension analysis of positive-sense
RNA generated by the NiV polymerase in the sin-
gle-step minigenome system, using the primer
indicated in Figure 6A. The upper and lower panels
show a phosphorimage scan of the same gel, with
the intervening region excised.
(B) Quantification of the levels of 30le (gray bars)
and gs (white bars) initiation products from repli-
cates of the experiment shown in (A). The bars
show the mean and SD from three independent
experiments for each mutant, except for the
H1347A/R1348A (HR) mutant (n=6).
(C) Northern blot analysis of RNAs generated in
the single-step minigenome system. The upper
panel shows negative-sense minigenome tem-
plate RNA generated by T7 RNA polymerase. The
lower panel shows the corresponding positive-
sense antigenome RNA and CAT mRNA gener-
ated by the NiV polymerase. Note that the anti-
genome RNA co-migrates with a background
band.
(D) Quantification of levels of mRNA from repli-
cates of the northern blots shown in (C). The bars
show the mean and SD from three independent
experiments for each mutant.
(E) Northern blot analysis of RNAs generated in the
multi-step minigenome system. The upper panel
shows negative-sense minigenome template RNA
generated by T7 RNA polymerase and amplified
by NiV polymerase. The lower panel shows the
corresponding positive-sense antigenome RNA
and CAT mRNA generated by the NiV polymerase.
(F) Quantification of levels of antigenome (gray
bars) and minigenome (hatched bars) from repli-
cates of the northern blots shown in (E). The bars
show the mean and SD from three independent
experiments for each mutant.
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 11
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
that was available in these reactions compared with the reac-
tions with WT L protein (note that there is a large difference in
the amount of encapsidated (nuclease-resistant) minigenome
template available if the minigenome can be replicated, e.g.,
compare the L WT + MC and L WT + SS samples in
Figures S6C and S6E). These results show that the intrusion
loop has two separate and distinguishable roles in mRNA tran-
scription and genome replication.
DISCUSSION
NiV depends on its L-P polymerase complex to perform all the
enzymatic activities required to generate capped and polyade-
nylated mRNAs and replicative RNAs. Here we determined the
structure of the L-P complex and performed functional analysis
of some features that are shared with other nsNSV polymerase
and features that have been specifically observed in the NiV
polymerase.
Both transcription and RNA replication are dependent on the
polymerase being able to associate with the template RNP com-
plex, dissociate N protein from the RNA, and feed the RNA into
the template entry channel. It is thought that P plays a key role
in this process. We detected a region of one of the P molecules,
P4, snaking along the surface of L on the opposite face of the po-
lymerase from P2-XD and close to the template entry channel
(Figure 4A). The Q454R/C457E double substitution, which we
used to disrupt the L interaction with P4, did not affect L protein
expression (Figures S6J and S6M) but reduced transcription and
RNA replication compared with WT levels (Figures 6 and S6R).
These findings suggest that this L-P interface plays an equivalent
role in both processes. The P4 extension binding close to the
template entrance channel might allow P4-XD to play a role in
helping to guide the RNP toward the template entrance channel.
Another feature that might be involved in guiding RNA is the
palm insert (amino acids 603–709), which is close to the putative
nascent RNA exit channel (Figures 2E–2G). Studies with other
paramyxovirus polymerases have shown this region is important
for polymerase activity but did not distinguish if it affected tran-
scription and/or replication.
27,50
Here, we show that the NiV palm
insert plays a key role during transcription and RNA replication
(Figures 6 and S6R). Interestingly, the corresponding palm insert
region of PIV3 L is also largely disordered, but the C-terminal
segment of the region forms a b-strand that augments a b-sheet
in the CTD (‘‘b-latch’’) that may restrict the conformation of the L
C-terminal globular domains.
27
In the NiV L-P structure, the palm
insert and the C-terminal globular domains are not visible; by
analogy to the PIV3 L structure, however, the C-terminal region
of the NiV L RdRp palm insert could interact with the C-terminal
globular domains.
The findings presented here demonstrate the multifunctional
nature of the NiV L protein CAP domain. By analogy with the
rhabdovirus polymerases, the CAP domain contains residues
necessary for the GTPase and PRNTase steps of cap addition
12
and is expected to be required for mRNA transcription. However,
studies presented here show the importance of the CAP domain
for RNA replication in addition to transcription. We examined the
roles of two ZF motifs in CAP, which are conserved among many
nsNSVs (Figures 3A–3C and S3A–S3C). As noted above, ZF1
appears to aid structural integrity of the L protein. However, mu-
tations in both ZF motifs reduced the relative amounts of a trun-
cated form of L, which is presumably generated by cleavage at
the palm insert (Figure S6K). This observation could suggest
direct or long-range interactions between the ZF motifs and
the palm insert or that the ZF motifs influence the conformational
dynamics of the RdRp core. Substitutions in each of the ZF mo-
tifs inhibited replication in addition to transcription (Figures 6 and
S6R), suggesting that they might play a role common to both
processes. Intriguingly, a structure of the VSV polymerase in as-
sociation with RNP within virions showed that it is oriented such
that the ZF motifs of the CAP are proximal to the N protein of the
RNP.
35
Thus, it is possible that the ZF motifs aid polymerase as-
sociation with the RNP template.
The other feature of the CAP domain that is required for both
transcription and replication is the intrusion loop. The intrusion
loop can be identified as a loop in all nsNSV polymerase struc-
tures that have been determined so far (Figures S3D–
S3I).
27,37–39,41,42
The intrusion and priming loops are less or-
dered in the NiV L protein structure than in other paramyxovirus
L protein structures (Figures 3F and S3D–S3F). Here, we exam-
ined the role of the intrusion loop with a panel of substitution mu-
tants. Two of the mutants that were tested were completely
defective in mRNA transcription, namely the H1347A/R1348A
and F1358A mutants (Figures 6,7, and S6R). The HR motif cat-
alyzes the PRNTase reaction required for mRNA cap addition,
12
and so the H1347A/R1348A mutation could inhibit transcription
by inhibiting cap addition. The mechanism of transcription inhibi-
tion by the F1358A mutation is not known. Both the HR motif and
F1358 also play a role in RNA replication (which does not involve
a cap addition step) (Figures 6,7, and S6R). Likewise, other mu-
tations in the intrusion loop, N1344A, L1345A, and Q1355A, also
inhibited RNA replication, with Q1355A having a more minor ef-
fect. Importantly, the N1344A and L1345A mutations had a more
profound inhibitory effect on RNA replication than mRNA tran-
scription (Figures 7A, 7B, and S6R).
While transcription and replication are distinct processes,
many of the initial steps are shared, including template binding,
promoter recognition, and RNA synthesis initiation at the 30end
of the le region and nascent RNA elongation. The observation
that the N1344A and L1345A mutants could perform each of
these steps during transcription suggests that they were defec-
tive in a step that is specific to RNA replication. These findings
are reminiscent of findings made with RSV polymerase, which
showed that substitution of the threonine residue of the GxxT
motif in the priming loop (T1267) or the arginine residue of
the HR motif of the intrusion loop (R1339) inhibited RNA replica-
tion.
51
Analysis of the abortive replicative RNAs generated by the
RSV L T1267A and R1339A mutants showed that both mutants
could initiate RNA replication at le position 1U, but rather than
elongating the products to make antigenome RNA, they released
the RNA before they reached the end of the le promoter region.
51
Given that elongation during RNA replication is thought to be
facilitated by concurrent encapsidation, it seems likely that mu-
tations in the intrusion loop (and the priming loop in the case of
RSV) either directly or indirectly inhibit encapsidation of the
nascent replicative RNA. We suggest that, given that the intru-
sion loop (including the residues we mutated for functional
ll
OPEN ACCESS
12 Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
analyses) is adjacent to the nascent RNA exit channel
(Figures S3K and S3L), it might help to guide the nascent RNA
appropriately for encapsidation.
Perhaps the most distinctive feature of the complex compared
with that of other nsNSV L-P structures is the ⍺-helical cap struc-
ture at the N-terminal end of the P OD (Data S2). Both cryo-EM
3D variability analysis and molecular dynamics analysis suggest
that the tip of the long coiled-coil that is capped by a bundle of
helices is flexible. This flexibility could be important during
RNA replication, in which it is thought that a region near the N ter-
minus of P molecules that are associated with L delivers unas-
sembled N protein onto the nascent RNA.
In summary, we have determined the structure of NiV L-P
complex and performed functional analyses to identify features
that play key roles in transcription and replication. These findings
enhance our understanding of nsNSV polymerases and the
mechanisms by which they engage in transcription or RNA repli-
cation and have the potential to allow for rational drug design
against NiV infection.
Limitations of the study
The in vitro RNA synthesis assay we performed in purified NiV
L-P complexes involves a naked RNA oligonucleotide template
rather than encapsidated RNA. While this assay is suitable for
showing if the NiV L-P complex is active, the polymerase might
behave differently on this template compared with an encapsi-
dated RNA template. While we could not resolve the C-terminal
globular domains of NiV L and the intrusion and priming loops,
likely due to flexibility, additional studies will be required to deter-
mine whether binding of ligands, including nucleotides and/or
nucleic acids, causes these regions to become ordered.
RESOURCE AVAILABILITY
Lead contact
Further information and requests for resources and reagents should be
directed to and will be fulfilled by the lead contact, Jonathan Abraham
(jonathan_abraham@hms.harvard.edu).
Materials availability
Reagents generated in this study are available from the lead contact upon
request with completed material transfer agreements.
Data and code availability
dProtein Data Bank (PDB) and Electron Microscopy Data Bank (EMDB)
identification numbers for the cryo-EM structures and maps reported
in this manuscript are available as of the date of publication. Identifica-
tion numbers are listed in the key resources table. The per-residue root-
mean-square-fluctuation (RMSF) values from MD simulations reported
in this study are provided in Data S4.
dThis paper does not report original code.
dAny additional information required to reanalyze the data reported in this
paper is available from the lead contact.
ACKNOWLEDGMENTS
Cryo-EM data were collected at the Harvard Cryo-EM Center for Structural
Biology at Harvard Medical School. We would like to thank Dr. Bernard
Moss (National Institutes of Health) for providing the pTM1 plasmid and Dr.
Klaus Conzelmann (Ludwig-Maximilians-University Munich) for providing the
BSR-T7 cells used for the minigenome assays. We thank Dr. Bridget Gollan
for her help with generating illustrations for the graphical abstract. Y.B.
received funding support from the National Natural Science Foundation of
China (Excellent Young Scientists Fund [Overseas] and 82404516) and the
Shanghai Municipal Commission of Education. This work was supported by
an award from the Bill & Melinda Gates Foundation (INV-040438) to R.F.
AUTHOR CONTRIBUTIONS
Conceptualization, S.H., H.K., P.Y., J.P., Y.B., R.F., and J.A.; investigation,
S.H., H.K., P.Y., Z.Y., B.L., S.M., J.P., Y.B., R.F., and J.A.; writing – original
draft, S.H., H.K., P.Y., Y.B., R.F., and J.A.; writing – review and editing, S.H.,
H.K., P.Y., Z.Y., B.L., S.M., J.P., M.S., Y.B., R.F., and J.A.; funding acquisition,
M.S., R.F., and J.A.
DECLARATION OF INTERESTS
R.F. is the recipient of a sponsored research agreement with Merck & Co.
STAR+METHODS
Detailed methods are provided in the online version of this paper and include
the following:
dKEY RESOURCES TABLE
dEXPERIMENTAL MODEL AND STUDY PARTICIPANT DETAILS
BInsect cells
BBacterial cells
BCell lines
dMETHOD DETAILS
BCloning and insect cell expression of NiV L-P complexes
BNiV L-P complex protein purification
BMass photometry analysis of NiV L-P protein
BIn vitro RNA synthesis assay
BCryo-EM sample preparation and data collection
BCryo-EM image processing
BModel building and refinement
B3D variability analysis
BNegative stain electron microscopy
BIn silico modeling, molecular dynamics simulations and molecular
docking
BAlphaFold 3 modeling
BDesign and cloning of the NiV minigenome system
BReconstitution of minigenome replication and transcription
BHarvesting of transfected cells for Western blot and luciferase
assays
BNorthern blot and hybridization
BPrimer extension analysis
BFigure preparation
dQUANTIFICATION AND STATISTICAL ANALYSIS
SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.cell.
2024.12.021.
Received: May 23, 2024
Revised: November 1, 2024
Accepted: December 17, 2024
Published: January 20, 2025
REFERENCES
1. Amarasinghe, G.K., Ayllo
´n, M.A., Ba
`o, Y., Basler, C.F., Bavari, S., Blasdell,
K.R., Briese, T., Brown, P.A., Bukreyev, A., Balkema-Buschmann, A., et al.
(2019). Taxonomy of the order Mononegavirales: update 2019. Arch. Virol.
164, 1967–1980. https://doi.org/10.1007/s00705-019-04247-4.
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 13
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
2. Faus-Cotino, J., Reina, G., and Pueyo, J. (2024). Nipah virus: A multidi-
mensional update. Viruses 16, 179. https://doi.org/10.3390/v16020179.
3. Gurley, E.S., Hegde, S.T., Hossain, K., Sazzad, H.M.S., Hossain, M.J.,
Rahman, M., Sharker, M.A.Y., Salje, H., Islam, M.S., Epstein, J.H., et al.
(2017). Convergence of humans, bats, trees, and culture in Nipah virus
transmission, Bangladesh. Emerg. Infect. Dis. 23, 1446–1453. https://
doi.org/10.3201/eid2309.161922.
4. Arunkumar, G., Chandni, R., Mourya, D.T., Singh, S.K., Sadanandan, R.,
Sudan, P., and Bhargava, B.; Nipah Investigators People and Health Study
Group (2019). Outbreak investigation of Nipah virus disease in Kerala, In-
dia, 2018. J. Infect. Dis. 219, 1867–1878. https://doi.org/10.1093/infdis/
jiy612.
5. Chadha, M.S., Comer, J.A., Lowe, L., Rota, P.A., Rollin, P.E., Bellini, W.J.,
Ksiazek, T.G., and Mishra, A. (2006). Nipah virus-associated encephalitis
outbreak, Siliguri, India. Emerg. Infect. Dis. 12, 235–240. https://doi.org/
10.3201/eid1202.051247.
6. Ogino, M., Green, T.J., and Ogino, T. (2024). The complete pathway for co-
transcriptional mRNA maturation within a large protein of a non-
segmented negative-strand RNA virus. Nucleic Acids Res. 52, 9803–
9820. https://doi.org/10.1093/nar/gkae659.
7. Ouizougun-Oubari, M., and Fearns, R. (2023). Structures and
mechanisms of nonsegmented, negative-strand RNA virus polymerases.
Annu. Rev. Virol. 10, 199–215. https://doi.org/10.1146/annurev-virology-
111821-102603.
8. Liang, B. (2020). Structures of the Mononegavirales polymerases. J. Virol.
94, e00175-20. https://doi.org/10.1128/JVI.00175-20.
9. Pyle, J.D., Whelan, S.P.J., and Bloyet, L.M. (2021). Structure and function
of negative-strand RNA virus polymerase complexes. Enzymes 50, 21–78.
https://doi.org/10.1016/bs.enz.2021.09.002.
10. Halpin, K., Bankamp, B., Harcourt, B.H., Bellini, W.J., and Rota, P.A.
(2004). Nipah virus conforms to the rule of six in a minigenome replication
assay. J. Gen. Virol. 85, 701–707. https://doi.org/10.1099/vir.0.19685-0.
11. Jordan, P.C., Liu, C., Raynaud, P., Lo, M.K., Spiropoulou, C.F., Symons,
J.A., Beigelman, L., and Deval, J. (2018). Initiation, extension, and termina-
tion of RNA synthesis by a paramyxovirus polymerase. PLoS Pathog. 14,
e1006889. https://doi.org/10.1371/journal.ppat.1006889.
12. Ogino, T., and Green, T.J. (2019). RNA Synthesis and capping by non-
segmented negative strand RNA viral polymerases: lessons from a proto-
typic virus. Front. Microbiol. 10, 1490. https://doi.org/10.3389/fmicb.
2019.01490.
13. Blocquel, D., Beltrandi, M., Erales, J., Barbier, P., and Longhi, S. (2013).
Biochemical and structural studies of the oligomerization domain of the Ni-
pah virus phosphoprotein: evidence for an elongated coiled-coil homo-
trimer. Virology 446, 162–172. https://doi.org/10.1016/j.virol.2013.07.031.
14. Bourhis, J.M., Yabukarski, F., Communie, G., Schneider, R., Volchkova,
V.A., Fre
´ne
´at, M., Ge
´rard, F.C., Ducournau, C., Mas, C., Tarbouriech, N.,
et al. (2022). Structural dynamics of the C-terminal X domain of Nipah
and Hendra viruses controls the attachment to the C-terminal tail of the
nucleocapsid protein. J. Mol. Biol. 434, 167551. https://doi.org/10.1016/
j.jmb.2022.167551.
15. Bruhn, J.F., Barnett, K.C., Bibby, J., Thomas, J.M.H., Keegan, R.M., Rig-
den, D.J., Bornholdt, Z.A., and Saphire, E.O. (2014). Crystal structure of
the Nipah virus phosphoprotein tetramerization domain. J. Virol. 88,
758–762. https://doi.org/10.1128/JVI.02294-13.
16. Schiavina, M., Salladini, E., Murrali, M.G., Tria, G., Felli, I.C., Pierattelli, R.,
and Longhi, S. (2020). Ensemble description of the intrinsically disordered
N-terminal domain of the Nipah virus P/V protein from combine d NMR and
SAXS. Sci. Rep. 10, 19574. https://doi.org/10.1038/s41598-020-76522-3.
17. Kleiner, V.A., and Fearns, R. (2024). How does the polymerase of non-
segmented negative strand RNA viruses commit to transcription or
genome replication? J. Virol. 98, e0033224. https://doi.org/10.1128/jvi.
00332-24.
18. Cao, D., Gao, Y., Chen, Z., Gooneratne, I., Roesler, C., Mera, C., D’Cunha,
P., Antonova, A., Katta, D., Romanelli, S., et al. (2024). Structures of the
promoter-bound respiratory syncytial virus polymerase. Nature 625,
611–617. https://doi.org/10.1038/s41586-023-06867-y.
19. Cao, D., Gao, Y., Roesler, C., Rice, S., D’Cunha, P., Zhuang, L., Slack, J.,
Domke, M., Antonova, A., Romanelli, S., et al. (2020). Cryo-EM structure of
the respiratory syncytial virus RNA polymerase. Nat. Commun. 11, 368.
https://doi.org/10.1038/s41467-019-14246-3.
20. Gilman, M.S.A., Liu, C., Fung, A., Behera, I., Jordan, P., Rigaux, P., Yse-
baert, N., Tcherniuk, S., Sourimant, J., Ele
´oue
¨t, J.F., et al. (2019). Structure
of the respiratory syncytial virus polymerase complex. Cell 179, 193–
204.e14. https://doi.org/10.1016/j.cell.2019.08.014.
21. Kleiner, V.A., O Fischmann, T., Howe, J.A., Beshore, D.C., Eddins, M.J.,
Hou, Y., Mayhood, T., Klein, D., Nahas, D.D., Lucas, B.J., et al. (2023).
Conserved allosteric inhibitory site on the respiratory syncytial virus and
human metapneumovirus RNA-dependent RNA polymerases. Commun.
Biol. 6, 649. https://doi.org/10.1038/s42003-023-04990-0.
22. Pan, J., Qian, X., Lattmann, S., El Sahili, A., Yeo, T.H., Jia, H., Cressey, T.,
Ludeke, B., Noton, S., Kalocsay, M., et al. (2020). Structure of the human
metapneumovirus polymerase phosphoprotein complex. Nature 577,
275–279. https://doi.org/10.1038/s41586-019-1759-1.
23. Yu, X., Abeywickrema, P., Bonneux, B., Behera, I., Anson, B., Jacoby, E.,
Fung, A., Adhikary, S., Bhaumik, A., Carbajo, R.J., et al. (2023). Structural
and mechanistic insights into the inhibition of respiratory syncytial virus
polymerase by a non-nucleoside inhibitor. Commun. Biol. 6, 1074.
https://doi.org/10.1038/s42003-023-05451-4.
24. Tarbouriec h, N., Curran, J., Ruigrok, R.W., and Burmeister, W.P. (2000).
Tetrameric coiled coil domain of Sendai virus phosphoprotein. Nat. Struct .
Biol. 7, 777–781. https://doi.org/10.1038/79013.
25. Jensen, M.R., Yabukarski, F., Communie, G., Condamine, E., Mas, C.,
Volchkova, V., Tarbouriech, N., Bourhis, J.M., Volchkov, V., Blackledge,
M., and Jamin, M. (2020). Structural description of the Nipah virus phos-
phoprotein and its interaction with STAT1. Biophys. J. 118, 2470–2488.
https://doi.org/10.1016/j.bpj.2020.04.010.
26. Abramson, J., Adler, J., Dunger, J., Evans, R., Green, T., Pritzel, A., Ron-
neberger, O., Willmore, L., Ballard, A.J., Bambrick, J., et al. (2024). Accu-
rate structure prediction of biomolecular interactions with AlphaFold 3.
Nature 630, 493–500. https://doi.org/10.1038/s41586-024-07487-w.
27. Xie, J., Ouizougun-Oubari, M., Wang, L., Zhai, G., Wu, D., Lin, Z., Wang,
M., Ludeke, B., Yan, X., Nilsson, T., et al. (2024). Structural basis for dimer-
ization of a paramyxovirus polymerase complex. Nat. Commun. 15, 3163.
https://doi.org/10.1038/s41467-024-47470-7.
28. Stourac, J., Vavra, O., Kokkonen, P., Filipovic, J., Pinto, G., Brezovsky, J.,
Damborsky, J., and Bednar, D. (2019). Caver Web 1.0: ident ification of tun-
nels and channels in proteins and analysis of ligand transport. Nucleic
Acids Res. 47, W414–W422. https://doi.org/10.1093/nar/gkz378.
29. Robert, X., and Gouet, P. (2014). Deciphering key features in prot ein struc-
tures with the new ENDscript server. Nucleic Acids Res. 42, W320–W324.
https://doi.org/10.1093/nar/gku316.
30. Sievers, F., Wilm, A., Dineen, D., Gibson, T.J., Karplus, K., Li, W., Lopez,
R., McWilliam, H., Remmert, M., So
¨ding, J., et al. (2011). Fast, scalable
generation of high-quality protein multiple sequence alignments using
Clustal Omega. Mol. Syst. Biol. 7, 539. https://doi.org/10.1038/msb.
2011.75.
31. Peng, Q., Yuan, B., Cheng, J., Wang, M., Gao, S., Bai, S., Zhao, X., Qi, J.,
Gao, G.F., and Shi, Y. (2023). Molecular mechanism of de novo replication
by the Ebola virus polymerase. Nature 622, 603–610. https://doi.org/10.
1038/s41586-023-06608-1.
32. McIlhatton, M.A., Curran, M.D., and Rima, B.K. (1997). Nucleotide
sequence analysis of the large (L) genes of phocine distemper virus and
canine distemper virus (corrected sequence). J. Gen. Virol. 78, 571–576.
https://doi.org/10.1099/0022-1317-78-3-571.
ll
OPEN ACCESS
14 Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
33. Poch, O., Blumberg, B.M., Bougueleret, L., and Tordo, N. (1990).
Sequence comparison of five polymerases (L proteins) of unsegmented
negative-strand RNA viruses: theoretical assignment of functional do-
mains. J. Gen. Virol. 71, 1153–1162. https://doi.org/10.1099/0022-1317-
71-5-1153.
34. Ogino, M., Green, T.J., and Ogino, T. (2022). GDP polyribonucleotidyl-
transferase domain of vesicular stomatitis virus polymerase regulates
leader-promoter escape and polyadenylation-coupled termination during
stop-start transcription. PLoS Pathog. 18, e1010287. https://doi.org/10.
1371/journal.ppat.1010287.
35. Si, Z., Zhou, K., Tsao, J., Luo, M., and Zhou, Z.H. (2022). Locations and
in situ structure of the polymerase complex inside the virion of vesicular
stomatitis virus. Proc. Natl. Acad. Sci. USA 119, e2111948119. https://
doi.org/10.1073/pnas.2111948119.
36. Yuan, B., Peng, Q., Cheng, J., Wang, M., Zhong, J., Qi, J., Gao, G.F., and
Shi, Y. (2022). Structure of the Ebola virus polymerase complex. Nature
610, 394–401. https://doi.org/10.1038/s41586-022-05271-2.
37. Abdella, R., Aggarwal, M., Okura, T., Lamb, R.A., and He, Y. (2020). Struc-
ture of a paramyxovirus polymerase complex reveals a unique methyl-
transferase-CTD conformation. Proc. Natl. Acad. Sci. USA 117, 4931–
4941. https://doi.org/10.1073/pnas.1919837117.
38. Horwitz, J.A., Jenni, S., Harrison, S.C., and Whelan, S.P.J. (2020). Struc-
ture of a rabies virus polymerase complex from electron cryo-microsco py.
Proc. Natl. Acad. Sci. USA 117, 2099–2107. https://doi.org/10.1073/pnas.
1918809117.
39. Jenni, S., Bloyet, L.M., Diaz-Avalos, R., Liang, B., Whelan, S.P.J., Grigor-
ieff, N., and Harrison, S.C. (2020). Structure of the vesicular stomatitis virus
L protein in complex with its phosphoprotein cofactor. Cell Rep. 30,53–
60.e5. https://doi.org/10.1016/j.celrep.2019.12.024.
40. Liang, B., Li, Z., Jenni, S., Rahmeh, A.A., Morin, B.M., Grant, T., Grigorieff,
N., Harrison, S.C., and Whelan, S.P.J. (2015). Structure of the L protein of
vesicular stomatitis virus from electron cryomicroscopy. Cell 162,
314–327. https://doi.org/10.1016/j.cell.2015.06.018.
41. Cong, J., Feng, X., Kang, H., Fu, W., Wang, L., Wang, C., Li, X., Chen, Y.,
and Rao, Z. (2023). Structure of the Newcastle disease virus L protein in
complex with tetrameric phosphoprotein. Nat. Commun. 14, 1324.
https://doi.org/10.1038/s41467-023-37012-y.
42. Li, T., Liu, M., Gu, Z., Su, X., Liu, Y., Lin, J., Zhang, Y., and Shen, Q.T.
(2024). Structures of the mumps virus polymerase complex via cryo-elec-
tron microscopy. Nat. Commun. 15, 4189. https://doi.org/10.1038/
s41467-024-48389-9.
43. Punjani, A., and Fleet, D.J. (2021). 3D variability analysis: resolving contin-
uous flexibility and discrete heterogeneity from single particle cryo-EM.
J. Struct. Biol. 213, 107702. https://doi.org/10.1016/j.jsb.2021.107702.
44. Cox, R.M., Sourimant, J., Toots, M., Yoon, J.J., Ikegame, S., Govindara-
jan, M., Watkinson, R.E., Thibault, P., Makhsous, N., Lin, M.J., et al.
(2020). Orally efficacious broad-spectrum allosteric inhibitor of paramyxo-
virus polymerase. Nat. Microbiol. 5, 1232–1246. https://doi.org/10.1038/
s41564-020-0752-7.
45. Cox, R.M., Wolf, J.D., Lieberman, N.A., Lieber, C.M., Kang, H.J., Sticher,
Z.M., Yoon, J.J., Andrews, M.K., Govindarajan, M., Krueger, R.E., et al.
(2024). Therapeutic mitigation of measles-like immune amnesia and exac-
erbated disease after prior respiratory virus infections in ferrets. Nat. Com-
mun. 15, 1189. https://doi.org/10.1038/s41467-024-45418-5.
46. Wolf, J.D., Sirrine, M.R., Cox, R.M., and Plemper, R.K. (2024). Structural
basis of paramyxo- and pneumovirus polymerase inhibition by non-nucle-
oside small-molecule antivirals. Antimicrob. Agents Chemother. 68,
e0080024. https://doi.org/10.1128/aac.00800-24.
47. Ogino, T., and Banerjee, A.K. (2007). Unconventional mechanism of mRNA
capping by the RNA-dependent RNA polymerase of vesicular stomatitis
virus. Mol. Cell 25, 85–97. https://doi.org/10.1016/j.molcel.2006.11.013.
48. Ogino, T., Yadav, S.P., and Banerjee, A.K. (2010). Histidine-mediated RNA
transfer to GDP for unique mRNA capping by vesicular stomatitis virus
RNA polymerase. Proc. Natl. Acad. Sci. USA 107, 3463–3468. https://
doi.org/10.1073/pnas.0913083107.
49. Ogino, T., and Banerjee, A.K. (2010). The HR motif in the RNA-dependent
RNA polymerase L protein of Chandipura virus is required for unconven-
tional mRNA-capping activity. J. Gen. Virol. 91, 1311–1314. https://doi.
org/10.1099/vir.0.019307-0.
50. Duprex, W.P., Collins, F.M., and Rima, B.K. (2002). Modulating the func-
tion of the measles virus RNA-dependent RNA polymerase by insertion
of green fluorescent protein into the open reading frame. J. Virol. 76,
7322–7328. https://doi.org/10.1128/jvi.76.14.7322-7328.2002.
51. Braun, M.R., Deflube
´, L.R., Noton, S.L., Mawhorter, M.E., Tremaglio, C.Z.,
and Fearns, R. (2017). RNA elongation by respiratory syncytial virus poly-
merase is calibrated by conserved region V. PLoS Pathog. 13, e1006803.
https://doi.org/10.1371/journal.ppat.1006803.
52. Pettersen, E.F., Goddard, T.D., Huang, C.C., Couch, G.S., Greenblatt,
D.M., Meng, E.C., and Ferrin, T.E. (2004). UCSF Chimera–a visualization
system for exploratory research and analysis. J. Comput. Chem. 25,
1605–1612. https://doi.org/10.1002/jcc.20084.
53. Goddard, T.D., Huang, C.C., Meng, E.C., Pettersen, E.F., Couch, G.S.,
Morris, J.H., and Ferrin, T.E. (2018). UCSF ChimeraX: meeting modern
challenges in visualization and analysis. Protein Sci. 27, 14–25. https://
doi.org/10.1002/pro.3235.
54. Adams, P.D., Afonine, P.V., Bunko
´czi, G., Chen, V.B., Echols, N., Headd,
J.J., Hung, L.W., Jain, S., Kapral, G.J., Grosse Kunstleve, R.W., et al.
(2011). The Phenix software for automated determination of macromolec-
ular structures. Methods 55, 94–106. https://doi.org/10.1016/j.ymeth.
2011.07.005.
55. Mastronarde, D.N. (2005). Automated electron microscope tomography
using robust prediction of specimen movements. J. Struct. Biol. 152,
36–51. https://doi.org/10.1016/j.jsb.2005.07.007.
56. Zheng, S.Q., Palovcak, E., Armache, J.P., Verba, K.A., Cheng, Y., and
Agard, D.A. (2017). MotionCor2: anisotropic correction of beam-induced
motion for improved cryo-electron microscopy. Nat. Methods 14,
331–332. https://doi.org/10.1038/nmeth.4193.
57. Rohou, A., and Grigorieff, N. (2015). CTFFIND4: fast and accurate defocus
estimation from electron micrographs. J. Struct. Biol. 192, 216–221.
https://doi.org/10.1016/j.jsb.2015.08.008.
58. Zivanov, J., Nakane, T., Forsberg, B.O., Kimanius, D., Hagen, W.J., Lin-
dahl, E., and Scheres, S.H. (2018). New tools for automated high-resolu-
tion cryo-EM structure determination in RELION-3. eLife 7, e42166.
https://doi.org/10.7554/eLife.42166.
59. Punjani, A., Rubinstein, J.L., Fleet, D.J., and Brubaker, M.A. (2017). cryo-
SPARC: algorithms for rapid unsupervised cryo-EM structure determina-
tion. Nat. Methods 14, 290–296. https://doi.org/10.1038/nmeth.4169.
60. Wagner, T., Merino, F., Stabrin, M., Moriya, T., Antoni, C., Apelbaum, A.,
Hagel, P., Sitsel, O., Raisch, T., Prumbaum, D., et al. (2019). SPHIRE-crY-
OLO is a fast and accurate fully automated particle picker for cryo-EM.
Commun. Biol. 2, 218. https://doi.org/10.1038/s42003-019-0437-z.
61. Sanchez-Garcia, R., Gomez-Blanco, J., Cuervo, A., Carazo, J.M., Sor-
zano, C.O.S., and Vargas, J. (2021). DeepEMhancer: a deep learning so-
lution for cryo-EM volume post-processing. Commun. Biol. 4, 874.
https://doi.org/10.1038/s42003-021-02399-1.
62. Emsley, P., Lohkamp, B., Scott, W.G., and Cowtan, K. (2010). Features
and development of coot. Acta Crystallogr. D Biol. Crystallogr. 66,
486–501. https://doi.org/10.1107/S0907444910007493.
63. Williams, C.J., Headd, J.J., Moriarty, N.W., Prisant, M.G., Videau, L.L.,
Deis, L.N., Verma, V., Keedy, D.A., Hintze, B.J., Chen, V.B., et al. (2018).
MolProbity: more and better reference data for improved all-atom struc-
ture validation. Protein Sci. 27, 293–315. https://doi.org/10.1002/
pro.3330.
64. Kelley, L.A., Mezulis, S., Yates, C.M., Wass, M.N., and Sternberg, M.J.E.
(2015). The Phyre2 web portal for protein modeling, prediction and anal-
ysis. Nat. Protoc. 10, 845–858. https://doi.org/10.1038/nprot.2015.053.
ll
OPEN ACCESS
Cell 188, 1–16, February 6, 2025 15
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
65. Sastry, G.M., Adzhigirey, M., Day, T., Annabhimoju, R., and Sherman, W.
(2013). Protein and ligand preparation: parameters, protocols, and influ-
ence on virtual screening enrichments. J. Comput. Aided Mol. Des. 27,
221–234. https://doi.org/10.1007/s10822-013-9644-8.
66. Bowers, K.J., Chow, E., Xu, H., Dror, R.O., Eastwood, M.P., Gregersen,
B.A., Klepeis, J.L., Kolossvary, I., Moraes, M.A., Sacerdoti, F.D., et al.
(2006). Scalable algorithms for molecular dynamics simulations on com-
modity clusters. In Proceedings of the 2006 ACM/IEEE Conference on
Supercomputing, p. 84.
67. Kwon, J.J., Hajian, B., Bian, Y., Young, L.C., Amor, A.J., Fuller, J.R., Fra-
ley, C.V., Sykes, A.M., So, J., Pan, J., et al. (2022). Structure-function anal-
ysis of the SHOC2-MRAS-PP1C holophosphatase complex. Nature 609,
408–415. https://doi.org/10.1038/s41586-022-04928-2.
68. Jorgensen, W.L., Maxwell, D.S., and Tirado-Rives, J. (1996). Development
and testing of the OPLS all-atom force field on conformational energetics
and properties of organic liquids. J. Am. Chem. Soc. 118, 11225–11236.
https://doi.org/10.1021/ja9621760.
69. Fuerst, T.R., Earl, P.L., and Moss, B. (1987). Use of a hybrid vaccinia virus-
T7 RNA polymerase system for expression of target genes. Mol. Cell. Biol.
7, 2538–2544. https://doi.org/10.1128/mcb.7.7.2538-2544.1987.
70. Buchholz, U.J., Finke, S., and Conzelmann, K.K. (1999). Generation of
bovine respiratory syncytial virus (BRSV) from cDNA: BRSV NS2 is not
essential for virus replication in tissue culture, and the human RSV leader
region acts as a functional BRSV genome promoter. J. Virol. 73, 251–259.
https://doi.org/10.1128/JVI.73.1.251-259.1999.
71. Noton, S.L., Aljabr, W., Hiscox, J.A., Matthews, D.A., and Fearns, R.
(2014). Factors affecting de novo RNA synthesis and back-priming by
the respiratory syncytial virus polymerase. Virology 462–463, 318–327.
https://doi.org/10.1016/j.virol.2014.05.032.
72. Kucukelbir, A., Sigworth, F.J., and Tagare, H.D. (2014). Quantifying the
local resolution of cryo-EM density maps. Nat. Methods 11, 63–65.
https://doi.org/10.1038/nmeth.2727.
ll
OPEN ACCESS
16 Cell 188, 1–16, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
STAR+METHODS
KEY RESOURCES TABLE
REAGENT or RESOURCE SOURCE IDENTIFIER
Antibodies
b-tubulin rabbit polyclonal antibody LI-COR Cat# NC9872627
Strep-Tag Classic antibody (Strep-tag II) Bio-Rad Cat# MCA2489, RRID: AB_609795
IRDye 680RD Donkey anti-rabbit IgG secondary antibody LI-COR Cat# 926-68073, RRID: AB_10954442
IRDye 800CW Goat anti-mouse IgG secondary antibody LI-COR Cat# 926-32210, RRID: AB_621842
Bacterial and virus strains
MAX Efficiency DH10Baccells Thermo Fisher Scientific Cat# 10361012
Chemicals, peptides, and recombinant proteins
Biotin Sigma-Aldrich Cat # B4501-5G
cOmplete, Mini, EDTA-free protease inhibitor cocktail Millipore Sigma Cat# 11836170001
Dithiothreitol (DTT) Sigma-Aldrich Cat# 10708984001
Strep-TactinXT 4Flowhigh-capacity resin IBA Lifesciences Cat# 2-5030-010
NiV (Bangladesh strain) L (full-length, with C-terminal
triple strep tag)-P (full-length, with N-terminal
polyhistidine tag)
This paper N/A
NiV L (Bangladesh stain) L (full-length, with residues
604–704 deleted -triple strep-tag)- P (full-length,
with N-terminal polyhistidine tag)
This paper N/A
[a
32
P] GTP, 250 mCi Revvity Cat# BLU006H250UC
NTP 100 mM each Thermo Fisher Scientific Cat# R0481
Calf intestinal alkaline phosphatase New England Biolabs Cat# M0525
Lipofectamine 3000 Transfection reagent Invitrogen Cat# L3000015
Aprotinin Thermo Fisher Scientific Cat# J11388MA
Micrococcal nuclease S7 Roche Cat# 10107921001
[ɣ
32
P] ATP, 250mCi Revvity Cat #BLU002A250UC
Sensiscript reverse transcriptase QIAGEN Cat# 205211
Moloney murine leukemia virus reverse transcriptase Promega Cat# M1701
AccuGel 19:1 (40%) National Diagnostics Cat# EC-850
Tris-Borate-EDTA Fisher BioReagents Cat# BP1333-1
Urea Sigma Cat# U6504
T4 Polynucleotide kinase New England Biolabs Cat# M0201
Critical commercial assays
Promega Dual-Luciferase Reporter Assay System Promega Cat# E1910
Monarch total RNA Miniprep kit New England Biolabs Cat# T2010S
Deposited data
Cryo-EM map of NiV L-P complex This paper EMD-44465
Model of NiV L-P complex This paper PDB ID: 9BDQ
Experimental models: Cell lines
Sf9 cells (Spodoptera frugiperda) ThermoFisher Cat# 11496015, RRID: CVCL_JF76
BSR-T7 cells Dr. Klaus Conzelmann N/A
Oligonucleotides
NiV (Bangladesh strain) leader sequence 1-17 RNA
(5’ AUUUUCCCUUGUUUGGU 3’)
Integrated DNA technologies N/A
NiV (Bangladesh strain) positions 90-109 primer
(5’ GATATATTTCTTGAGGATCC 3’), PAGE purified
Invitrogen N/A
(Continued on next page)
ll
OPEN ACCESS
Cell 188, 1–16.e1–e8, February 6, 2025 e1
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
Continued
REAGENT or RESOURCE SOURCE IDENTIFIER
DynaMarker prestain marker for RNA high Diagnocine Cat# DM260S
Single-stranded DNA ladder Simplex Sciences Cat# ss20
Recombinant DNA
pFastbac Dual vector Thermofisher Cat# 10712024
NiV (Bangladesh strain) L (full-length, with
C-terminal triple strep tag)-P (full-length,
with N-terminal polyhistidine tag)
in pFastBac dual vector
This paper N/A
NiV (Bangladesh stain) L (full-length,
with residues 604–704 deleted triple
strep-tag-P (full-length, with
N-terminal polyhistidine tag)
in pFastBac dual vector
This paper N/A
Multi-cycle minigenome with CAT
and Renilla luciferase reporter gene
This paper N/A
Single-step minigenome with a deletion
of promoter element 2 within the L
nontranslated region and substitutions
at position 1-4 relative to the
5’ end of the trailer region
This paper N/A
pTM1 plasmid Dr. Bernard Moss N/A
Codon optimized NiV (Bangladesh strain)
L open reading frames in modified pTM1
containing T7 promoter and internal
ribosome entry site with Strep tag on C terminus
This paper N/A
Codon optimized NiV (Bangladesh strain) P open
reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry site
This paper N/A
Codon optimized NiV (Bangladesh strain) N open
reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry site
This paper N/A
NiV (Bangladesh strain) L (with D832A) open reading
frames in modified pTM1 containing T7 promoter
and internal ribosome entry site with
Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with H1347A and R1348A)
open reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry
site with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with Q454R and
C457E in pTM1) open reading frames in
modified pTM1 containing T7 promoter
and internal ribosome entry site with
Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with residues
604–704 deleted) open reading frames in
modified pTM1 containing T7 promoter
and internal ribosome entry site
with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with C1236S and
C1239S) open reading frames in modified
pTM1 containing T7 promoter and internal
ribosome entry site with Strep tag on C terminus
This paper N/A
(Continued on next page)
ll
OPEN ACCESS
e2 Cell 188, 1–16.e1–e8, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
Continued
REAGENT or RESOURCE SOURCE IDENTIFIER
NiV (Bangladesh strain) L (with C1428S and
C1429S) open reading frames in modified
pTM1 containing T7 promoter and internal
ribosome entry site with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with T1341A) open
reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry site
with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with N1344A) open
reading frames in modified pTM1 containing T7
promoter and internal ribosome entry site
with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with L1345A) open
reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry
site with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with Q1355A) open
reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry
site with Strep tag on C terminus
This paper N/A
NiV (Bangladesh strain) L (with F1358A) open
reading frames in modified pTM1 containing
T7 promoter and internal ribosome entry
site with Strep tag on C terminus
This paper N/A
Software and algorithms
UCSF Chimera 1.15 Pettersen et al.
52
https://www.cgl.ucsf.edu/chimera/,
RRID:SCR_004097
UCSF ChimeraX 1.5 Goddard et al.
53
https://www.cgl.ucsf.edu/chimerax/,
RRID:SCR_015872
PyMOL 2.5.5 The PyMOL Molecular Graphics
System, Version 3.0
Schro
¨dinger, LLC.
https://pymol.org/2/, RRID:SCR_000305
Phenix 1.20.1-4487 Adams et al.
54
https://www.phenix-online.org,
RRID: SCR_014224
SerialEM 4.1-beta Mastronarde
55
http://bio3d.colorado.edu/SerialEM/,
RRID: SCR_017293
MotionCor2 1.6.4 Zheng et al.
56
https://emcore.ucsf.edu/cryoem-software,
RRID: SCR_016499
CTFFind4 4.1.14 Rohou and Grigorieff
57
http://grigoriefflab.janelia.org/ctffind4,
RRID: SCR_016732
Relion 3.1.1 Zivanov et al.
58
https://www3.mrc-lmb.cam.ac.uk/
relion/index.php/Main_Page,
RRID: SCR_016274
cryoSPARC v4.5.1 Punjani et al.
59
https://cryosparc.com/,
RRID: SCR_016501
cryOLO 1.9.9 Wagner et al.
60
https://cryolo.readthedocs.io/en/stable/
index.html, RRID: SCR_018392
DeepEMhancer 20220530_cu10 Sanchez-Garcia et al.
61
https://github.com/rsanchezgarc/
deepEMhancer, RRID: SCR_022573
Coot 0.9 Emsley et al.
62
http://www2.mrc-lmb.cam.ac.uk/
personal/pemsley/coot/,
RRID: SCR_014222
MolProbity 4.5.2 Williams et al.
63
http://molprobity.biochem.duke.edu,
RRID: SCR_014226
(Continued on next page)
ll
OPEN ACCESS
Cell 188, 1–16.e1–e8, February 6, 2025 e3
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
EXPERIMENTAL MODEL AND STUDY PARTICIPANT DETAILS
Insect cells
Sf9 cells (Spodoptera frugiperda) (Thermo Fisher Scientific Cat# 11496015) cells used for recombinant NiV L-P production were
maintained in Sf-900II SFM (Thermo Fisher Scientific Cat# 10902088) media at 27C.
Bacterial cells
Escherichia coli MAX Efficiency DH10Baccells (Thermo Fisher Scientific Cat# 10361012) were used for bacmid production to
produce recombinant baculoviruses.
Cell lines
BSR-T7 cells were maintained Glasgow’s MEM (Thermo Fisher Scientific Cat# 11710035) with 10% (v/v) Fetal Bovine Serum (Thermo
Fisher Scientific Cat# A5256701), 1% (v/v) MEM Non-Essential Amino Acids Solution (Thermo Fisher Scientific Cat# 11140050), 1%
(v/v) GlutaMAX(Thermo Fisher Scientific Cat# 35050061), and 1 mg/ml of Geneticin (G418 sulfate) (Thermo Fisher Scientific Cat#
10131035) at 37 C, 5–10% CO
2
, and 100% relative humidity.
METHOD DETAILS
Cloning and insect cell expression of NiV L-P complexes
The coding sequence of NiV L (GenBank: AAY43917.1) and P (GenBank: AAY43912.1) were synthesized and codon-optimized for
Bac-to-Bac expression system using pFastBac Dual transfer vector. The sequences of NiV L and P were fused with a C-terminal
triple-strep tag after a TEV protease cleavage site and an N-terminal His
6
tag followed by a TEV protease cleavage site, respectively.
NiV L was inserted downstream of the polyhedrin promoter and P was inserted downstream of the p10 promoter in the pFastBac
Dual. Recombinant bacmid containing NiV L and P genes was isolated after transforming into MAX Efficiency DH10Baccompetent
cells (Thermo Fisher Scientific Cat# 10361012) following the user guide of the Bac-to-Bac Baculovirus Expression System (Invitro-
gen). Viral stock generated from purified bacmids was amplified and used for protein expression. Two liters of Spodoptera frugiperda
9 (Sf9) cells (Thermo Fisher Scientific Cat# 11496015) maintained in GibcoSf-900TM II SFM media (Thermo Fisher Scientific Cat#
10902104) at a density of 2.5x10
6
cells/ml were infected with amplified viruses to co-express NiV L and P proteins at 27C for 72 h.
NiV L-P complex protein purification
The pellet of Sf9 cells expressing the NiV L-P complex was resuspended in lysis buffer (50 mM Tris, pH 7.5, 500 mM NaCl, 10% (v/v)
glycerol, 6 mM MgSO
4
, 1 mM dithiothreitol (DTT), 1% (v/v) Triton X-100) supplemented with cOmplete, EDTA-free protease inhib-
itor cocktail (Sigma-Aldrich Cat# 11836170001), carried out for 5 min at 4C. After high-speed centrifugation at 50,000 g for 2 h at 4C,
the supernatant containing the target proteins was incubated with Strep-TactinXT Sepharose resin (IBA Lifesciences Cat#
2-5030-010) at 4C for over 1 h and then loaded onto a gravity column. The beads were washed using buffer A (50 mM Tris pH
Continued
REAGENT or RESOURCE SOURCE IDENTIFIER
Refeyn DISCover
MP
v2.3 REFEYN https://www.refeyn.com/
Maestro 2024–3 Schrodinger https://www.schrodinger.com/
maestro, RRID: SCR_016748
Image Studio Lite version 5.2 LI-COR https://www.licor.com/bio/
image-studio-lite/,
RRID: SCR_013715
Prism 10 version 10.3.1 GraphPad Prism https://www.graphpad.com,
RRID: SCR_002798
UNICORNversion 7.8 Cytiva https://www.cytivalifesciences.com/
en/us/shop/chromatography/
software/unicorn-7-p-05649
AlphaFold Server Abramson et al.
26
https://deepmind.google/technologies/
alphafold/alphafold-server/,
RRID: SCR_025885
CAVER web version 1.2 Stourac et al.
28
https://loschmidt.chemi.muni.cz/
caverweb/
Clustal Omega Sievers et al.
30
https://www.ebi.ac.uk/jdispatcher/
msa/clustalo, RRID: SCR_001591
ll
OPEN ACCESS
e4 Cell 188, 1–16.e1–e8, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
8.0, 500 mM NaCl, 10% glycerol (v/v), 2 mM MgSO
4
, 1 mM DTT), and the bound proteins were eluted using 50 mM biotin in buffer A.
The eluates were analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and the putative L and P
bands were subjected to LC-MS/MS analysis (Harvard Center for Mass Spectrometry) to confirm their identities. Specifically, the
NiV L band that is near the 250 kDa marker band on SDS-PAGE (Figure S1A) was cut and sent for trypsin digestion, followed by
LC-MS/MS sequencing. Fragments of NiV L detected by LC-MS/MS are highlighted in gray in Data S1. NiV L has a coverage
over 82%. The sequence C-terminal of residue 2244 is a glycine-serine linker and triple-strep tag. Additionally, the NiV P band
near the 100 kDa marker band on SDS-PAGE (Figure S1A) was cut and sent for LC-MS/MS sequencing. Fragments detected are
highlighted in gray in Data S1. NiV P has a coverage over 94%. The sequence N-terminal of residue 1 is a histidine tag and
serine-glycine linker.
The fractions containing L-P protein were further purified using a size-exclusion column (Superose 6 Increase 10/300 GL, GE
healthcare) equilibrated with size exclusion chromatography (SEC) buffer (20 mM Tris, pH 7.5, 500 mM NaCl, 2 mM MgSO
4
,
1 mM DTT) using an A
¨KTA pureprotein purification system (Cytiva) with UNICORNversion 7.8. The peak fractions near 11 ml
were pooled and concentrated. Protein homogeneity was examined by negative stain electron microscopy and mass photometry
(see below).
Mass photometry analysis of NiV L-P protein
Mass photometry analyses were carried out with a Refeyn Two
MP
mass photometer (Refeyn LTD, Oxford, UK) at room temperature.
Glass coverslips and gaskets were cleaned with HPLC-grade water and isopropanol and dried under filtered gas before use. NiV L-P
was diluted to 200 nM in SEC buffer. Eighteen microliters of buffer were used to find the camera focus prior to loading 2 ml of the
sample onto the gasket. Acquisition camera image size was set to medium. Data were collected as a 1 min movie and then processed
using ratiometric imaging. To correlate ratiometric contrast to mass, the Refeyn Two
MP
instrument was calibrated using molecular
standards of monomeric bovine serum albumin (BSA) (66 kDa), dimeric BSA (132 kDa), and thyroglobulin (MW 660 kDa) with a
molecular weight error less than 5%. The experiment was performed twice. Data were analyzed using Discover
MP
version 2.3 soft-
ware (Refeyn LTD, Oxford, UK).
In vitro RNA synthesis assay
L protein was quantified by SDS-PAGE and densitometric comparison with a BSA standard curve on the same gel. L-P complexes
(8–90 nM final concentration) were incubated with 2 mM NiV le 1–17 template with the sequence 3’-UGGUUUGUUCCCUUUUA-5’ in a
buffer containing 50 mM Tris-HCl, pH 7.5, 8 mM MgCl
2
, 5 mM DTT and 10% glycerol (v/v) for 10 min at 30C. Reactions were initiated
by the addition of ATP and CTP to a final concentration of 1 mM each and 10 mCi of [a-
32
P]-GTP tracer (Revvity, 3000 Ci/mmol and
10 mCi/ml, final concentration 100–150 nM) in a total reaction volume of 50 ml and allowed to proceed for 1 h at 30C. The polymerase
complex was inactivated by heating to 95C for 5 min. Reactions were subsequently treated with 10 U of calf intestinal alkaline phos-
phatase (NEB, 10,000 U/ml) for 1 h at 37C. 200 ml of 0.25% SDS in nuclease-free water were added to each and RNA products were
isolated by extraction with 250 ml of acid-phenol:chloroform (Invitrogen). After the addition of 170 mM NaCl and 15 mg glycogen
(Invitrogen), RNA products were precipitated with ethanol overnight at -20C. Pellets were washed once with ice-cold 70% (v/v)
ethanol, briefly air-dried and resuspended in 8 ml water. An equal volume of 2 x STOP buffer (20 mM EDTA, 0.01% each bromophenol
blue and xylene cyanol in deionized formamide) was added before heat denaturing samples for 5 min at 95C. Samples were flash
cooled on ice and loaded onto a 20% acrylamide sequencing gel containing 7 M urea in Tris-borate-EDTA buffer. Gels were vacuum
dried at 80C onto Whatman 3MM paper. RNA products were visualized by phosphor imaging (Typhoon IP, GE Healthcare Life
Technologies).
Cryo-EM sample preparation and data collection
An aliquot of 4 ml of NiV L-P complex at 0.7 mg/ml was applied to a freshly glow-discharged Quantifoil Cu 1.2/1.3 400 mesh grid (Elec-
tron Microscopy Sciences Cat# Q4100CR1.3). The sample was blotted for 3 s after incubation for 15 s at 4C with a relative humidity
of 100%, then plunge-frozen in liquid ethane using Vitrobot Mark IV (Thermo Fisher Scientific, USA), and stored in liquid nitrogen.
Cryo-EM data were collected on a Titan Krios microscope (Thermo Fisher Scientific, USA) (300 kV) equipped with a K3 Summit direct
electron detector (Gatan, USA) at the Harvard Cryo-Electron Microscopy Center for Structural Biology. Movie stacks were automat-
ically recorded using SerialEM 4.1-beta
55
in counting mode at a nominal magnification of 105,000 x, corresponding to a physical pixel
size of 0.83 A
˚. The defocus was set to from -1.0 to -2.5 mm. A total exposure dose of 53 e
-
/A
˚
2
was fractionated into 50 frames for each
movie stack. We obtained one cryo-EM dataset of NiV L-P complex including a total number of 7803 movie stacks.
Cryo-EM image processing
All images were processed using Relion 3.1,
58
movie frames were gain-normalized and motion-corrected using MotionCor2 version
1.6.4
56
, and contrast transfer function (CTF) correction was performed using CTFFind4 version 4.1.14,
57
as implemented in Relion
3.1. Particle-picking was performed in crYOLO version 1.9.9
60
and 1,871,392 particles from 7,803 micrographs were picked. Picked
particles were imported to Relion, extracted, binned to a pixel size of 1.66 A
˚and subjected to 2D classification. Good classes of par-
ticles from the 2D classification were selected and imported to cryoSPARC version 4.5.1
59
to generate an initial model. Three rounds
of 3D classification with C1 symmetry were imposed on 1,842,231 particles, 477,936 particles of which were selected and subjected
ll
OPEN ACCESS
Cell 188, 1–16.e1–e8, February 6, 2025 e5
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
to auto-refinement to generate a 2.7 A
˚map. CTF refinement and Bayesian polishing were performed to improve the resolution.
Consequently, a final map was generated to 2.3 A
˚. To improve the local resolution of P ⍺-helical cap structure, focused 3D auto-
refinement was performed, allowing us to obtain a masked a 3.0 A
˚map. We also used DeepEMhancer version 20220530_cu10
61
for cryo-EM volume post-processing.
Model building and refinement
The structures of the five domains (RdRp, CAP, CD, MTase domain and CTD domain) of L and of monomeric P were predicted using
Phyre2 version 2.0.
64
These predicted structures and the crystal structure of the NiV P OD (PDB code: 4N5B,
15
residues 476–576)
were used to build the atomic model of NiV L-P complex. The RdRp and CAP domains of L and tetrameric P OD plus a single P XD
were docked and rigid-body fitted well into the cryo-EM map using UCSF Chimera. Extra residues of P were built manually in Coot
version 0.9.
62
Based on interpretable density, we could model L residues 5–1463, with the exception of residues 601–709, 1148–
1153, 1266–1289 (priming loop), and 1342–1362 (intrusion loop). We could model P1 residues 479–584, P2 residues 479–708 (except
for residues 581–591 and 612–630), P3 residues 477–579, and P4 residues 479–596. We performed manual model building to
improve local fit using Coot
62
and real space refinement using Phenix.
54
MolProbity
63
was used to validate the model. Statistics
are provided in Table S1.
3D variability analysis
3D variability analysis (3DVA) was performed using cryoSPARC.
43,59
Particle stacks and reference map from Relion final 3D auto-
refine were imported. 3DVA was performed using a mask generated from Relion 3.1. Twenty components were generated, of which
major components showed flexibility in the P segment. The frames from these components were visualized and recorded in UCSF
Chimera
52
as a volume series (see Video S1).
Negative stain electron microscopy
For negative stain electron microscopy experiments, 4 ml of purified protein samples were applied to glow-discharged continuous
carbon films supported by 400-mesh copper grids (Electron Microscopy Sciences Cat#: FCF400-Cu-50) stained with 1.5% (w/v)
uranyl formate. Stained grids were imaged on a Philips CM10 transmission electron microscope (100 kV) equipped with a Gatan
UltraScan 894 (2k x 2k) CCD camera. Images were collected at a magnification of 52,000x (2.06 A
˚/pixel) and processed using cry-
oSPARC.
59
Particles were selected using Blob picker and particles were subjected to several rounds of reference-free alignment and
2D classification.
In silico modeling, molecular dynamics simulations and molecular docking
The cryo-EM structure of the NiV L-P complex was prepared before modelling and simulations. The module of Protein Preparation in
Schro
¨dinger Maestro
65
was applied to cap termini, repair residues, optimize H-bond assignments, and run restrained minimizations
following default settings.
The Schro
¨dinger Desmond MD engine 2024–3
66
was used for simulations.
67
An orthorhombic water box was applied to bury pre-
pared protein systems with a minimum distance of 10 A
˚to the edges from the protein. Water molecules were described using the SPC
model. Na
+
ions were placed to neutralize the total net charge. All simulations were performed following the OPLS4 force field.
68
The
ensemble class of NPT was selected with the simulation temperature set to 300K (Nose-Hoover chain) and the pressure set to
1.01325 bar (Martyna-Tobias-Klein). A set of default minimization steps pre-defined in the Desmond protocol was adopted to relax
the MD system. The simulation time was set to 100 ns for the protein system with three duplicate MD runs. One frame was recorded
per 200 ps during the sampling phase. Post-simulation analysis of the RMSF was performed using a Schro
¨dinger simulation inter-
action diagram. RMSF values from the Caof each residue were used for plotting.
The molecular docking was performed using the Glide module in Schro
¨dinger. The binding pocket on NiV L was defined by select-
ing residues E922, S925, S928, Y1001, I1068, and H1165/Y1165 along with surrounding residues. Similarly, the binding pocket on
PIV3 L (PDB: 8KDC)
27
was defined by selecting residues E863, S866, S869, Y942, I1009, T1010, and Y1106 along with surrounding
residues. The 3D conformation of the compound GHP-88309 was prepared using LigPrep. Glide SP, in standard precision mode,
was used without constraints to generate binding poses. Up to 20 docked poses were allowed to be generated for GHP-88309 inside
the defined pockets for each protein. Among these poses, the top 10 poses underwent the post-docking minimization. The prior
study by Cox et al.
44
was referred to guide the visual inspection of minimized poses. Representative conformations that could repro-
duce the predicted binding pose
44
were selected and investigated to be reported in this manuscript.
AlphaFold 3 modeling
Predicted structures for RNA-bound NiV L (Bangladesh strain, GenBank: AAY43917.1) or PIV3 L (GenBank: WCF97250.1) were
generated using AF3.
26
Modeling was performed using the leader RNA 1–16 (underlined) plus six adenines as the template RNA
(50-UUUUCCCUUGUUUGGUAAAAAA-30for NiV and 50-UCUUCUCUUGUUUGGUAAAAAA-30for PIV3) and the complementary
sequence of leader RNA 1–9 (underlined) plus three adenines as the nascent RNA (50-AAAACCAAACAA-30) along with a GTP mole-
cule, two magnesium ions, and two zinc ions. Note that non-specific adenines were added to the template and nascent RNAs to
ensure that their 30and 50ends, respectively, remained single stranded in the model.
ll
OPEN ACCESS
e6 Cell 188, 1–16.e1–e8, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
Design and cloning of the NiV minigenome system
The minigenome system was designed based on sequences from the NiV Bangladesh strain (GenBank: AY988601.1). The cis-acting
elements inserted into the multi-cycle replication minigenome were determined based on a previously described NiV minigenome
template.
10
The multi-cycle minigenome contained in 3’ to 5’ order, the first 112 nt of the NiV genome sequence, including the 55
nt leader region, Ngene start signal, and 46 nt of non-translated sequence from the 3’ end of the Ngene, which includes promoter
element 2 (CNNNNN)
3
, 519 nt derived from the bacterial chloramphenicol acetyltransferase (CAT) gene, 105 nt of sequence derived
from the N-P gene junction region (nucleotides 2247–2351 of the NiV Bangladesh complete genome) that includes the Ngene end
signal, trinucleotide intergenic, and Pgene start signal, followed by 939 nt Renilla luciferase reporter gene sequence, and then the 5’
100 nt of the NiV genome, including 56 nt of non-translated sequence from the 5’ end of the L gene, which contains the complement of
promoter element 2, the Lgene end signal and the 30 nt 5’ trailer region. Together with inserted restriction sites, the total minigenome
length is 1794 nt. The minigenome cassette was flanked by a hepatitis delta virus ribozyme sequence adjacent to the leader region
and a T7 promoter sequence adjacent to the trailer region (Figure S6A). This minigenome was used as a template to generate a single-
step minigenome limited to the antigenome synthesis step of replication. This minigenome had a deletion of promoter element 2
within the L nontranslated region and substitutions at position 1–4 relative to the 5’ end of the trailer region, which inactivated the
trailer promoter at the 3’ end of the antigenome and created an optimal T7 promoter sequence (Figure S6A). The multi-cycle mini-
genome followed the ‘‘rule of six’’ except for two additional G residues contributed by the T7 promoter; the single-step minigenome
followed the rule of six including the additional G residues contributed by the T7 promoter. Codon optimized versions of NiV
Bangladesh strain N, P, and L open reading frames (Synbio) were inserted into a modified version of plasmid pTM1 (a kind gift
from Dr. Bernard Moss),
69
which contains a T7 promoter and internal ribosome entry site. Each open reading frame was inserted
such that the initiating codon was inserted into the plasmid NcoI site, and the 30end of the open reading frame was flanked with
a 17 nt poly A sequence.
27
In addition, the open reading frame of L contained a C-terminal strep tag. All plasmids were sequenced
in their entirety (Plasmidsaurus) and their integrity and purity was confirmed by agarose gel electrophoresis prior to each transfection.
Reconstitution of minigenome replication and transcription
BSR-T7 cells (a kind gift from Dr. Klaus Conzelmann)
70
in six-well dishes were transfected with (per well) 0.3 mg of the relevant
minigenome plasmid, 0.4 mg of N, 0.2 mg of P, 0.1 mg of L and 0.04 mg of firefly expression minigenome using Lipofectamine 3000
(Invitrogen Cat# L3000015). Each transfection reaction was set up in duplicate. Followed by incubation at 37C for 22–24 h, the trans-
fection mixture was replaced with OptiMem containing 2% (v/v) fetal bovine serum. Cells were harvested 44–48 h after transfection.
Harvesting of transfected cells for Western blot and luciferase assays
For Western blot and luciferase analyses, cells from one transfected well were lysed in 500 ml passive lysis buffer (Promega). A 100 ml
aliquot of the lysate was passed through a QIAshredder column (Qiagen Cat# 79656) and a 14 ml aliquot was subjected to electro-
phoresis on an 8% SDS-polyacrylamide gel alongside a PageRuler Plus Prestained Protein Ladder (Thermo Fisher Scientific Cat#
26619). Proteins were transferred to nitrocellulose by Western blotting and the blots were probed with beta-tubulin rabbit polyclonal
antibody (LI-COR) at 1:1000 and a Strep-tag classic mouse monoclonal antibody (Bio-Rad Cat# MCA2489) at 1:500 dilution, followed
by incubation with IRDye 680RD-conjugated donkey anti-rabbit (LI-COR Cat# 926-68073) and IRDye 800CW-conjugated goat anti-
mouse (LI-COR Cat# 926-32210) antibodies, each at 1:20,000 dilution. Blots were scanned using an OdysseyCLx imager (LICOR).
For quantification a box of equivalent size as used for other bands of the same protein species was drawn at the corresponding po-
sition on the -L sample lane and used to set the background to 0. The resulting values were then normalized to the corresponding
value from the L WT reaction, which was set to 1.
For luciferase assays, aliquots of cell lysates were diluted 1:30 and assessed for firefly and Renilla luciferase activity using a Dual-
Luciferase Reporter Assay System (Promega Cat# E1910) according to the manufacturer’s instructions. Quantification of the relative
luciferase activity for each sample was performed by normalizing the Renilla luciferase value to the firefly luciferase value. The -L
value was subtracted to account for background. The resulting values were then normalized to the corresponding value from the
L WT reaction, which was set to 1.
Northern blot and hybridization
For analysis of total intracellular RNA cells from a transfected well were pelleted and RNA was extracted using a Monarch total RNA
miniprep kit (NEB Cat# T2010S). In the case of micrococcal treatment, each transfection reaction was set up in duplicate wells.
Following harvest, the cell pellet from one well of the duplicate was resuspended in 100 ml of nuclease lysis buffer 10 mM NaCl,
10 mM Tris, pH 7,5, 1.5 mM MgCl
2
, 1% Triton X-100, 0.5% sodium deoxycholate, 10 mM CaCl
2
,1ml of aprotinin (Thermo Fisher
Scientific Cat# J11388MA) and then lysis buffer from the Monarch total RNA miniprep kit (NEB Cat# T2010S) was added directly.
The cell pellet from the other well was resuspended in 100 ml of nuclease lysis buffer and 10 ml of micrococcal nuclease S7 (Roche
Cat# 10107921001) was added. The reaction containing micrococcal nuclease was incubated at 30C for 1 h with occasional
shaking, following which the Monarch total RNA miniprep kit (NEB) lysis buffer was added. RNA from the untreated and treated cells
was isolated using the Monarch total RNA miniprep kit (NEB) according to the manufacturer’s instructions. Isolated RNA was sub-
jected to denaturing gel electrophoresis alongside a molecular weight ladder (DynaMarker prestain marker for RNA high; Diagnocine
Cat# DM260S) in a 1.5% (w/v) agarose gel containing 0.44 M formaldehyde in MOPS buffer. The RNA was transferred to a nylon
ll
OPEN ACCESS
Cell 188, 1–16.e1–e8, February 6, 2025 e7
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
membrane (Cytiva), which was stained with methylene blue to confirm that equal amounts of RNA were loaded in each lane and to
allow visualization of the molecular weight markers. Negative- and positive-sense
32
P-labled CAT-specific riboprobes corresponding
to minigenome (or complementary) sequence were synthesized by T7 RNA polymerase and purified by phenol-chloroform extrac-
tion. The appropriate riboprobe was hybridized to the membrane in 6X SSC, 2X Denhardt’s solution, 0.1% SDS, and 100 mg/ml of
sheared salmon sperm DNA for a minimum of 18 h at 65C. The membranes were washed at 65C in 2X SSC-0.1% SDS for 2 h
and in 0.1X SSC-0.1% SDS for 15 min. Phosphorimager analysis was performed using an Azure Sapphire FL Biomolecular Imager
(IS4000) and signals were quantified using Image Studio (LI-COR). The CAT mRNA would be expected to be 636 nt (excluding the
poly A tail which is heterogeneous in length), the antigenome and genome RNAs would be expected to be 1,776 nt in the case of the
single-step minigenome, and the replicated antigenome and genome RNAs would be expected to be 1,794 nt in the case of the multi-
cycle minigenome. To quantify the RNA levels, a box of equivalent size as used for other bands of the same RNA species was drawn
at the corresponding position on the -L sample lane and used to set the background to 0. The resulting values were then normalized to
the corresponding value from the L WT reaction, which was set to 1.
Primer extension analysis
Primer extension reactions were carried out using 4–8 mg of total intracellular RNA isolated from transfected cells. RNA was extracted
using a Monarch Total RNA Miniprep Kit (NEB). The RNA initiated from position 1U at the 3’ end of the le region, and RNA initiated at
the first gs signal was detected with a
32
P-end-labeled primer (5’-GATATATTTCTTGAGGATCC-3’), which corresponds to nucleo-
tides 90-109 from the 3’ end of the minigenome le region. Thus, the cDNA products derived from RNA initiated at the 3’ end of le
or the gs signal would be expected to be 109 nt or 54 nt, respectively. Each reaction was prepared with either Sensiscript reverse
transcriptase (QIAGEN Cat# 205211) or Moloney murine leukemia virus reverse transcriptase (Promega Cat# M1701), following
the manufacturer’s instructions. Using Sensiscript reverse transcriptase, the RNA samples were reverse transcribed at 50C for
1.5 hours. With M-MLV, the RNA samples were incubated for at 42C for 1.5 hours. After incubation with reverse transcriptase,
2X STOP buffer (deionized formamide, 20 mM EDTA, 0.1% (w/v) bromophenol blue, and xylene cyanol) was added. A single-
stranded DNA ladder (Simplex Cat# ss20) was end-labeled using T4 Polynucleotide Kinase (NEB Cat# M0201), following the man-
ufacturer’s instructions. The primer extension products were then separated by electrophoresis on 8% polyacrylamide gels contain-
ing 7 M urea in Tris-borate-EDTA buffer. After electrophoresis, the gels were dried onto 3MM paper using a vacuum drier at 80C.
Phosphorimager analysis was performed using an Azure Sapphire FL Biomolecular Imager (IS4000), and signals were quantified us-
ing Image Studio (LI-COR). A box of equivalent size to those used for other bands of the same RNA species was drawn at the cor-
responding position on the negative control lane and used to set the background to 0. The resulting values were then normalized to
the corresponding value from the L WT reaction, which was set to 1.
Figure preparation
UCSF Chimera version 1.15, ChimeraX version 1.5, and PyMOL version 2.5.5 (Schro
¨dinger LLC) were used for structure visualization
and figure generation. Prism 10 version 10.3.1 (GraphPad Prism) was used to prepare the bar charts shown in Figures 6,7, and S6.
QUANTIFICATION AND STATISTICAL ANALYSIS
The data analysis function in Microsoft Excel (version 16.62) was used to conduct the unpaired, two-tailed Student’s t-test to
compare mean RMSF values from two groups. Statistical significance was defined as p< 0.05. Data from the Image Studio Lite
(version 5.2) and luciferase readings were inputted into Prism 10 (GraphPad Prism version 10.3.1) and means and standard deviations
were calculated by Prism 10. The details of tests, including the number of replicates, are included in figure legends.
ll
OPEN ACCESS
e8 Cell 188, 1–16.e1–e8, February 6, 2025
Please cite this article in press as: Hu et al., Structural and functional analysis of the Nipah virus polymerase complex, Cell (2025), https://
doi.org/10.1016/j.cell.2024.12.021
Article
Supplemental figures
(legend on next page)
ll
OPEN ACCESS
Article
Figure S1. Purification and cryo-EM reconstruction of NiV L-P, activity assay, and structural alignment with other nsNSV L proteins, related
to Figure 1
(A) Size exclusion chromatography profile of NiV L-P complex. The complex has a retention volume of 11 mL. A sample obtained from the 11 mL peak fraction
was examined using SDS-PAGE, with a Coomassie-stained gel shown (inset). NiV L has a molecular weight of 260 kDa. NiV P has a molecular weight of 80 kDa
but has an apparent size that is larger than its molecular weight (100 kDa).
(B) RNA synthesis activity assay with radioactive products migrated on a denat uring polyacrylamide gel, as described for Figures 1C and 1D. Different amounts of
polymerase complexes were used in this experiment as indicated (values indicate nM amounts of L). This analysis clearly shows products that are longer than the
input template. Based on the observed migration patterns of the longer products (with incremental single nucleotide additions) and prior findings that nsNSV
polymerases are prone to reiterative stuttering,
11,71
we conclude that these additional bands are due to polymerase stuttering on U tracts within the promoter,
although further experiments are required to show this conclusively.
(C) Workflow used for cryo-EM data processing of the NiV L-P complex. Low -resolution features in 3D volumes consistent with the presence of flexible portions of
the L C-terminal globular domains (CD, MTase, and CTD) are indicated with a dashed circled.
(D) Fourier shell correlation (FSC) curves. The threshold used to estimate the resolution is 0.143.
(E) Local resolution estimates of the NiV L-P complex obtained using ResMap.
72
(F) Structural alignment of NiV L (residues 5–1463) with the L proteins of the indicated nsNSVs calculated over C-⍺positions. The alignment was performed using
PyMOL. RMSD, root-mean-square deviation.
ll
OPEN ACCESS Article
(legend on next page)
ll
OPEN ACCESS
Article
Figure S2. Partial sequence alignments of nsNSV L proteins, active site predictions, and negative stain images of NiV L, related to Figure 2
(A) Partial sequence alignment of nsNSV L proteins showing motifs A, C, and F. Motif C GDNE/Q is the catalytic site in the RdRp domain and is conserved among
all the L proteins aligned. The fourth residue is glutamine in most sequenced viruses but is glutamate in a few viral sequences, including NiV L. Motif A D722, motif
C D832, and motif F R551 are indicated by black triangles. See Table S2 for virus name abbreviations and GenBank accession numbers.
(B) PIV3 L polymerase active site in the 2.7 A
˚cryo-EM structure of the PIV3 L-P complex.
27
Active site residues D773 (motif C), D663 (motif A), and R552 (motif F)
are shown as sticks. A magnesium ion (green sphere) was resolved in the cryo-EM structure interacting with D773 in the catalytic center.
(C) AF3
26
prediction of the PIV3 L active site in the presence of a duplex RNA, GTP, and two magnesium ions (green spheres). D663 (motif A) and D773 (motif C)
are predicted to coordinate metals, and R552 (motif F) is positioned to interact with the phosphate of the incoming nucleotide. Template and nascent RNA are
shown in dark and light gray, respectively. Parts of motifs A and D are shown as semi-transparent for clarity.
(D and E) pLDDT of the active sites of the AF3 model of NiV L-RNA (D) and PIV3 L-RNA (E). Residues with very high confidence (pLDDT > 90) are shown in deep
blue, confidence (90 > pLDDT > 70) in light blue, low confidence (70 > pLDDT > 50) in yellow, and very low confidence (pLDDT < 50) in orange.
(F–I) Amino acid sequence alignment of palm insert regions of NiV and HeV (F), MojV and LayV (G), NiV and CedPV (H), and NiV and GhV (I). CedPV has a
particularly long palm insert.
(J) Phylogenetic tree of the indicated parahenipaviruses (MojV and LayV) and henipaviruses (GhV, NiV, HeV, and CedPV) based on L amino acid sequences
generated using Clustal Omega.
30
(K) SDS-PAGE gel of eluted fraction from Streptactin purification of NiV L
D-P.I.
-P run under reducing conditions. Imaging was performed with a stain-free gel
system.
(L) Negative stain image of NiV L-P WT. After Streptactin purification, NiV L-P was diluted to 0.03 mg/mL for imaging by negative stain electron microscopy. The
scale bar is 50 nm.
(M) Negative stain image of NiV L
D-P.I.
-P. After Streptactin purification, NiV L
D-P.I.
-P was diluted to 0.06 mg/mL for negative stain electron microscopy. The scale
bar is 50 nm.
ll
OPEN ACCESS Article
Figure S3. CAP domain features for additional polymerase structures, related to Figure 3
(A–C) AF3
26
predicted ZFs of HeV (A), LayV (B), and MojV (C). Zinc ions are shown as gray spheres, and residues coordinating zinc ions are shown as sticks.
(D–I) Cryo-EM structures of PIV3 L-P (PDB: 8KDC)
27
(D), NDV L-P (PDB: 7YOU)
41
(E), PIV5 L-P (PDB: 6V85)
37
(F), VSV L-P (PDB: 6U1X)
39
(G), RSV L-P (6PZK)
20
(H),
and RSV L-P-RNA (PDB: 8SNX)
18
(I). The L protein RdRp domains are shown in surface representation with partially clipped surfaces, and the CAP domains are
shown as ribbon diagrams. The priming (green) and intrusion (purple) loops are shown. The disordered portions within these loops that were not resolved in the
cryo-EM structures are shown as dashed lines. HR residues are indicated. Except for NDV, whose intrusion loop HR motif arginine (R1269) was in a disordered
segment, the HR residues could be resolved. Nucleic acids are shown in orange.
(J) pLDDT scores of an AF3
26
model of the NiV L RdRp and CAP domains in the absence of RNA. The priming and intrusion loops are indicated and shown with
thicker ribbon representation.
(legend continued on next page)
ll
OPEN ACCESS
Article
(K) AF3 model of RNA-bound NiV L showing the location of the priming (green) and intrusion (purple) loops with respect to the putative nascent RNA channel, the
general location of which is indicated by the yellow oval. Intrusion loop residues N1344, L1345, and R1348, which are indicated, were subjected to mutational
analysis using minigenome assays (see Figure S6R). Priming loop residue T1276 is equivalent to RSV priming loop residue T1267, which is a part of the
GxxT motif.
(L) Cryo-EM structure of RSV L-P bound to promoter RNA (PDB: 8SNX)
18
showing the location of the priming (green) and intrusion (purple) loops with respect to
the putative nascent RNA channel, the general location of which is indicated by the yellow oval. RSV L intrusion loop residues N1335, Y1336, and R1339 are
equivalent to NiV L residues N1344, L1345, and R1348. RSV L priming loop residue T1267 is equivalent to NiV L residue T1276.
ll
OPEN ACCESS Article
Figure S4. Tetrameric P is flexible when assembled onto the L polymerase core, related to Figure 4
(A) Per-residue root-mean-square-fluctuation (RMSF) from 100 ns molecular dynamics (MD) simulations of the protein complex. Three thin lines represent
individual MD runs. A thick line represents averaged RMSF values. For the RdRp/CAP portion of the plot, regions denoted as ‘‘D’’ are stretches of residues that
had no observable or interpretable density in cryo-EM maps and therefore were not modeled in atomic coordinates, and high RMSF fluctuations in these regions
are likely due to interruption of the polypeptide chain.
(B) Mean RMSF values calculated during MD simulations for L (RdRp/CAP), P1, P2, P3, and P4 are plotted. The comparison of mean RMSF values between two
groups was performed using an unpaired, two-tailed Student’s t test. Error bars represent standard errors. ***p< 0.001.
(C) Cryo-EM structure of the NiV L-P complex provided as a reference for regions of P with increased flexibility.
ll
OPEN ACCESS
Article
Figure S5. GHP-88309 structure, previously described activity profile, and conservation of inhibitor binding site, related to Figure 5
(A) Structure of the broad-spectrum paramyxovirus antiviral GHP-88309 in 2D (top) and 3D (bottom) representations.
(B) Summary of results of antiviral activity assays for GHP-88309 as determined by Cox et al. using minigenome assays for NiV, PIV3, MeV, or recombinant SeV.
44
(C) Partial sequence alignment of the L proteins of the indicated viruses. Residues implicated in inhibitor resistance by Cox et al.
44
or near the drug in docking
experiments are indicated.
(D) pLDDT of AF3
26
model of NiV L-RNA complex.
ll
OPEN ACCESS Article
Figure S6. Validation of the single-step minigenome assay, mutant L expression, gene expression activities of L mutants measured by
luciferase activity, and summary of functional studies, related to Figures 6 and 7
(A) Sequence alignment of the 30ends of the antigenome RNAs of the multi-cycle and single-step minigenomes (Figures 6A and 6B). The alignment shows the
internal NiV promoter element (promoter element 2; blue type), the complement of the trailer region (green type), and the T7 promoter (black type). The T7 RNA
polymerase initiates opposite the C residues that are underlined to synthesize negative-sense RNA. Dashes indicate residues that were deleted in the single-step
minigenome. Substitutions introduced into the trailer promoter of the single-step minigenome are shown in lowercase type and were designed to ablate the
promoter and enhance T7 promoter activity. The multi-cycle minigenome followed the rule of six (i.e., its nucleotide length was divisible by six) except for the
(legend continued on next page)
ll
OPEN ACCESS
Article
additional G residues contributed by the T7 promoter, and the single-step minigenome followed the rule of six, including the additional G residues contributed by
the T7 promoter.
(B and C) Northern blot analysis of negative-sense (minigenome sense) RNA generated by either T7 RNA polymerase and NiV polymerase in the case of the multi-
cycle minigenome or by T7 RNA polymerase alone in the case of the single-step minigenome. The blots show RNA generated in cells transfected with plasmids
expressing either WT L or L with a substitution in the GDNE motif of the RdRp (L D832A). Cells were transfected with plasmids expressing either multi-cycle (MC)
or single-step (SS) replication minigenomes, or minigenome (MG) was omitted from the transfection (MG). (B) shows total RNA harvested from transfected cells.
(C) shows RNA from a parallel transfection in which the cell lysate was treated with micrococcal nuclease prior to RNA purification to reduce levels of
unencapsidated RNA and distinguish encapsidated minigenome template RNA.
(D and E) Quantification of the bands from replicates of the experiments shown in (B) and (C), respectively.
(F and G) RNA from the same transfections as used for (B) and (C), in which the northern blot was probed to detect positive-sense RNA (antigenome and CAT
mRNA).
(H and I) Quantification of the bands from replicates of the experiments shown in (F) and (G), respectively. (H) shows quantification of CAT mRNA, and (I) shows
quantification of nuclease-resistant antigenome RNA.
The data shown in (B), (C), (F), and (G) are representative of two independent experiments. The bars in (D), (E), (H), and (I) show the two data points and the mean
for the two independent experiments, with normalization to L WT + MC minigenome in each case.
This experiment shows that the single-step replication minigenome had a stronger T7 promoter than the multi-cycle replication minigenome (B andD). However,
the single-step replication minigenome was not amplified by WT L protein (C and E) and consequently produced relatively low levels of CAT mRNA (F and H) and
antigenome (G and I).
(J–L) Western blot analysis of minigenome-transfected cell lysates probed with a strep-tag-specific antibody to detect L (green), and tubulin (red) was detected
with a specific antibody as a loading control. The band marked with an asterisk is the appropriate molecular weight to be truncated L protein that had been
cleaved at the palm insert.
(M–O) Quantification of the full-length NiV L band from replicates of the experiments sho wn in (J)–(L), respectively. Each bar represents the mean and SD derived
from three independent experiments.
(P and Q) Renilla luciferase activity of L mutants in assays with a single-step minigenome. The bars show the mean and SD from three independent experiments,
except for the H1347A/R1348A (HR) and Q454R/C457E (P4) mutants (n=7).
(R) Summary of functional studies with NiV minigenome system. *Activity levels determined by primer extension analysis of RNA initiated at the 30end of the
le region (replication) or gs signal (transcription).
ll
OPEN ACCESS Article