ArticlePDF Available

Abstract and Figures

Over 950,000 whole genome sequences of SARS-CoV-2 have been determined for viruses isolated from around the world. These sequences have been critical for understanding the spread and evolution of SARS-CoV-2. Using global phylogenomics, we show that mutations frequently occur in the C-terminal end of ORF7a. We have isolated one of these mutant viruses from a patient sample and used viral challenge experiments to link this isolate (ORF7aΔ115) to a growth defect. ORF7a has been implicated in immune modulation, and we show that the C-terminal truncation negates anti-immune activities of the protein, which results in elevated type I interferon response to the viral infection. Collectively, this work indicates that ORF7a mutations occur frequently and that these changes affect viral mechanisms responsible for suppressing the immune response.
ORF7a truncation results in loss-of-function. A) Amino-acid (aa) sequence alignment of SARS-CoV-2 ORF7aWT, ORF7aΔ115 and SARS-CoV-1 ORF7a. Gaps show non-matching positions, red shows 17 aa sequence resulting from a frameshift in the ORF7a mutant. Beta strands (arrows) and alpha helices (coil) are shown above the alignment. B) Diagram of SARS-CoV-2 ORF7a Ig-like fold. Disulfide bonds that stabilize the β-sandwich structure are shown with red lines. The portion of the protein eliminated by the deletion is shown in gray. C) C-terminally Flag-tagged ORF7aWT and ORF7aΔ115 were cloned and overexpressed in HEK 293T-hACE2. Protein expression was confirmed with Western Blot using anti-Flag antibody. β-actin (ACTB) was used as a loading control. (D) Flag-tagged ORF7aWT or (E) ORF7aΔ115 expressed in HEK 293T-hACE2 cells. Immunostaining preformed using an anti-Flag antibody (green). Cell nuclei were stained with Hoechst 33342 (blue). White scale bar is 10 μm. F) HEK 293T cells with integrated ISRE-luciferase reporter were transfected with pLV-mCherry (“-“), ORF7aWT or ORF7aΔ115 plasmids. Transfected cells were treated with 5 ng/mL of human recombinant IFN-α2b for 24 h and induction of type I IFN signaling was measured with luciferase assay. Fold induction vs non-treated control was determined and normalized to mCherry control. Means (n = 6) were compared with one-way ANOVA (p-value = 0.00162). Pairwise comparisons were performed using post hoc Tuckey’s test. Data is shown as mean ± sd. ∗∗, p < 0.01; ns – not significant, p > 0.05. G) ORF7a and ORF7b sgRNAs were identified by RT-PCR. Lower MW band corresponding to ORF7b sgRNA is indicated with asterisk (*). Diagram on the right shows primer (arrows) positions. Specificity of PCR products was confirmed with sanger sequencing. GAPDH was used as a control for cDNA synthesis. H) Lysates from SARS-CoV-2 infected cells were probed with antibodies raised against ORF7b protein. ACTB was used as a loading control.
… 
The ORF7aΔ115 mutation results in replication and immune suppression defects. VeroE6 (A) or HEK 293T-hACE2 (B) were infected with ORF7aWT or ORF7aΔ115 SARS-CoV-2 strain at MOI = 0.05. Viral RNA was measured in supernatant using RT-qPCR at different timepoints after infection. Measured RNA levels were normalized (using ΔCt method) to RNA levels at 0.5 hpi. (C) Total RNA was extracted from infected HEK 293T-hACE2 cells and viral RNA was measured. ΔΔCt method was used to normalize viral RNA levels to host RNA (ACTB) and 0.5 hpi timepoint. Data in A-C are presented as mean ± SD of 3 biological replicates. Significance levels: ∗ p<0.05, ∗∗ p<0.01, *** p<0.001 or ns (no significant difference, p>0.05). D) Volcano plot showing IFN-I response in 293T-hACE2 cells infected with ORF7aWT SARS-CoV-2 strain at MOI = 0.05. Expression of IFN-I response genes was studied 24 hpi using RT-qPCR array targeting 91 human transcripts (88 targets and 3 references). Experiment was performed in 3 biological replicates. Dashed lines show regulation (≥2-fold) and statistical significance (p-value < 0.05) thresholds. Each dot represents mean (n = 3) normalized expression of a single gene relative to non-infected host. Genes that passed the threshold are labeled. E) Volcano plot showing IFN-I response in ORF7aΔ115 vs ORF7aWT infection. F) Data shown in D) and Figure S3A was plotted as a heatmap. Genes were classified into three groups. Group #1 genes are oppositely regulated between the two viral strains. Group #2 genes are upregulated by both ORF7aWT and ORF7aΔ115. Group #3 genes are downregulated in both. Genes that have statistically significant difference in expression between two viral strains are marked with asterisk. Gray shows genes with no detectable expression over 40 cycles of RT-qPCR.
… 
Content may be subject to copyright.
Report
SARS-CoV-2 genomic surveillance identifies
naturally occurring truncation of ORF7a that limits
immune suppression
Graphical abstract
Highlights
dORF7a mutations are found in SARS-CoV-2 genomes
isolated from around the globe
dThe ORF7aD115 isolate displays a replication defect
dAn ORF7a mutation limits viral suppression of the interferon
response
Authors
Artem Nemudryi, Anna Nemudraia,
Tanner Wiegand, ..., Diane Bimczok,
Mark A. Jutila, Blake Wiedenheft
Correspondence
bwiedenheft@gmail.com
In brief
Nemudryi et al. use global SARS-CoV-2
phylogenomics to identify mutations that
frequently occur in the C-terminal end of
ORF7a accessory protein. They isolate
one of these mutant viruses from a patient
sample and demonstrate that ORF7a
truncation affects viral mechanisms
responsible for suppressing host immune
response.
Nemudryi et al., 2021, Cell Reports 35, 109197
June 1, 2021 ª2021 The Author(s).
https://doi.org/10.1016/j.celrep.2021.109197 ll
Report
SARS-CoV-2 genomic surveillance identifies
naturally occurring truncation of ORF7a
that limits immune suppression
Artem Nemudryi,
1,3,5
Anna Nemudraia,
1,5
Tanner Wiegand,
1,6
Joseph Nichols,
1,6
Deann T. Snyder,
1
Jodi F. Hedges,
1
Calvin Cicha,
1
Helen Lee,
1
Karl K. Vanderwood,
2
Diane Bimczok,
1
Mark A. Jutila,
1
and Blake Wiedenheft
1,4,7,
*
1
Department of Microbiology and Immunology, Montana State University, Bozeman, MT 59717, USA
2
Gallatin City-County Health Department, Bozeman, MT 59715, USA
3
Twitter: @artemnemudryi
4
Twitter: @WiedenheftLab
5
These authors contributed equally
6
These authors contributed equally
7
Lead contact
*Correspondence: bwiedenheft@gmail.com
https://doi.org/10.1016/j.celrep.2021.109197
SUMMARY
Over 950,000 whole-genome sequences of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2)
have been determined for viruses isolated from around the world. These sequences are critical for under-
standing the spread and evolution of SARS-CoV-2. Using global phylogenomics, we show that mutations
frequently occur in the C-terminal end of ORF7a. We isolate one of these mutant viruses from a patient sam-
ple and use viral challenge experiments to link this isolate (ORF7a
D115
) to a growth defect. ORF7a is impli-
cated in immune modulation, and we show that the C-terminal truncation negates anti-immune activities
of the protein, which results in elevated type I interferon response to the viral infection. Collectively, this
work indicates that ORF7a mutations occur frequently, and that these changes affect viral mechanisms
responsible for suppressing the immune response.
INTRODUCTION
The spillover of severe acute respiratory syndrome coronavirus 2
(SARS-CoV-2) into the human population has resulted in a global
pandemic of coronavirus disease 2019 (COVID-19) with more
than 2.8 million deaths worldwide (https://coronavirus.jhu.edu/)
(Andersen et al., 2020). Although much of the research on this vi-
rus is focused on the Spike protein, recent reports demonstrate
that accessory proteins of SARS-CoV-2 might be involved in
COVID-19 pathogenesis by modulating antiviral host responses
(Young et al., 2020;Zhang et al., 2020). SARS-CoV-2 uses mul-
tiple strategies to evade host immunity. Accessory proteins
ORF3b, ORF6, and ORF7a antagonize various steps of type I
interferon (IFN-I) production and signaling, while the proposed
function of ORF8 is downregulating antigen presentation (Konno
et al., 2020;Lei et al., 2020;Miorin et al., 2020;Tan et al., 2020;
Xia et al., 2020). To occlude signal transmission from IFN recep-
tors, ORF7a subverts phosphorylation of STAT2, suppressing
transcriptional activation of antiviral IFN-stimulated genes
(ISGs) that can otherwise restrict viral replication (Martin-Sancho
et al., 2020;Xia et al., 2020). Although the proposed function for
ORF7a is intracellular, antibodies against ORF7a are elevated in
the serum of COVID-19 patients (Hachim et al., 2020). Work that
is currently under review indicates that recombinant ORF7a pro-
tein interacts with CD14
+
monocytes and triggers expression of
pro-inflammatory cytokines, including interleukin-6 (IL-6) and tu-
mor necrosis factor alpha (TNF-a)(Zhou et al., 2021). However, it
is unclear if these interactions happen in COVID-19 patients.
Together, these data suggest that ORF7a plays a dual role in
SARS-CoV-2 infection by modulating both the IFN and inflam-
matory responses.
Here, we show that C-terminal mutations in ORF7a occur
frequently in samples isolated from patients around the globe.
These mutations are not derived from a single lineage, and
they do not persist over time. Using samples collected from in-
fected patients, we have isolated a virus containing a deletion
mutation in ORF7a that truncates the C-terminal half of the pro-
tein. In vitro viral challenge experiments demonstrate that this
mutation results in a replication defect and obviates viral sup-
pression of the immune response. Collectively, these data sug-
gest that ORF7a truncations are defective in suppressing the
host immune response, which may explain why these mutations
quickly disappear in the immunocompetent population.
RESULTS
SARS-CoV-2 genomic surveillance identifies
truncations of ORF7a
Genome sequencing has been used to track the rise and spread
of new SARS-CoV-2 lineages over the course of the pandemic
Cell Reports 35, 109197, June 1, 2021 ª2021 The Author(s). 1
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
ll
OPEN ACCESS
(https://www.gisaid.org/)(Bedford et al., 2020;Korber et al.,
2020). As part of this effort, we sequenced SARS-CoV-2 ge-
nomes isolated from patients in Bozeman, Montana (Figure 1A;
Table S1). In total, we determined 55 whole-genome sequences
of SARS-CoV-2 viruses isolated from patients between April and
July 2020 using an amplicon-based Nanopore protocol from the
ARTIC network (https://artic.network/)(Quick et al., 2017;Tyson
et al., 2020). To place the local outbreak in the context of the
global SARS-CoV-2 evolutionary trends, we aligned 55 whole-
genome sequences isolated from patients in Bozeman to
4,235 genomes subsampled (10 genomes per month per coun-
try) from 181,003 SARS-CoV-2 genomes downloaded from the
GISAID. The subsampling protocol removes redundancy and
bias introduced by uneven distribution of the global SARS-
CoV-2 sequencing effort. The resulting alignment was used to
build a phylogenetic tree (Figure 1B). Of the 55 SARS-CoV-2 ge-
nomes determined from patients in Bozeman, only 1 was derived
from the WA1 lineage (lineage A in Figure 1B) that was intro-
duced to Washington state from Wuhan, China, in January
2020 (Bedford et al., 2020;Holshue et al., 2020). The remaining
54 genomes associate with clade B.1 and its offshoots, which
are characterized by the D614G mutation in the spike and have
prevailed globally since the spring of 2020 (Rambaut et al.,
2020a)(Figure 1B). The genetic variability in SARS-CoV-2 circu-
lating in Bozeman (April to July 2020) represents 14 independent
viral lineages, reflecting multiple introductions of the virus to the
community from many different sources (Figure S1A).
During the annotation of these genomes, we noticed a reoc-
curring (7 out of 55 genomes) 115-nt deletion in the gene-encod-
ing accessory protein ORF7a (27,549–27,644 nt). The ORF7a
D115
Figure 1. SARS-CoV-2 genomic surveillance identifies global reoccurrence of ORF7a truncations
(A) Symptom onset (purple) and PCR-based SARS-CoV-2 test results (coral) for patients in Bozeman, Montana, are shown with vertical bars. Seven-day moving
averages, shown with lines, were used to indicate epidemiological trends.
(B) Phylogenetic analysis of SARS-CoV-2 genomes sampled in Boz eman and globally. The tree was constructed from an alignment of 55 Bozeman samples and
4,871 genomes subsampled from GISAID. Subsampling was performed using Augur utility (https://nextstrain.org) by selecting 10 genomes per country per
month since the start of the pandemic. Outer ring shows SARS-CoV-2 lineages assigned to genome sequences (Rambaut et al., 2020a). Major lineages include A
(pink) that is associated with initial outbreak in China and B (blue) that emerged later in Europe. Mino r lineages (i.e., C-N) are offshoots of lineage B. Red branches
identify truncated ORF7a variants (n = 205) detected in the global data and merged into the alignment. The red dot highlights 7 of the 55 ORF7a variants that were
isolated in Bozeman between April and July (2020). White dots highlight 48 viral genomes isolated in Bozeman that have wild-type ORF7a sequences.
(C) Distribution of different mutations that occur along the ORF7a coding sequence.
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
2Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
mutation was found in patient swabs collected over a period of
1.5 months (Table S1). RT-PCR and Sanger sequencing veri-
fied that all these seven ORF7a
D115
variants are bona fide muta-
tions and not a sequencing artifact (Figures S1B and S1C). In
addition to the ORF7a
D115
deletion, these genomes also share
10 single-nucleotide variants (SNVs) compared with the Wu-
han-Hu-1 reference genome. Seven of the 10 SNVs are found
frequently worldwide and are a signature of the B1.1 lineage of
SARS-CoV-2 (Rambaut et al., 2020a). The remaining three muta-
tions are rare, do not co-occur in any other genomes on GISAID,
and lead to amino acid changes in ORF3a (Q38P, L95F) and N
(R195I) proteins. Interestingly, one of the genomes in the
ORF7a
D115
cluster has two additional SNVs that likely were ac-
quired during circulation in the community. This virus was
sampled from a patient 41 days after sampling the first
ORF7a
D115
variant, which agrees well with estimated SARS-
CoV-2 mutation rates (2 nt/month) (van Dorp et al., 2020).
To determine if the ORF7a
D115
genotype is unique to Boze-
man, we downloaded an alignment of 180,971 SARS-CoV-2 ge-
nomes from GISAID, extracted ORF7a sequences, and deter-
mined their mutational profiles. In total, we identified 845
unique ORF7a gene variants that are different from the Wuhan-
Hu-1 reference sequence. Next, we looked for mutations with
major effect on the ORF7a amino acid sequence. We identified
189 unique ORF7a variants (825 total sequences) with frame-
shifts, deletions, or missense mutations causing premature
stop codons (Figure 1C). To understand the evolutionary history
of these mutations, we integrated these genome sequences into
our phylogenetic analysis. Detected ORF7a variants appear to
have emerged independently on every continent and were not
confined to any single lineage (Figure 1B; Table S1). Next, we
aligned translated ORF7a sequences to look for patterns. Most
of the identified ORF7a mutations (126 out of 189) truncate the
C terminus but preserve the N-terminal half of the protein (Fig-
ure 1C; Figure S1B). A portion (36.5%) of these truncated
ORF7a variants appeared in two or more patient samples (116
maximum [max]). Although some of these ORF7a mutants
appear to have arisen on multiple independent occasions, others
come from genomes that form monophyletic clades, suggesting
that viruses with ORF7a truncations are capable of transmission
within a host population (Figure S1B).
ORF7a truncation has a loss-of-function effect
ORF7a is a type I transmembrane (TM) protein with an N-terminal
immunoglobulin-like (Ig-like) ectodomain, stalk, TM domain, and
short cytosolic tail (Figure 2A). The two bsheets of the ectodo-
main are held together by two disulfide bonds (i.e., Cys23-
Cys58 and Cys35-Cys67) (Figure 2B). The cytosolic tail of
ORF7a contains a dilysine (KRKTE) endoplasmic reticulum (ER)
retrieval signal (ERRS) that mediates protein trafficking to the
ER-Golgi intermediate compartment (ERGIC) (Nelson et al.,
2005).
The D115 nt mutation in ORF7a introduces a premature stop
codon that eliminates b5, b6, and b7, two of the cystines that
form disulfides (Cys58 and Cys67), the TM, and the cytosolic
tail (Figures 2A and 2B). This truncation is expected to destabilize
the protein structure and significantly impact protein function
and trafficking in the host cell. To determine how loss of TM
and sorting signal affects protein localization, we cloned and ex-
pressed FLAG-tagged wild-type and truncated ORF7a proteins
in HEK293T-hACE2 cells. The wild-type ORF7a accumulated in
the perinuclear region of the cell, which is consistent with previ-
ously reported ERGIC localization (Figure 2D) (Martin-Sancho
et al., 2020;Nelson et al., 2005). In contrast, the truncated
ORF7a is distributed throughout the cytoplasm and does not
associate with specific subcellular compartments, which is
consistent with the loss of the TM domain and the ERRS signals
required for protein targeting (Figure 2E).
Extent of the truncation and the defect in intracellular targeting
suggests possible loss of ORF7a function. Along with several
SARS-CoV-2 accessory proteins, ORF7a has been implicated
in suppression of host IFN response to the infection (Xia et al.,
2020). To test the effect of the truncated ORF7a protein on IFN
signaling, we generated reporter HEK293T cells with integrated
luciferase reporter gene controlled by the interferon-stimulated
response element (ISRE). In these cells, activation of interferon
response drives reporter expression, which is quantified
using a luciferase assay. To test immune suppression, we over-
expressed the wild-type and the deletion mutant of ORF7a
(ORF7a
D115
) from plasmids and the stimulated cells with IFNa2b.
The ORF7a WT protein suppressed IFN response by 34%
(p = 0.0074) compared with cells transfected with control
plasmid, which is consistent with recent results by Xia et al.
(2020). In contrast, the ORF7a
D115
protein failed to suppress
IFN response and showed no significant difference to a control
(mCherry) plasmid (Figure 2F). These results indicate that the
D115 mutation in ORF7a has a loss-of-function effect that
negates antagonism of IFN signaling.
ORF7a mutation has no collateral effect on ORF7b
The ORF7a gene overlaps with the downstream ORF7b gene
(Figure S2). To determine if the D115mutationinORF7aim-
pacts ORF7b, we first examined whether it eliminates a tran-
scription-regulatory sequence (TRS) that is required for subge-
nomic RNA (sgRNA) synthesis. None of the TRSs identified via
direct RNA sequencing (RNA-seq) of SARS-CoV-2 overlap
with the D115 mutation site (Kim et al., 2020), suggesting
that ORF7b transcription is not affected (Figure S2). To
confirm this prediction, we used RT-PCR with primers that
detect ORF7a and ORF7b sgRNAs (Figure 2G). In this assay,
we infected 293T-hACE2 cells (MOI = 0.05) with ORF7a
WT
or
ORF7a
D115
viral strains and extracted total RNA from cells at
24 h postinfection (hpi). Both viruses produced specific RT-
PCR products corresponding to ORF7a and ORF7b sgRNAs.
Finally, to verify that the ORF7a
D115
variant does not impact
ORF7b translation, we probed cell lysates 24 hpi with an
anti-ORF7b antibody (Figure 2H). We detected ORF7b protein
in both viral strains with band intensities that suggest similar
expression levels. Collectively, these data indicate that the
ORF7a
D115
mutation has no collateral effect on the adjacent
gene.
ORF7a
D115
isolate displays a replication defect
To investigate the phenotype of SARS-CoV-2 D115 variant, we
infected 293T-hACE2 and Vero E6 cells with ORF7a
WT
and
ORF7a
D115
viruses at an MOI of 0.05 and measured viral RNA
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 3
Report
ll
OPEN ACCESS
replication using qRT-PCR (Figures 3A and 3B). In this assay, we
used viral strains sharing the same haplotype (C241T, C3037T,
C14408T, A23403G) that defines the SARS-CoV-2 B.1 lineage
and its derivatives (Figure S3A) (Rambaut et al., 2020a). Over
the course of 120 h, ORF7a
WT
viral RNA level increased 960-
fold in supernatants from Vero E6 cells (Figure 3A) and 270-
fold in supernatants from 293T-hACE2 cells (Figure 3B). The
ORF7a
D115
virus displayed limited replication that resulted in
only 18.6-fold and 7-fold RNA level increase at 120 hpi in VeroE6
and 293T-hACE2, respectively. This growth defect reaches
statistical significance starting at 6 hpi in Vero E6 cells (p =
0.0135) and 24 hpi in 293T-hACE2 cells (p = 6.2 310
5
).
To determine if this phenotype results from a defect in viral
egress only or if viral RNA replication in the host is also
affected, we examined early steps in the infection. We ex-
tracted total RNA from the infected 293T-hACE2 cells at 0.5,
6, and 24 hpi and measured SARS-CoV-2 RNA with qRT-
PCR. Compared with wild type, ORF7a
D115
RNA level was
reduced 2.1-fold inside the host at 6 hpi (p = 0.002) and 4.3-
fold at 24 hpi (p = 0.02) (Figure 3C). This decrease in viral
Figure 2. ORF7a truncation results in loss of function
(A) Amino acid (aa) sequence alignment of SARS-CoV-2 ORF7a
WT
, ORF7a
D115
, and SARS-CoV-1 ORF7a. Gaps show non-matching positions; red shows 17-aa
sequence resulting from a frameshift in the ORF7a mutant. Beta strands (arrows) and alpha helices (coil) are shown above the alignment.
(B) Diagram of SARS-CoV-2 ORF7a Ig-like fold. Disulfide bonds that stabilize the bsandwich structure are shown with red lines. The portion of the protein
eliminated by the deletion is shown in gray.
(C) C-terminally FLAG-tagged ORF7a
WT
and ORF7a
D115
were cloned and overexpressed in HEK293T-hACE2. Protein expression was confirmed with western
blot using anti-FLAG antibody. b-Actin (ACTB) was used as a loading control.
(D and E) FLAG-tagged ORF7a
WT
(D) or ORF7a
D115
(E) expressed in HEK293T-hACE2 cells. Immunostaining was performed using an anti-FLAG antibody (green).
Cell nuclei were stained with Hoechst 33342 (blue). White scale bar is 10 mm.
(F) HEK293T cells with integrated ISRE-luciferase reporter were transfected with pLV-mCherry (‘‘-’’), ORF7a
WT
, or ORF7a
D115
plasmids. Transfected cells were
treated with 5 ng/mL human recombinant IFN-a2b for 24 h, and induction of type I IFN signaling was measured with luciferase assay. Fold induction versus non-
treated control was determined and normalized to mCherry control. Means (n = 6) were compared with one-way ANOVA (p = 0.00162). Pairwise comparisons
were performed using post hoc Tukey’s test. Data are shown as mean ±SD. **p < 0.01;
ns
p > 0.05.
(G) ORF7a and ORF7b sgRNAs were identified by RT-PCR. Lower molecular weight band corresponding to ORF7b sgRNA is indicated with asterisk (*). Diagram
on the right shows primer (arrows) positions. Specificity of PCR products was confirmed with Sanger sequencing. GAPDH was used as a control for cDNA
synthesis.
(H) Lysates from SARS-CoV-2-infected cells were probed with antibodies raised against ORF7b protein. ACTB was used as a loading control.
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
4Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
RNA in the host cell differs from the corresponding decrease in
the supernatant RNA at 24 hpi (4.3-fold versus 16.8-fold,
respectively; Figures 3B and 3C). This disparity suggests that
observed growth defect in the SARS-CoV-2 D115 variant might
result from defects in genomic RNA replication and transcrip-
tion, as well as in viral egress.
Figure 3. The ORF7a
D115
mutation results in replication and immune suppression defects
(A and B) VeroE6 (A) or HEK293T-hACE2 (B) cells were infected with ORF7a
WT
or ORF7a
D115
SARS-CoV-2 strain at MOI = 0.05. Viral RNA was measured in
supernatant using qRT-PCR at different time points after infection. Measured RNA levels were normalized (using DCt method) to RNA levels at 0.5 hpi.
(C) Total RNA was extracted from infected HEK293T-hACE2 cells, and viral RNA was measured. The DDCt method was used to normalize viral RNA levels to host
RNA (ACTB) and 0.5 hpi time point. Data in (A)–(C) are presented as mean ±SD of three biolo gical replicates. Significance levels: *p < 0.05, **p < 0.01, ***p < 0.001,
or
ns
p > 0.05 (no significant difference).
(D) Volcano plot showing IFN-I response in 293T-hACE2 cells infected with ORF7a
WT
SARS-CoV-2 strain at MOI = 0.05. Expression of IFN-I response genes was
studied 24 hpi using qRT-PCR array targeting 91 human transcripts (88 targets and 3 references). Experiment was performe d in three biological replicates.
Dashed lines show regulation (R2-fold) and statistical significance (p < 0.05) thresholds. Each dot represents mean (n = 3) normalized expression of a single gene
relative to non-infected host. Genes that passed the threshold are labele d.
(E) Volcano plot showing IFN-I response in ORF7a
D115
versus ORF7a
WT
infection.
(F) Data shown in (D) and Figure S3A were plotted as a heatmap. Genes were classified into three groups. Group #1 genes are oppositely regulated between
the two viral strains. Group #2 genes are upregulated by both ORF7a
WT
and ORF7a
D115
. Group #3 genes are downregulated in both. Genes that have statistically
significant difference in expression between two viral strains are marked with asterisk. Gray shows genes with no detectable expression over 40 cyclesof
qRT-PCR.
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 5
Report
ll
OPEN ACCESS
ORF7a
D115
virus induces elevated IFN-I response in vitro
Suppressed or delayed IFN responses characterize SARS-CoV-
2 infections (Lei et al., 2020). The C-terminal truncation of ORF7a
limits suppression of the type I IFN signaling (Figure 2F). There-
fore, we hypothesized that ORF7a
D115
virus isolated from the
community is deficient in antagonizing IFN response.
To test this hypothesis, we measured expression of type I IFN
response genes upon infection with ORF7a
WT
and ORF7a
D115
strains using a qRT-PCR array targeting 91 human transcripts
(Table S2). The array included three reference genes that we
used to normalize expression of 88 target genes. In agreement
with recent transcriptomic studies, we detected a type I IFN-
mediated antiviral response after infecting HEK293T-hACE2
with SARS-CoV-2 isolates (Figure 3D; Figure S3B) (Banerjee
et al., 2020;Blanco-Melo et al., 2020). Out of 88 type I IFN
response genes included in the qRT-PCR array, 42 were differ-
entially expressed (R2-fold change, p < 0.05) in ORF7a
WT
-in-
fected cells and 55 in ORF7a
D115
infection, as compared with
the non-infected control (Figure 3D; Figure S3B). Both viruses
stimulated expression of multiple type I interferon genes, antiviral
ISGs, ILs, and proinflammatory cytokines (Table S2).
To understand the effect of ORF7a
D115
on the host type I IFN
response, we compared expression of genes targeted with
qRT-PCR array between the two viruses. In general, the ORF7a
-
D
115 variant induces elevated type I IFN response to the infec-
tion (Figures S3C and S3D). The median expression fold change
of 88 genes targeted in the qRT-PCR array was 1.72 in ORF7a
WT
and 2.51 in ORF7a
D115
virus (p = 1.4E12; Figure S3C). Analysis
of genes differentially expressed between the two viral infections
results in an asymmetric volcano plot (Figure 3E). A subset of
ISGs had expression levels that are significantly different (p <
0.05) between the two viral infections (Figures 3E and 3F; Table
S2). This subset includes sensors (TLR7), signal transducers
(MYD88, OAS2), transcriptional regulators (IRF3,IRF5), and re-
striction factors (GBP1,IFITM3,MX1) known to combat infec-
tions by RNA viruses (Schoggins, 2019). One of these differen-
tially expressed ISGs (IFITM3) restricts SARS-CoV-2 entry (Shi
et al., 2021;Zang et al., 2020). Additionally, polymorphisms in
IFITM3,MX1, and TLR7 are linked to the severity of COVID-19
(Andolfo et al., 2020;Zhang et al., 2020a;Pati et al., 2021). Upre-
gulation of ISGs indicates that the ORF7a
D115
virus is inept at
antagonizing the IFN-I response to the infection, which is consis-
tent with results from our ISRE reporter assay (Figure 2F).
DISCUSSION
In this study, we use genomic surveillance and phylogenetics to
identify ORF7a variants that have emerged globally throughout
the SARS-CoV-2 pandemic. Local occurrence of ORF7a muta-
tions has been reported by others (Addetia et al., 2020;Holland
et al., 2020;Rosenthal et al., 2020). However, the effect of these
mutations on viral fitness and host immune responses has not
been investigated. Here, we isolated a virus with a deletion mu-
tation in ORF7a (ORF7a
D115
) and show that this genotype results
in an in vitro growth defect. This growth defect is associated with
elevated IFN response to SARS-CoV-2.
Interferon systems are a frontline of host defense that signals
infection and elicits an antiviral state (McNab et al., 2015). Pre-
treatment with type I, type II, and type III IFNs restricts SARS-
CoV-2 infection and replication in cell culture (Mantlo et al.,
2020;Miorin et al., 2020). SARS-CoV-2 deploys multiple proteins
(e.g., nsp6, nsp13, ORF6, ORF7a, ORF7b, and ORF9b) to shut
down IFN signaling and dampen innate immune responses
(Jiang et al., 2020;Lei et al., 2020;Xia et al., 2020), in particular,
when overexpressed ORF7a subverts STAT2 phosphorylation,
blocking IFN-dependent transcriptional activation of antiviral
ISGs (Xia et al., 2020). Both type I and III IFNs signal through
the JAK-STAT pathway and overlap substantially in the tran-
scriptional responses they induce (Kotenko and Durbin, 2017).
Therefore, ORF7a likely suppresses both IFN types; however,
its role in suppressing the type III IFN response has yet to be veri-
fied (Xia et al., 2020). Using the ISRE reporter system, we show
that C-terminal truncation negates the ability of ORF7a to sup-
press IFN signaling (Figure 2). The ORF7a
D115
variant shows
limited replication in HEK293T-hACE2, as well as in Vero E6,
the latter of which does not express type I IFN genes (Emeny
and Morgan, 1979;Osada et al., 2014). Although type I IFN genes
are deleted in Vero E6 cells, the type III IFNs are intact and are
activated upon infection with RNA viruses, which results in ISG
upregulation (Stoltz and Klingstro
¨m, 2010;Wang et al., 2020).
ISGs exert anti-viral activities that combat the infection (Schog-
gins, 2019). We hypothesize that the replication defect of the
ORF7a
D115
strain in Vero E6 and 293T-hACE2 cells results
from an IFN-dependent response that the virus fails to suppress.
Several groups have replaced ORF7a with a reporter gene
(Hou et al., 2020;Thi Nhu Thao et al., 2020;Xie et al., 2020).
Thi Nhu Thao et al. (2020) show that complete deletion of
ORF7a reduces replication of the synthetic virus, which
agrees well with growth defect in the ORF7a
D115
strain.
However, two other studies have not observed this effect.
We believe that our results with a naturally occurring
ORF7a deletion provide valuable insight to this discussion.
Finally, it should be noted that different mutations (i.e., com-
plete deletion versus partial) might have a different outcome
for the viral phenotype.
Cytokine profiling and RNA-seq of clinical samples reveal sup-
pressed or delayed IFN responses in COVID-19 patients (Aruna-
chalam et al., 2020;Hadjadj et al., 2020;Lucas et al., 2020). Dys-
regulated immune responses coupled with exacerbated
inflammatory cytokines drive pathogenesis of the disease and
correlate with its severity (Zhang et al., 2020b;Tay et al., 2020;
Zhou et al., 2020). Here, we detected enhanced IFN signaling
in cells infected with a naturally occurring ORF7a
D115
variant of
SARS-CoV-2.
Given the importance of the IFN response in COVID-19, we
anticipate that ORF7a mutations might affect SARS-CoV-2 path-
ogenicity. Clinical comparisons will be required to examine this
potential effect on COVID-19 features.
Deletions in ORF7a, ORF7b, and ORF8 have previously been
reported (Addetia et al., 2020;Holland et al., 2020;Rosenthal
et al., 2020;Su et al., 2020). Similar deletions emerged in Middle
East Respiratory Syndrome (MERS)- and SARS-CoV, which
were linked to viral attenuation (Chinese SARS Molecular Epide-
miology Consortium, 2004;Muth et al., 2018). A 382-nt deletion
in ORF8 is associated with milder SARS-CoV-2 infection in pa-
tients (Young et al., 2020). Although this mutation affects clinical
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
6Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
features of COVID-19, it has no evident effect on viral replication
or host transcriptional responses in human cell culture (Gamage
et al., 2020).
Several studies illustrate how persistent SARS-CoV-2 infec-
tions in immunocompromised hosts associate with accelerated
viral evolution (Avanzato et al., 2020;Choi et al., 2020). In these
hosts there is less selective pressure on viral anti-immune pro-
teins, which might permit loss-of-function mutations in accessory
genes. The ORF7a
D115
variant that we isolatedwas present in only
seven genomes over a course of 1.5 months in a local commu-
nity and then disappeared. We hypothesize that deletions in
accessory genes, such as ORF7a
D115
, might emerge in immuno-
compromised patients but do not persist in the immunocompe-
tent population. Similarly, a deletion mutation in ORF8 (i.e.,
D382), which was associated with milder COVID-19 disease, dis-
appeared soon after it was discovered (Young et al., 2020). How-
ever, founder effect or mutations that co-occur with other muta-
tions that increase transmissibility (e.g., spike mutations) can
result in fixation of an otherwise attenuating mutation. A new
more transmissible lineage of SARS-CoV-2 (i.e., B.1.1.7) has
recently emerged (Davies et al., 2021;Rambaut et al., 2020b).
Aside from changes in the Spike protein, this lineage also includes
a premature stop codon in the accessory ORF8 that truncates
most of the protein (Volz et al., 2021). This lineage might illustrate
the second scenario for fixing mutations in accessory genes.
Limitations of study
The viral replication defect and cellular immune response assay
presented here were performed using HEK293T-hACE2 and
Vero E6 cells that are not the natural targets of SARS-CoV-2.
However, both cell lines are permissive for SARS-CoV-2 infec-
tion and are used extensively to study efficacy of antiviral com-
pounds (Riva et al., 2020), virion assembly (Klein et al., 2020),
SARS-CoV-2 entry (Hoffmann et al., 2020), replication (Thi Nhu
Thao et al., 2020), and anti-immune activities of SARS-CoV-2
proteins (Xia et al., 2020). Although we acknowledge that our
conclusions are limited to Vero E6 and 293T-hACE2 cells, we
anticipate that the results presented here will provide important
context for clinical comparisons, similar to those recently pub-
lished for ORF8 (Young et al., 2020).
Studies of naturally occurring SARS-CoV-2 variants have limita-
tions, including unavailability of isogenic controls. Here, we
compared two viral isolates, ORF7a
WT
and ORF7a
D115
,thatin
addition to D115 differ at several other nucleotide positions (Fig-
ure S3A). Although ISRE reporter assay unambiguously links the
immunosuppression defect to the truncation of ORF7a, it is not
possible to rule out a role for these mutations on the viral life cycle.
STAR+METHODS
Detailed methods are provided in the online version of this paper
and include the following:
dKEY RESOURCES TABLE
dRESOURCE AVAILABILITY
BLead contact
BMaterials availability
BData and code availability
dEXPERIMENTAL MODEL AND SUBJECT DETAILS
BCell cultures
BPlasmids
BAntibodies
BHuman clinical sample collection and preparation
BProduction and titration of coronavirus stocks
BSymptom onset data and clinical test results
dMETHOD DETAILS
BQuantitative reverse transcription PCR (qRT-PCR)
BRT-PCR and SARS-CoV-2 genome sequencing
BSARS-CoV-2 genome assembly
BRT-PCR and Sanger sequencing
BPhylogenetic and ORF7a mutational analysis
BWestern blot
BImmunocytochemistry
BIFN inhibition assay
BSARS-CoV-2 replication assays
BType I interferon response assay
dQUANTIFICATION AND STATISTICAL ANALYSIS
SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.
celrep.2021.109197.
ACKNOWLEDGMENTS
We are grateful to members of Bozeman Health that provided de-identified pa-
tient samples, specifically C. Nero, D. Smoot, W. Wallace, C. Faurot-Daniels,
V. Lawrence, and M. Blauvelt. We are also grateful to M. Flenniken, K. Daugh-
enbaugh, and other members of the COVID task force at MSU for assistanc e
establishing the COVID testing center. We thank the Lefcort lab for generous
use of their fluorescent confocal microscope and Dr. Pincus for providing
the HEK 293T-hACE2 cells. Research in the Wiedenheft lab is supported by
the National Institutes of Health (1R35GM134867), the Montana State Univer-
sity Agricultural Experimental Station, the MJ Murdock Charitable Trust, the
Gianforte Foundation, and the MSU Office of the Vice President for Research.
We thank the GISAID’s EpiFlu Database and contributing laboratories (Table
S4). The phylogenetic analysis in this paper would not have been possible
without their willingness to share data. The graphical abstract was created
with BioRender.
AUTHOR CONTRIBUTIONS
Conceptualization, B.W., A. Nemudryi, and A. Nemudraia; methodology, B.W.,
A. Nemudraia, A. Nemudryi, D.T.S., and J.F.H.; sample acquisition, D.B. and
B.W.; investigation & data collection, A. Nemudraia, A. Nemudryi, D.T.S.,
J.F.H., J.N., H.L., and K.K.V.; genomics and bioinformatics analysis, T.W., A.
Nemudryi, and C.C.; writing – original draft, B.W., A. Nemudryi, A. Nemudraia,
and T.W.; writing – review & editing, B.W., A. Nemudryi, A. Nemudraia, T.W.,
and M.A.J.
DECLARATION OF INTERESTS
B.W. is the founder of SurGene LLC and VIRIS Detection Systems Inc. B.W., A.
Nemudryi, and A. Nemudraia are inventors on patents related to CRISPR-Cas
systems and applications thereof.
Received: February 3, 2021
Revised: April 4, 2021
Accepted: May 10, 2021
Published: June 1, 2021
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 7
Report
ll
OPEN ACCESS
REFERENCES
Addetia, A., Xie, H., Roychoudhury, P., Shrestha, L., Loprieno, M., Huang,
M.-L., Jerome, K.R., and Greninger, A.L. (2020). Identification of multiple large
deletions in ORF7a resulting in in-frame gene fusions in clinical SARS-CoV-2
isolates. J. Clin. Virol. 129, 104523.
Andersen, K.G., Rambaut, A., Lipkin, W.I., Holmes, E.C., and Garry, R.F.
(2020). The proximal origin of SARS-CoV-2. Nat. Med. 26, 450–452.
Andolfo, I., Russo, R., Lasorsa, A.V., Cantalupo, S., Rosato, B.E., Bonfiglio, F.,
Frisso, G., Pasquale, A., Cassese, G.M., Servillo, G., et al. (2020). Common
variants at 21q22.3 locus influence MX1 gene expression and susceptibility
to severe COVID-19. medRxiv. https://doi.org/10.1101/2020.12.18.20248470.
Arunachalam, P.S., Wimmers, F., Mok, C.K.P., Perera, R.A.P.M., Scott, M.,
Hagan, T., Sigal, N., Feng, Y., Bristow, L., Tak-Yin Tsang, O., et al. (2020). Sys-
tems biological assessment of immunity to mild versus severe COVID-19
infection in humans. Science 369, 1210–1220.
Avanzato, V.A., Matson, M.J., Seifert, S.N., Pryce, R., Williamson, B.N., An-
zick, S.L., Barbian, K., Judson, S.D., Fischer, E.R., Martens, C., et al. (2020).
Case Study: Prolonged Infectious SARS-CoV-2 Shedding from an Asymptom-
atic Immunocompromised Individual with Cancer. Cell 183, 1901–1912.e9.
Banerjee, A., El-Sayes, N., Budylowski, P., Richard, D., Maan, H., Aguiar, J.A.,
Baid, K., D’Agostino, M.R., Ang, J.C., Tremblay, B.J.M., et al. (2020). Experi-
mental and natural evidence of SARS-CoV-2 infection-induced activation of
type I interferon responses. bioRxiv. https://doi.org/10.1101/2020.06.18.
158154.
Bedford, T., Greninger, A.L., Roychoudhury, P., Starita, L.M., Famulare, M.,
Huang, M.-L., Nalla, A., Pepper, G., Reinhardt, A., Xie, H., et al.; Seattle Flu
Study Investigators (2020). Cryptic transmission of SARS-CoV-2 in Washing-
ton state. Science 370, 571–575.
Blanco-Melo, D., Nilsson-Payant, B.E., Liu, W.-C., Uhl, S., Hoagland, D., Møl-
ler, R., Jordan, T.X., Oishi, K., Panis, M., Sachs, D., et al. (2020). Imbalanced
Host Response to SARS-CoV-2 Drives Development of COVID-19. Cell 181,
1036–1045.e9.
Chinese SARS Molecular Epidemiology Consortium (2004). Molecular evolu-
tion of the SARS coronavirus during the course of the SARS epidemic in China.
Science 303, 1666–1669.
Choi, B., Choudhary, M.C., Regan, J., Sparks, J.A., Padera, R.F., Qiu, X., Sol-
omon, I.H., Kuo, H.-H., Boucau, J., Bowman, K., et al. (2020). Persistence and
Evolution of SARS-CoV-2 in an Immunocompromised Host. N. Engl. J. Med.
383, 2291–2293.
Davies, N.G., Abbott, S., Barnard, R.C., Jarvis, C.I., Kucharski, A.J., Munday,
J.D., Pearson, C.A.B., Russell, T.W., Tully, D.C., Washburne, A.D., et al.;
CMMID COVID-19 Working Group; COVID-19 Genomics UK (COG-UK) Con-
sortium (2021). Estimated transmissibility and impact of SARS-CoV-2 lineage
B.1.1.7 in England. Science 372, eabg3055.
Emeny, J.M., and Morgan, M.J. (1979). Regulation of the interferon system: ev-
idence that Vero cells have a genetic defect in interferon production. J. Gen.
Virol. 43, 247–252.
Gamage, A.M., Tan, K.S., Chan, W.O.Y., Liu, J., Tan, C.W., Ong, Y.K., Thong,
M., Andiappan, A.K., Anderson, D.E., Wang, Y., and Wang, L.F. (2020). Infec-
tion of human Nasal Epithelial Cells with SARS-CoV-2 and a 382-nt deletion
isolate lacking ORF8 reveals similar viral kinetics and host transcriptional pro-
files. PLoS Pathog. 16, e1009130.
Grubaugh, N.D., Gangavarapu, K., Quick, J., Matteson, N.L., De Jesus, J.G.,
Main, B.J., Tan, A.L., Paul, L.M., Brackney, D.E., Grewal, S., et al. (2019). An
amplicon-based sequencing framework for accurately measuring intrahost vi-
rus diversity using PrimalSeq and iVar. Genome Biol 20,8.
Hachim, A., Kavian, N., Cohen, C.A., Chin, A.W.H., Chu, D.K.W., Mok, C.K.P.,
Tsang, O.T.Y., Yeung, Y.C., Perera, R.A.P.M., Poon, L.L.M., et al. (2020). ORF8
and ORF3b antibodies are accurate serological markers of early and late
SARS-CoV-2 infection. Nat. Immunol. 21, 1293–1301.
Hadjadj, J., Yatim, N., Barnabei, L., Corneau, A., Boussier, J., Smith, N., Pe
´re
´,
H., Charbit, B., Bondet, V., Chenevier-Gobeaux, C., et al. (2020). Impaired type
I interferon activity and inflammatory responses in severe COVID-19 patients.
Science 369, 718–724.
Harcourt, J., Tamin, A., Lu, X., Kamili, S., Sakthivel, S.K., Murray, J., Queen, K.,
Tao, Y., Paden, C.R., Zhang, J., et al. (2020). Severe Acute Respiratory Syn-
drome Coronavirus 2 from Patient with Coronavirus Disease, United States.
Emerg. Infect. Dis. 26, 1266–1273.
Hoffmann, M., Kleine-Weber, H., Schroeder, S., Kr
uger, N., Herrler, T.,
Erichsen, S., Schiergens, T.S., Herrler, G., Wu, N.H., Nitsche, A., et al.
(2020). SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is
Blocked by a Clinically Proven Protease Inhibitor. Cell 181, 271–280.e8.
Holland, L.A., Kaelin, E.A., Maqsood, R., Estifanos, B., Wu, L.I., Varsani, A.,
Halden, R.U., Hogue, B.G., Scotch, M., and Lim, E.S. (2020). An 81-Nucleotide
Deletion in SARS-CoV-2 ORF7a Identified from Sentinel Surveillance in Ari-
zona (January to March 2020). J. Virol. 94, e00711-20.
Holshue, M.L., DeBolt, C., Lindquist, S., Lofy, K.H., Wiesman, J., Bruce, H.,
Spitters, C., Ericson, K., Wilkerson, S., Tural, A., et al.; Washington State
2019-nCoV Case Investigation Team (2020). First Case of 2019 Novel Corona-
virus in the United States. N. Engl. J. Med. 382, 929–936.
Hou, Y.J., Okuda, K., Edwards, C.E., Martinez, D.R., Asakura, T., Dinnon, K.H.,
3rd, Kato, T., Lee, R.E., Yount, B.L., Mascenik, T.M., et al. (2020). SARS-CoV-2
Reverse Genetics Reveals a Variable Infection Gradient in the Respiratory
Tract. Cell 182, 429–446.e14.
Jiang, H.W., Zhang, H.N., Meng, Q.F., Xie, J., Li, Y., Chen, H., Zheng, Y.X.,
Wang, X.N., Qi, H., Zhang, J., et al. (2020). SARS-CoV-2 Orf9b suppresses
type I interferon responses by targeting TOM70. Cell. Mol. Immunol. 17,
998–1000.
Kim, D., Lee, J.-Y., Yang, J.-S., Kim, J.W., Kim, V.N., and Chang, H. (2020). The
Architecture of SARS-CoV-2 Transcriptome. Cell 181, 914–921.e10.
Klein, S., Cortese, M., Winter, S.L., Wachsmuth-Melm, M., Neufeldt, C.J., Ceri-
kan, B., Stanifer, M.L., Boulant, S., Bartenschlager, R., and Chlanda, P. (2020).
SARS-CoV-2 structure and replication characterized by in situ cryo-electron
tomography. bioRxiv. https://doi.org/10.1101/2020.06.23.167064.
Konno, Y., Kimura, I., Uriu, K., Fukushi, M., Irie, T., Koyanagi, Y., Sauter, D.,
Gifford, R.J., Nakagawa, S., and Sato, K.; USFQ-COVID19 Consortium
(2020). SARS-CoV-2 ORF3b Is a Potent Interferon Antagonist Whose Activity
Is Increased by a Naturally Occurring Elongation Variant. Cell Rep. 32, 108185.
Korber, B., Fischer, W.M., Gnanakaran, S., Yoon, H., Theiler, J., Abfalterer, W.,
Hengartner, N., Giorgi, E.E., Bhattacharya, T., Foley, B., et al.; Sheffield
COVID-19 Genomics Group (2020). Tracking Changes in SARS-CoV-2 Spike:
Evidence that D614G Increases Infectivity of the COVID-19 Virus. Cell 182,
812–827.e19.
Kotenko, S.V., and Durbin, J.E. (2017). Contribution of type III interferons to
antiviral immunity: location, location, location. J. Biol. Chem. 292, 7295–7303.
Lei, X., Dong, X., Ma, R., Wang, W., Xiao, X., Tian, Z., Wang, C., Wang, Y., Li, L.,
Ren, L., et al. (2020). Activation and evasion of type I interferon responses by
SARS-CoV-2. Nat. Commun. 11, 3810.
Lemieux, J.E., Siddle, K.J., Shaw, B.M., Loreth, C., Schaffner, S.F., Gladden-
Young, A., Adams, G., Fink, T., Tomkins-Tinch, C.H., Krasilnikova, L.A., et al.
(2021). Phylogenetic analysis of SARS-CoV-2 in Boston highlights the impact
of superspreading events. Science 371, eabe3261.
Li, H. (2018). Minimap2: pairwise alignment for nucleotide sequences. Bioin-
formatics 34, 3094–3100.
Loveday, E.K., Hain, K.S., Kochetkova, I., Hedges, J.F., Robison, A., Snyder,
D.T., Brumfield, S.K., Young, M.J., Jutila, M.A., Chang, C.B., and Taylor,
M.P. (2021). Effect of Inactivation Methods on SARS-CoV-2 Virion Protein
and Structure. Viruses 13, 562.
Lucas, C., Wong, P., Klein, J., Castro, T.B.R., Silva, J., Sundaram, M., Elling-
son, M.K., Mao, T., Oh, J.E., Israelow, B., et al.; Yale IMPACT Team (2020).
Longitudinal analyses reveal immunological misfiring in severe COVID-19. Na-
ture 584, 463–469.
Mantlo, E., Bukreyeva, N., Maruyama, J., Paessler, S., and Huang, C. (2020).
Antiviral activities of type I interferons to SARS-CoV-2 infection. Antiviral
Res. 179, 104811.
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
8Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
Martin-Sancho, L., Lewinski, M.K., Pache, L., Stoneham, C.A., Yin, X., Pratt,
D., Churas, C., Rosenthal, S.B., Liu, S., De Jesus, P.D., et al. (2020). Functional
Landscape of SARS-CoV-2 Cellular Restriction. bioRxiv. https://doi.org/10.
1101/2020.09.29.319566.
McNab, F., Mayer-Barber, K., Sher, A., Wack, A., and O’Garra, A. (2015). Type
I interferons in infectious disease. Nat. Rev. Immunol. 15, 87–103.
Miorin, L., Kehrer, T., Sanchez-Aparicio, M.T., Zhang, K., Cohen, P., Patel,
R.S., Cupic, A., Makio, T., Mei, M., Moreno, E., et al. (2020). SARS-CoV-2
Orf6 hijacks Nup98 to block STAT nuclear import and antagonize interferon
signaling. Proc. Natl. Acad. Sci. USA 117, 28344–28354.
Muth, D., Corman, V.M., Roth, H., Binger, T., Dijkman, R., Gottula, L.T., Gloza-
Rausch, F., Balboni, A., Battilani, M., Rihtari
c, D., et al. (2018). Attenuation of
replication by a 29 nucleotide deletion in SARS-coronavirus acquired during
the early stages of human-to-human transmission. Sci. Rep. 8, 15177.
Nelson, C.A., Pekosz, A., Lee, C.A., Diamond, M.S., and Fremont, D.H. (2005).
Structure and intracellular targeting of the SARS-coronavirus Orf7a accessory
protein. Structure 13, 75–85.
Osada, N., Kohara, A., Yamaji, T., Hirayama, N., Kasai, F., Sekizuka, T., Kur-
oda, M., and Hanada, K. (2014). The genome landscape of the african green
monkey kidney-derived vero cell line. DNA Res. 21, 673–683.
Pati, A., Padhi, S., Suvankar, S., and Panda, A.K. (2021). Minor Allele of Inter-
feron-Induced Transmembrane Protein 3 Polymorphism (rs12252) Is Covered
Against Severe Acute Respiratory Syndrome Coronavirus 2 Infectio n and Mor-
tality: A Worldwide Epidemiological Investigation. J. Infect. Dis. 223, 175–178.
Quick,J., Grubaugh, N.D.,Pullan, S.T., Claro,I.M., Smith, A.D.,Gangavarapu,K.,
Oliveira, G., Robles-Sikisaka, R., Rogers, T.F., Beutler, N.A., et al. (2017). Multi-
plex PCR method for MinION and Illumina sequencing of Zika and other virus
genomes directly from clinical samples. Nat. Protoc. 12, 1261–1276.
Rambaut, A., Holmes, E.C., O’Toole, A
´., Hill, V., McCrone, J.T., Ruis, C., du
Plessis, L., and Pybus, O.G. (2020a). A dynamic nomenclature proposal for
SARS-CoV-2 lineages to assist genomic epidemiology. Nat. Microbiol. 5,
1403–1407.
Rambaut, A., Loman, N., Pybus, O., Barclay, W., Barrett, J., Carabelli, A., Con-
nor, T., Peacock, T., Robertson, D.L., and Volz, E.; on behalf of COVID-19 Ge-
nomics Consortium UK (CoG-UK) (2020b). Preliminary genomic characterisa-
tion of an emergent SARS-CoV-2 lineage in the UK defined by a novel set of
spike mutations. https://virological.org/t/563.
Riva, L., Yuan, S., Yin, X., Martin-Sancho, L., Matsunaga, N., Pache, L.,
Burgstaller-Muehlbacher, S., De Jesus, P.D., Teriete, P., Hull, M.V., et al.
(2020). Discovery of SARS-CoV-2 antiviral drugs through large-scale com-
pound repurposing. Nature 586, 113–119.
Rosenthal, S.H., Kagan, R.M., Gerasimova, A., Anderson, B., Grover, D., Hua,
M., Liu, Y., Owen, R., and Lacbawan, F. (2020). Identification of eight SARS-
CoV-2 ORF7a deletion variants in 2,726 clinical specimens. bioRxiv. https://
doi.org/10.1101/2020.12.10.418855.
Schoggins, J.W. (2019). Interferon-Stimulated Genes: What Do They All Do?
Annu. Rev. Virol. 6, 567–584.
Shi, G., Kenney, A.D., Kudryashova, E., Zani, A., Zhang, L., Lai, K.K., Hall--
Stoodley, L., Robinson, R.T., Kudryashov, D.S., Compton, A.A., and Yount,
J.S. (2021). Opposing activities of IFITM proteins in SARS-CoV-2 infection.
EMBO J. 40, e106501.
Stoltz, M., and Klingstro
¨m, J. (2010). a/binterferon (IFN-a/b)-independent in-
duction of IFN-l1 (interleukin-29) in response to Hantaan virus infection.
J. Virol. 84, 9140–9148.
Su, Y.C.F., Anderson, D.E., Young, B.E., Linster, M., Zhu, F., Jayakumar, J.,
Zhuang, Y., Kalimuddin, S., Low, J.G.H., Tan, C.W., et al. (2020). Discovery
and Genomic Characterization of a 382-Nucleotide Deletion in ORF7b and
ORF8 during the Early Evolution of SARS-CoV-2. MBio 11, e01610-20.
Tan, Y., Schneider, T., Leong, M., Aravind, L., and Zhang, D. (2020). Novel
Immunoglobulin Domain Proteins Provide Insights into Evolution and Patho-
genesis of SARS-CoV-2-Related Viruses. MBio 11, e00760-20.
Tay, M.Z., Poh, C.M., Re
´nia, L., MacAry, P.A., and Ng, L.F.P. (2020). The trinity
of COVID-19: immunity, inflammation and intervention. Nat. Rev. Immunol. 20,
363–374.
Thi Nhu Thao, T., Labroussaa, F., Ebert, N., V’kovski, P., Stalder, H., Portmann,
J., Kelly, J., Steiner, S., Holwerda, M., Kratzel, A., et al. (2020). Rapid
reconstruction of SARS-CoV-2 using a synthetic genomics platform. Nature
582, 561–565.
Tyson, J.R., James, P., Stoddart, D., Sparks, N., Wickenhagen, A., Hall, G.,
Choi, J.H., Lapointe, H., Kamelian, K., Smith, A.D., et al. (2020). Improvements
to the ARTIC multiplex PCR method for SARS-CoV-2 genome sequencing us-
ing nanopore. bioRxiv. https://doi.org/10.1101/2020.09.04.283077.
van Dorp, L., Richard, D., Tan, C.C.S., Shaw, L.P., Acman, M., and Balloux, F.
(2020). No evidence for increased transmissibility from recurrent mutations in
SARS-CoV-2. Nat. Commun. 11, 5986.
Volz, E., Mishra, S., Chand, M., Barrett, J.C., Johnson, R., Geidelberg, L., Hins-
ley, W.R., Laydon, D.J., Dabrera, G., O’Toole, A
´., et al. (2021). Transmission of
SARS-CoV-2 Lineage B.1.1.7 in England: Insights from linking epidemiological
and genetic data. medRxiv. https://doi.org/10.1101/2020.12.30.20249034.
Wang, C., Shan, L., Qu, S., Xue, M., Wang, K., Fu, F., Wang, L., Wang, Z., Feng,
L., Xu, W., and Liu, P. (2020). The Coronavirus PEDV Evades Type III Interferon
Response Through the miR-30c-5p/SOCS1 Axis. Front. Microbiol. 11, 1180.
Wu, F., Zhao, S., Yu, B., Chen, Y.M., Wang, W., Song, Z.G., Hu, Y., Tao, Z.W.,
Tian, J.H., Pei, Y.Y., et al. (2020). A new coronavirus associated with human
respiratory disease in China. Nature 579, 265–269.
Xia, H., Cao, Z., Xie, X., Zhang, X., Chen, J.Y.-C., Wang, H., Menachery, V.D.,
Rajsbaum, R., and Shi, P.-Y. (2020). Evasion of Type I Interferon by SARS-
CoV-2. Cell Rep. 33, 108234.
Xie, X., Muruato, A., Lokugamage, K.G., Narayanan, K., Zhang, X., Zou, J., Liu,
J., Schindewolf, C., Bopp, N.E., Aguilar, P.V., et al. (2020). An Infectious cDNA
Clone of SARS-CoV-2. Cell Host Microbe 27, 841–848.e3.
Young, B.E., Fong, S.-W., Chan, Y.-H., Mak, T.-M., Ang, L.W., Anderson, D.E.,
Lee, C.Y.-P., Amrun, S.N., Lee, B., Goh, Y.S., et al. (2020). Effects of a major
deletion in the SARS-CoV-2 genome on the severity of infection and the inflam-
matory response: an observational cohort study. Lancet 396, 603–611.
Zang, R., Case, J.B., Yutuc, E., Ma, X., Shen, S., Gomez Castro, M.F., Liu, Z.,
Zeng, Q., Zhao, H., Son, J., et al. (2020). Cholesterol 25-hydroxylase sup-
presses SARS-CoV-2 replication by blocking membrane fusion. Proc. Natl.
Acad. Sci. USA 117, 32105–32113.
Zhang, Y., Zhang, J., Chen, Y., Luo, B., Yuan, Y., Huang, F., Yang, T., Yu, F.,
Liu, J., Liu, B., et al. (2020). The ORF8 Protein of SARS-CoV-2 Mediates Im-
mune Evasion through Potently Downregulating MHC-I. bioRxiv. http s://doi.
org/10.1101/2020.05.24.111823.
Zhang, Y., Qin, L., Zhao, Y., Zhang, P., Xu, B., Li, K., Liang, L., Zhang, C., Dai,
Y., Feng, Y., et al. (2020a). Interferon-Induced Transmembrane Protein 3 Ge-
netic Variant rs12252-C Associated With Disease Severity in Coronavirus Dis-
ease 2019. J. Infect. Dis. 222, 34–37.
Zhang, B., Zhou, X., Qiu, Y., Song, Y., Feng, F., Feng, J., Song, Q., Jia, Q., and
Wang, J. (2020b). Clinical characteristics of 82 cases of death from COVID-19.
PLoS ONE 15, e0235458.
Zhou, F., Yu, T., Du, R., Fan, G., Liu, Y., Liu, Z., Xiang, J., Wang, Y., Song, B.,
Gu, X., et al. (2020). Clinical course and risk factors for mortality of adult inpa-
tients with COVID-19 in Wuhan, China: a retrospective cohort study. Lancet
395, 1054–1062.
Zhou, Z., Huang, C., Zhou, Z., Huang, Z., Su, L., Kang, S., Chen, X., Chen, Q.,
He, S., Rong, X., et al. (2021). Structural Insight Reveals SARS-CoV-2 ORF7a
as an Immunomodulating Factor for Human CD14+ Monocytes. iScience 24,
102187.
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 9
Report
ll
OPEN ACCESS
STAR+METHODS
KEY RESOURCES TABLE
REAGENT or RESOURCE SOURCE IDENTIFIER
Biological samples
Nasopharyngeal swabs Bozeman Health Deaconess
Hospital, MT, USA
N/A
Bacterial and virus strains
SARS-CoV-2 Isolate USA-WA1/2020 BEI Resources NR-52281
Experimental models: cell lines
HEK293T-hACE2 cells BEI Resources NR-52511
Vero E6 ATCC Cat#CRL-1586
Recombinant DNA
pEGFP-N1 Clontech 6085-1
pLV-mCherry Addgene 36084
pORF7a-CoV2 This study N/A
pORF7a D115-Cov2 This study N/A
pGreenFire1-ISRE SBI TR016PA-1
Antibodies
Rabbit anti-ACTB ABclonal Cat. AC026, RRID:AB_2768234
Mouse anti-FLAG ThermoFisher Scientific Cat: MA1-91878-1MG, RRID:AB_2537619
Sheep anti-ORF7b MRC I PPU, College of Life
Sciences, University of Dundee
N/A
Goat anti-mouse IgG peroxidase-conjugated Jackson ImmunoResearch Cat: 115-035-003, RRID:AB_10015289
Goat anti-rabbit IgG peroxidase-conjugated Jackson ImmunoResearch Cat: 111-035-003, RRID:AB_2313567
Donkey Anti-Sheep IgG peroxidase-conjugated Jackson ImmunoResearch Cat: 713-035-147, RRID:AB_2340710
Mouse anti-FLAG Invitrogen Cat: MA1-91878-1MG, RRID:AB_2537619
Goat anti-mouse-AlexaFluor-488 Invitrogen Cat: A11001, RRID:AB_2534069
Chemicals, peptides, and recombinant proteins
QIAamp viral RNA mini kit QIAGEN 52904
2019-nCoV RUO kit IDT 10006713
Positive template control (PTC) plasmid IDT 10006625
TaqPath 1-Step RT-qPCR master mix Thermo Fisher Scientific A15300
SuperScript IV reverse transcriptase Thermo Fisher Scientific 18090010
R9.4.1 flow cells Nanopore Technologies FLO-MIN106
AMX, LNB, SFB, EB and SQB Nanopore Technologies SQK-LSK109
Flow cell priming kit Nanopore Technologies EXP-FLP002
NEBNext Ultra II end-prep New England Biolabs E7546S
NEBNext quick ligation module New England Biolabs E6056S
Native barcoding expansion kits Nanopore Technologies EXP-NBD104, EXP-NBD114
Q5 high-fidelity DNA polymerase New England Biolabs M0491S
DNA clean & concentrator kit Zymo Research D4005
Lipofectamine 3000 Invitrogen L3000-015
Pierce ECL western blotting substrate ThermoFisher Scientific #32106
X-Ray film Santa Cruz Biotech sc-201696
IFN-a2b InvivoGen rcyc-hifna2b
Luciferase assay system Promega E1500
(Continued on next page)
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
e1 Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
RESOURCE AVAILABILITY
Lead contact
Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, Blake Wie-
denheft (bwiedenheft@gmail.com).
Materials availability
Plasmids generated in this study are available upon request from the lead contact.
Data and code availability
The complete SARS-CoV-2 genome sequences are deposited to GISAID EpiCoV and GenBank databases. GISAID accession IDs
and URLs to GenBank records are provided in Table S1. Raw data for IFN response assay presented in Figures 3 and S3 is available
in Table S2.
Continued
REAGENT or RESOURCE SOURCE IDENTIFIER
Type I interferon response (SAB Target List)
H96 qPCR array
Bio-Rad N/A
SsoAdvanced universal SYBR green supermix Bio-Rad 1725270
Deposited data
URLs for SARs-CoV-2 genome sequences
are listed in the Table S1
GenBank N/A
Oligonucleotides
The oligonucleotides used in this study are
listed in the Table S3
IDT N/A
Software and algorithms
SDS software v1.4 Applied Biosystems 4379633
RStudio v1.2.1335 The R project https://www.r-project.org/
artic-ncov2019 ARTIC network https://artic.network/ncov-2019
MinKNOW software Oxford Nanopore
Technologies
https://community.nanoporetech.com/
protocols/experiment-companion-minknow/
v/mke_1013_v1_revbm_11apr2016
PrimePCR analysis software Bio-Rad https://www.bio-rad.com/en-us/category/qpcr-
analysis-software?ID=42a6560b-3ad7-43e9-
bb8d-6027371de67a
CD-HIT GitHub https://github.com/weizhongli/cdhit
BioStrings Bioconductor https://bioconductor.org/packages/release/bioc/
html/Biostrings.html
ggplot2 CRAN https://cran.r-project.org/web/packages/ggplot2/
index.html
Augur Nextstrain https://github.com/nextstrain/augur
Nextclade Nextstrain https://clades.nextstrain.org/
pangolin Github https://github.com/cov-lineages/pangolin
Trimal Github https://github.com/scapella/trimal
IQ-Tree IQ-Tree http://www.iqtree.org/
TreeTime GitHub https://github.com/jkanev/treetime
APE v5.3 CRAN https://cran.r-project.org/web/packages/ape/
index.html
tidyquant CRAN https://cran.r-project.org/web/packages/
tidyquant/index.html
stats CRAN https://stat.ethz.ch/R-manual/R-devel/library/
stats/html/00Index.html
UGENE Unipro http://ugene.net/
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 e2
Report
ll
OPEN ACCESS
EXPERIMENTAL MODEL AND SUBJECT DETAILS
Cell cultures
HEK293T-hACE2 cells (BEI, NR-52511) were generously provided by Dr. Seth Pincus (Montana State University) and maintained at
37C and 5% CO2 in Dulbecco’s Modified Eagle Medium (DMEM, GIBCO, cat. #12100-061) supplemented with 10% fetal bovine
serum (FBS, ATLAS Biologicals, Lot. #F31E18D1), sodium bicarbonate (3.7 g/L), 50 I.U./mL penicillin and 50 mg/mL streptomycin.
Vero E6 cells (ATCCcat. #CRL-1586) were maintained at 37C and 5% CO2 in Eagle’s Modified Eagle Medium (EMEM,
ATCC, cat. #30-2003) supplemented with 10% fetal bovine serum (FBS, ATLAS Biologicals, Lot. #F31E18D1) and 50 I.U./mL
penicillin and 50 mg/mL streptomycin. All cell lines tested negative for mycoplasma.
Plasmids
Human-codon optimized dsDNA gene fragments (gBlocks) encoding for SARS-CoV-2 C-terminally FLAG-tagged ORF7aWT and
ORF7aD115 proteins were synthesized by Integrated DNA Technologies (IDT). Sequence of ORF7a from Wuhan-Hu-1 reference
(GenBank MN908947.3) was used for ORF7aWT gBlock. The gene fragments were cloned into the pEGFP-N1 (Clontech cat#
6085-1) backbone to replace the eGFP gene. pLV-mCherry was a gift from Pantelis Tsoulfas (Addgene plasmid # 36084; http://
www.addgene.org/36084/; RRID:Addgene_36084).
Antibodies
Western blotting: Rabbit anti-ACTB (Cat: AC026, 1:20,000) antibodies were from ABclonal, Mouse anti-FLAG antibodies (Cat:
MA1-91878-1MG, 1:1000) were from ThermoFisher Scientific, sheep anti-ORF7b antibodies (1:120) were from MRC I PPU, Col-
lege of Life Sciences, University of Dundee, Scotland; Goat anti-mouse IgG peroxidase-conjugated (Cat: 115-035-003,
1:10,000), Goat anti-rabbit IgG peroxidase-conjugated (Cat: 111-035-003, 1:10,000) and Donkey Anti-Sheep IgG peroxidase-
conjugated (Cat: 713-035-147, 1:10,000) antibodies were from Jackson ImmunoResearch. Immunocytochemistry: Mouse
anti-FLAG antibodies (Cat: MA1-91878-1MG, 1:200), Goat anti-mouse-AlexaFluor-488 (Cat: A11001, 1:2,000) antibodies
were from Invitrogen.
Human clinical sample collection and preparation
Clinical samples were obtained with local IRB approval (protocol #DB033020) and informed consent from patients undergoing testing
for SARS-CoV-2 at Bozeman Health Deaconess Hospital. Patients tested for SARS-CoV-2 included both in-patients and out-pa-
tients. The latter included individuals who developed symptoms and sought medical care and asymptomatic (at the time of testing)
individuals who were exposed to known COVID-19 case and therefore were tested. Nasopharyngeal swabs from patients that tested
positive for SARS-CoV-2 were collected in viral transport media. RNA was extracted from all patient samples using QIAamp Viral RNA
Mini Kit (QIAGEN) a biosafety level 3 (BSL3) laboratory. All samples were heat-inactivated before removing from BSL3. Information
about age and sex of the patients is provided in Table S1.
Production and titration of coronavirus stocks
The nasopharyngeal swabs in the viral transport media were used to generate viral stocks as previously described (Harcourt et al.,
2020). Briefly, 100 ul of the viral transport media was two-fold serially diluted and applied to Vero E6 cells When infected cells
showed extensive cytopathic effect, the media was collected and utilized to generate a greater volume second passage in
Vero E6 cells. When CPE was apparent, supernatants were centrifuged 1,000 RCF for 5 minutes to remove cellular debris. Virus
titer in clarified supernatants was determined with plaque assay in Vero E6 cells as described (Loveday et al., 2021). For plaque
assay, Vero E6 cells were incubated with viral inoculum at limiting dilutions. Inoculated cells were overlayed with either 0.75%
methylcellulose, DMEM supplemented with 2% FBS and 1% pen-strep and incubated for 4 days. Cells were fixed and stained
with 0.5% methylene blue in 70% ethanol. Plaques were counted and the overall titer was calculated. Viral stock’s identity
was confirmed via whole genome sequencing on Oxford Nanopore. All SARS-CoV-2 experiments were performed in a BSL3
laboratory.
Symptom onset data and clinical test results
Suspect cases of COVID-19 were tested in a CLIA lab and instructed to self-quarantine until notified of the RT-qPCR test results. All
laboratory confirmed positive cases of COVID-19 were contacted via telephone by local public health nurses to complete contact
tracing. During this interview, the nurses collected recorded symptoms, symptom onset date, travel history, contact with other known
laboratory confirmed cases, close contacts and activities on the two days before symptom onset up until notification of a positive
test. Data collection was conducted as part of a public health response. The study was reviewed by the Montana State University
Institutional Review Board (IRB) For the Protection of Human Subjects (FWA 00000165) and was exempt from IRB oversight in accor-
dance with Code of Federal regulations, Part 46, section 101. All necessary patient/participant consent has been obtained and the
appropriate institutional forms have been archived.
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
e3 Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
METHOD DETAILS
Quantitative reverse transcription PCR (qRT-PCR)
qRT-PCR was performed using CDC primers (N1 and N2) and probes from the 2019-nCoV RUO Kit (IDT# 10006713). SARS-CoV-2
RNA was quantified using one-step qRT-PCR in ABI 7500 Fast Real-Time PCR System according to CDC protocol (https://www.fda.
gov/media/134922/download). In brief, 20 mL reactions included 8.5 mL of Nuclease-free Water, 1.5 mL of Primer and Probe mix (IDT,
10006713), 5 mL of TaqPath 1-Step RT-qPCR Master Mix (ThermoFisher, A15299) and 5 mL of the template. Nuclease-free water was
used as negative template control (NTC). Amplification was performed using following program: 25C for 2 min, 50C for 15 min, 95C
for 2 min followed by 45 cycles of 95C for 3 s and 55C for 30 s. Run data was analyzed in SDS software v1.4 (Applied Biosystems).
The NTC showed no amplification throughout the 40 cycles of qPCR.
RT-PCR and SARS-CoV-2 genome sequencing
For sequencing 10 mL of RNA from SARS-CoV-2 positive patient sample was reverse transcribed using SuperScript IV (Thermo Fisher
Scientific) according to the supplier’s protocol. The protocol developed by ARTIC Network was used to generate and sequence am-
plicon library that covers whole SARS-CoV-2 genome on Oxford Nanopore using ligation sequencing kit (SQK-LSK109) (https://artic.
network/ncov-2019)(Grubaugh et al., 2019;Tyson et al., 2020). Briefly, two multiplex PCR reactions were performed with primer
pools from ARTIC nCoV-2019 V3 Panel (IDT; Table S3) using Q5 High-Fidelity DNA Polymerase (New England Biolabs). Reactions
were performed with the following thermocycling conditions: 98C for 2min, 30 cycles of 98C for 15 s and 65C for 5 min. Two re-
sulting amplicon pools for each patient sample were combined and used for library preparation. After end repair (NEB E7546) sam-
ples were barcoded using Native Barcoding Expansion Kits EXP-NBD104 and EXP-NBD114 from Oxford Nanopore. A total of 24
barcoded samples were pooled together and Nanopore adaptors were ligated. After clean-up 20 ng of multiplexed library DNA
was loaded onto the MinION flowcell for sequencing. A total of 0.3 Gb of raw sequencing data was collected per patient sample.
Nuclease-free water was used as a negative control for library preparation and sequencing. Non-specific amplification in negative
control was additionally checked using an agarose gel. SARS-CoV-2 Isolate USA-WA1/2020 (BEI, NR-52281) was used as a positive
control for genome sequencing.
SARS-CoV-2 genome assembly
MinKNOW software was used to basecall raw Nanopore reads in high-accuracy mode. ARTIC bioinformatic pipeline for COVID-19
was used to analyze reads (https://artic.network/ncov-2019). Pipeline included demultiplexing with guppy barcoder and generating
consensus sequences with minimap2 and calling single nucleotide variants with nanopolish relative to Wuhan-Hu-1/2019 reference
genome (Li, 2018;Quick et al., 2017;Wu et al., 2020). 20X coverage was used as a threshold both for consensus and variant calling.
Nucleotide positions with less than 20X coverage were masked with Ns in the final consensuses. Consensus sequences were up-
loaded to GISAID (https://www.gisaid.org/), accession IDs are provided in Table S1.
RT-PCR and Sanger sequencing
To verify the ORF7aD115 mutation we used RT-PCR with primers that flank the deletion region (Table S3). Reactions were performed
with Q5 High-Fidelity DNA Polymerase in 25 mL volume (New England Biolabs) using following program: 98C for 2min, 35 cycles of
98C for 15 s and 65C for 5 min, 35 cycles. PCR products were analyzed on 1% agarose gels stained with SYBR Safe (Thermo Fisher
Scientific). The remaining volume of the reaction was purified using DNA Clean & Concentrator kit (Zymo Research) and sent to Pso-
magen for Sanger sequencing. Each PCR product was sequenced with both forward and reverse primers used for PCR. To detect
sgRNAs of ORF7a and ORF7b primers annealing to Leader region of SARS-CoV-2 and inside the ORF were used (Table S3). PCR
products were cut out from 1% agarose gel and purified with Zymoclean Gel DNA Recovery Kit. Purified products were sequenced
with forward and reverse primers. Sequence was verified by aligning to SARS-CoV-2 genome in UGENE software (Unipro).
Phylogenetic and ORF7a mutational analysis
An alignment of 181,003 SARS-CoV-2 genomes was downloaded from the GISAID database at 9:11 AM on 2020-11-11, and
sequences without corresponding metadata entries were removed. The nucleotide positions encoding ORF7a in the reference
sequence (Wuhan-Hu-1; EP_ISL_402125) were extracted for the remaining 180,971 sequences, and ORF7a sequences with
100% sequence identity were clustered using CD-HIT (with settings: -c 1 -aL 1). One representative was randomly selected from
each of the 846 ORF7a sequence clusters and mutations in each nucleotide and translated sequence were determined using the
Biostrings package in R. Graphs of these mutations were rendered with the ggplot2 package. To construct a phylogenetic tree of
global and Bozeman SARS-CoV-2 sequences, up to ten sequences from each country in the world were selected for each month
of the pandemic using the Filter utility in the Augur pipeline (–group-by country year month–sequences-per-group 10) (Bedford
et al., 2020;Lemieux et al., 2021). This resulted in an alignment of 4,235 GISAID sequences. Then, the Nextclade online tool was
used to determine the quality of 73 SARS-CoV-2 genomes (% genome covered) sequenced from Bozeman patients. 55 Bozeman
sequences that were determined to be of ‘‘Good’’ quality were merged with the alignment of GISAID sequences, as well as an align-
ment of 669 representative SARS-CoV-2 genomes that had non-synonymous mutations in ORF7a. The bat coronavirus RaTG13
(MN996532.2) and the Wuhan- Hu-1 SARS-CoV-2 sequences were also merged, resulting in an alignment of 4,959 full genome
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 e4
Report
ll
OPEN ACCESS
sequences, and positions that are hypervariable or prone to sequencing artifacts were masked with a VCF file downloaded from
EMBL at 12:45PM on 2020-05- 29. The alignment was trimmed of columns composed of 90% or more gaps using Trimal (-gt
0.9), and IqTree was used to construct a maximum likelihood phylogeny (-m GTR -ninit 2 -n 2 -me 0.05). Branch lengths and internal
nodes were rescaled on the resulting tree using TreeTime, and tree was re-rooted to the RaTG13 sequence using the APE package
before being visualized with the ggTree package in R. Clades were assigned to SARS-CoV-2 genomes using Nextclade and pangolin
(https://cov-lineages.org/) utilities.
Western blot
293T-hACE2 cells were transfected using Lipofectamine 3000 (Invitrogen, L3000-015) with plasmids encoding for ORF7aWT,
ORF7aD115 or control pEGFP-N1 plasmid. To detect ORF7b expression cells were infected with ORF7aWT or ORF7aD115
SARS-CoV-2 strains at MOI = 0.05. After 24 hours, cells were washed two times with PBS and lysed in RIPA buffer (150 mM sodium
chloride, 1% NP-40, 0.5% sodium deoxycholate, 0.2% SDS, 50 mM Tris, pH 8.0) at 4C for 30 min. The lysates were clarified by
centrifugation (10’000 g, 20 min) and stored at 80C. For western blot lysates were mixed with 6xLaemmli SDS-PAGE buffer
and heated at 98C for 5 min, resolved in 12% SDS-PAGE gel and transferred onto a PVDF (ORF7b) or nitrocellulose (ORF7a) mem-
brane using Mini Trans-Blot Electrophoretic Transfer Cell (Bio-Rad, #1703930). Membranes were blocked and probed with indicated
antibodies. The proteins were visualized using Pierce ECL Western Blotting Substrate (ThermoFisher Scientific, #32106) and
exposed to X-Ray film (sc-201696, Santa Cruz Biotech).
Immunocytochemistry
HEK293T-hACE2 were transfected with plasmid DNA using Lipofectamine 3000 (Invitrogen REF: L3000-015) according to manufac-
turer’s instructions. Cells were either transfected with expression vectors for SARS-CoV-2 ORF7aWT, SARS-CoV-2 ORF7aD115, or
mock transfected. 24 hours after transfection, cells were detached and seeded at 50% confluency on glass coverslips (Fisherbrand,
#12-545-80) pre-coated with poly-D-lysine (Cultrex, #3439-100-01) in a 24 well plate. The next day, media was removed, and cells
were washed with PBS and fixed with ice-cold 100% methanol on ice for 10 minutes. Half of the methanol was replaced with PBS
three times before the methanol solution was discarded and cells were washed three times in PBS (5 minutes each wash). Cells were
blocked with 1% BSA (Fisher Scientific BP9703-100) in PBS + 0.1% Tween-20 (PBST) for 30 minutes at room temperature. After
blocking, cells were stained with mouse monoclonal anti-Flag antibody (Invitrogen, MA1-91878) (1:200 in 1% BSA in PBST) overnight
at 4C. The next day, cells were washed 3X in PBS (5 min each) and stained with secondary antibody diluted 1:2000 in 1% BSA in
PBST for 1 h at RT in the dark (Goat anti-mouse AlexaFluor488; Invitrogen, A11001). After 3x PBS washes nuclei were stained with
Hoechst 33342 in PBS for 5 minutes at RT. Coverslips were then mounted with ProLong Gold antifade mounting media (Invitrogen,
P36934) on Superfrost Plus microscope slides (#22-037-246). Immunostained cells were imaged using a Leica SP8 confocal
microscope.
IFN inhibition assay
To generate ISRE reporter system, HEK293T cells were transduced with lentiviruses carrying pGreenFire1-ISRE (Cat. TR016PA-1,
SBI) construct and selected with puromycin (1 ug/ml) for 2 days. After selection cells (1x105 cells/well, 48-well plate) were transfected
with expression plasmids (250 ng) for SARS-CoV-2 ORF7aWT, SARS-CoV-2 ORF7aD115, or control plasmid pLV-mCherry. 16 hours
post transfection, cells were treated for 24 h with 5 ng/ml of human IFN-a2b (Cat. rcyc-hifna2b, InvivoGen) according to manufac-
turer’s recommendations. After treatment activation of IFN signaling was measured using Luciferase Assay System (Cat. E1500,
Promega) on Cytation 5 (BioTek) plate reader. Assay was performed in six replicates for each condition.
SARS-CoV-2 replication assays
HEK293T-hACE2 and Vero E6 cells were seeded in the 48-well plate and infected with ORF7aWT or ORF7aD115 SARS-CoV-2
strains at MOI = 0.05 in 50 mL of the cell culture media. Infections were performed in triplicates. Next, 0.5 h post infection (hpi)
500 mL of the fresh cell culture media was added to each well. The supernatants (140 ul) from the infected cells were harvested at
the following time points 0.5, 6, 24, 48, 72 and 120 hpi. The RNAs from the supernatants were extracted using QIAamp Viral RNA
Mini Kit (QIAGEN) and used as an input in qRT-PCR with CDC N1 and N2 primers and probes as described above. Relative quan-
tification was used (DCt method) to calculate viral replication versus 0.5 hpi time point as 2-DCt.
The infected cells were harvested 0.5, 6 and 24 hpi and washed with PBS. Total RNA from cells was extracted using RNeasy kit
(QIAGEN, Cat No./ID: 74104) with on-column DNase digestion step (QIAGEN, 79254) according to the manufacturers protocol. Viral
RNA was quantified using qRT-PCR CDC N1 primers and probe. Human ACTB endogenous control was quantified with TaqMan
assay (ThermoFisher, 4333762T). DDCt method was used to normalize viral RNA level to host RNA and 0.5 hpi. Viral replication
was calculated for each replicate as 2-DDCt.
Type I interferon response assay
HEK293T-hACE2 were seeded in the 48-well plate and infected with ORF7aWT or ORF7aD115 SARS-CoV-2 strains at MOI = 0.05 in
50 ul of the media. After 30 min on infection the 500 mL of the cell culture media was added to each well. At 24 hpi the cells were
washed once with PBS, detached using Trypsin and the total RNA was extracted using RNeasy kit (QIAGEN, Cat No./ID: 74104)
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
e5 Cell Reports 35, 109197, June 1, 2021
Report
ll
OPEN ACCESS
with on-column DNase digestion (QIAGEN, Cat No./ID: 79254) according to the manufacturer’s protocol. Non-infected cells were
processed using the same protocol with mock infection. After extraction, 0,5-1 mg of total RNA was reverse transcribed Reverse us-
ing SuperScript IV Reverse Transcriptase (Thermo Fisher Scientific) according to the supplier’s protocol. The cDNA was diluted 10-
fold in water and analyzed with Type I interferon response (SAB Target List) H96 qPCR array. Each reaction contained 2 mL of cDNA,
10 mL of 2X SsoAdvanced Universal SYBR Green Supermix (Bio-Rad) and 8 mL of water. Reactions were performed on QuantStudio 3
Real-Time PCR System instrument with following thermocycling conditions: 95C for 2min, 40 cycles of 95C for 5 s and 60C for 30
s. After amplification, melt curve analysis was performed to examine product specificity. Assay was performed in triplicates for non-
infected cells and both viruses. Run data was analyzed in PrimePCR Analysis software (Bio-Rad). Expression of IFN-I response genes
was normalized to geometric mean of 3 reference transcripts (TBP, GAPDH, HPRT1) and non-infected control (DDCt method).
QUANTIFICATION AND STATISTICAL ANALYSIS
All statistical analyses were performed in RStudio v1.2.1335. Moving averages (n = 7) for symptom onset and positive COVID-19 tests
data were calculated using geom_ma function from tidyquant package and plotted with ggplot2 in RStudio. qRT-PCR data is shown
as mean of three biological replicates (each with three technical replicates) ±standard deviation (SD). Data in ISRE reporter assay
was analyzed with one-way ANOVA with Tukey’s post hoc pairwise comparisons. Replication of viruses and expression of IFN-I
response genes was compared using Student’s t test. Medians of IFN-I response between two viruses were compared with Wilcoxon
signed-rank test. Significance levels in figures: * p < 0.05, ** p < 0.01, *** p < 0.001 or ns (no significant difference, p > 0.05).
Please cite this article in press as: Nemudryi et al., SARS-CoV-2 genomic surveillance identifies naturally occurring truncation of ORF7a that limits
immune suppression, Cell Reports (2021), https://doi.org/10.1016/j.celrep.2021.109197
Cell Reports 35, 109197, June 1, 2021 e6
Report
ll
OPEN ACCESS
... Global genomic surveillance during the COVID-19 pandemic has shown that deletion and substitution mutations in the SARS-CoV-2 ORF7a gene are frequent [22][23][24][25]. Though no study has determined the impact of these mutations in the clinical context, certain SARS-CoV-2 ORF7a deletion mutations are known to impair virus replication in vitro [22,26]. ...
... Global genomic surveillance during the COVID-19 pandemic has shown that deletion and substitution mutations in the SARS-CoV-2 ORF7a gene are frequent [22][23][24][25]. Though no study has determined the impact of these mutations in the clinical context, certain SARS-CoV-2 ORF7a deletion mutations are known to impair virus replication in vitro [22,26]. A SARS-CoV-2 strain with a truncated ORF7a (115 nucleotide deletion) was found to be defective in suppressing the host immune response [22]. ...
... Though no study has determined the impact of these mutations in the clinical context, certain SARS-CoV-2 ORF7a deletion mutations are known to impair virus replication in vitro [22,26]. A SARS-CoV-2 strain with a truncated ORF7a (115 nucleotide deletion) was found to be defective in suppressing the host immune response [22]. A mutation (A105V) in the TM domain of SARS-CoV-2 ORF7a that improved its stability was associated with severe disease outcome among a group of Romanian COVID-19 patients [24]. ...
Article
Full-text available
A feature of the SARS-CoV-2 Omicron subvariants BF.5 and BF.7 that recently circulated mainly in China and Japan was the high prevalence of the ORF7a: H47Y mutation, in which the 47th residue of ORF7a has been mutated from a histidine (H) to a tyrosine (Y). Here, we evaluated the effect of this mutation on the three main functions ascribed to the SARS-CoV-2 ORF7a protein. Our findings show that H47Y mutation impairs the ability of SARS-CoV-2 ORF7a to antagonize the type I interferon (IFN-I) response and to downregulate major histocompatibility complex I (MHC-I) cell surface levels, but had no effect in its anti-SERINC5 function. Overall, our results suggest that the H47Y mutation of ORF7a affects important functions of this protein, resulting in changes in virus pathogenesis.
... Mutations in non-Spike ORFs within the 30 kB SARS-CoV-2 genome can impact viral fitness [41], [42], [43], [44]. For example, the commonly occurring ORF7a C-terminal truncation attenuates virusmediated interferon response suppression [42]. ...
... Mutations in non-Spike ORFs within the 30 kB SARS-CoV-2 genome can impact viral fitness [41], [42], [43], [44]. For example, the commonly occurring ORF7a C-terminal truncation attenuates virusmediated interferon response suppression [42]. Non-Spike mutations that appear in the nasal swabs and global variants since mid-2021 [38], [45]. ...
Preprint
Full-text available
SARS-CoV-2 infection of immunocompromised individuals often leads to prolonged detection of viral RNA and infectious virus in nasal specimens, presumably due to the lack of induction of an appropriate adaptive immune response. Mutations identified in virus sequences obtained from persistently infected patients bear signatures of immune evasion and have some overlap with sequences present in variants of concern. We characterized virus isolates from two COVID-19 patients undergoing immunosuppressive cancer therapy, with all isolates obtained greater than 100 days after the initial COVID-19 diagnoses and compared to an isolate from the start of the infection. Isolates from an individual who never mounted an antibody response specific to SARS-CoV-2 despite the administration of convalescent plasma showed slight reductions in plaque size and some showed temperature-dependent replication attenuation on human nasal epithelial cell culture compared to the virus that initiated infection. An isolate from another patient who did mount a SARS-CoV-2 IgM response showed temperature dependent changes in plaque size as well as increased syncytia formation and escape from serum neutralizing antibody. Our results indicate that not all virus isolates from immunocompromised COVID-19 patients display clear signs of phenotypic change, but increased attention should be paid to monitoring virus evolution in this patient population.
... [13][14][15] GRP78, as a molecular chaperone in the ER, plays a critical role in regulating protein folding, assembly, and homeostasis not just for host cell proteins but also for external viral proteins. 13 [63][64][65] Additionally, ORF8 is intricately connected to the evasion of the immune system and the elicitation of inflammatory responses in the context of SARS-CoV-2 infection. [66][67][68] Taken together, GRP78, acting as a host chaperone for viral SARS-CoV-2 proteins, assists in the proper folding, assembly, and matura- ...
Article
Full-text available
The coronavirus disease 2019 (COVID-19) pandemic, driven by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has led to an unprecedented global surge in infections and fatalities. Notably, obesity has emerged as an important susceptibility factor for COVID-19; however, the pathological mechanisms for this remain poorly understood. Recent studies proposed a role for glucose-regulated protein 78 (GRP78), a protein implicated in both obesity and metabolic syndrome, which may function as a binding partner and/or co-receptor for SARS-CoV-2. Given its crucial involvement in diverse biological processes, GRP78 likely plays a major role in multiple facets of the viral life cycle and the pathology of COVID-19. This perspective review discusses the potential contributions of GRP78 to the dynamics of SARS-CoV-2 infection and pathology, particularly in the context of obesity. The primary objective is to facilitate a deeper understanding of the pathogenesis of COVID-19. Through this exploration, we aim to illuminate the complex interactions underpinning the nexus of COVID-19, obesity, and GRP78, ultimately paving the way for informed therapeutic strategies and preventive measures.
... As a result of the former, our molecular phylogenetic analyses suggested that the BA. Previous studies revealed that the deletion of certain accessory genes, such as ORF3a (25)(26)(27), ORF7a (26,28), and ORF8 (26,29), attenuate viral pathogenicity in experimen tally infected animal models. In this study, by creating SARS-CoV-2 mutant lines using reverse genetics, we presented results suggesting that multiple mutations downstream of the S gene in the BA.2 genome cooperatively reduce viral growth efficacy in vitro, thereby attenuating intrinsic pathogenicity. ...
Article
Full-text available
Previous studies on the Omicron BA.2 variant suggested that the virological characteristics of BA.2 are determined by the mutations in at least two different regions of the viral genome: in the BA.2 spike gene (enhancing viral fusogenicity and intrinsic pathogenicity) and the non- spike region of the BA.2 genome (leading to intrinsic pathogenicity attenuation). However, the mutations modulating the BA.2 virological properties remain elusive. In this study, we demonstrated that the L371F substitution in the BA.2 spike protein confers greater fusogenicity and intrinsic pathogenicity. Furthermore, we revealed that multiple mutations downstream of the spike gene in the BA.2 genome are responsible for attenuating intrinsic viral pathogenicity and replication capacity. As mutations in the SARS-CoV-2 variant spike proteins could modulate certain virological properties, such as immune evasion and infectivity, most studies have previously focused on spike protein mutations. Our results underpin the importance of non-spike protein-related mutations in SARS-CoV-2 variants. IMPORTANCE Most studies investigating the characteristics of emerging SARS-CoV-2 variants have been focusing on mutations in the spike proteins that affect viral infectivity, fusogenicity, and pathogenicity. However, few studies have addressed how naturally occurring mutations in the non- spike regions of the SARS-CoV-2 genome impact virological properties. In this study, we proved that multiple SARS-CoV-2 Omicron BA.2 mutations, one in the spike protein and another downstream of the spike gene, orchestrally characterize this variant, shedding light on the importance of Omicron BA.2 mutations out of the spike protein.
... (27,549-27,644 nt) was identified as part of regional genomic sequencing efforts in Bozeman, Montana, USA [17]. This ORF7a deletion (ORF7a Δ 115 ) was subsequently shown to have an in vitro growth defect associated with elevated IFN response suggesting an immunosuppressive role for ORF7a. ...
Preprint
Full-text available
SARS-CoV-2 molecular testing coupled with whole genome sequencing is instrumental for real-time genomic surveillance. Genomic surveillance is critical for monitoring the spread of variants of concern (VOC) as well as novel variant discovery. Since the beginning of the pandemic millions of SARS-CoV-2 genomes have been deposited into public sequence databases. This is the result of efforts of both national and regional diagnostic laboratories. Here we describe the results of SARS-CoV-2 genomic surveillance from February 2021 to June 2022 at a metropolitan hospital in the USA. We demonstrate that consistent daily sampling is sufficient to track the regional prevalence and emergence of VOC. Similar sampling efforts should be considered a viable option for local SARS-CoV-2 genomic surveillance at other regional laboratories.
Article
Full-text available
In vaccine development, many use the spike protein (S protein), which has multiple “spike-like” structures protruding from the spherical structure of the coronavirus, as an antigen. However, there are concerns about its effectiveness and toxicity. When S protein is used in a vaccine, its ability to attack viruses may be weak, and its effectiveness in eliciting immunity will only last for a short period of time. Moreover, it may cause “antibody-dependent immune enhancement”, which can enhance infections. In addition, the three-dimensional (3D) structure of epitopes is essential for functional analysis and structure-based vaccine design. Additionally, during viral infection, large amounts of extracellular vesicles (EVs) are secreted from infected cells, which function as a communication network between cells and coordinate the response to infection. Under conditions where SARS-CoV-2 (severe acute respiratory syndrome coronavirus 2) molecular vaccination produces overwhelming SARS-CoV-2 spike glycoprotein, a significant proportion of the overproduced intracellular spike glycoprotein is transported via EVs. Therefore, it will be important to understand the infection mechanisms of SARA-CoV-2 via EV-dependent and EV-independent uptake into cells and to model the infection processes based on 3D structural features at interaction sites.
Article
Full-text available
Introduction The severe acute respiratory syndrome coronavirus (SARS-CoV-2) spike (S) protein is essential in mediating membrane fusion of the virus with the target cells. Several reports demonstrated that SARS-CoV-2 S protein fusogenicity is reportedly closely associated with the intrinsic pathogenicity of the virus determined using hamster models. However, the association between S protein fusogenicity and other virological parameters remains elusive. Methods In this study, we investigated the virological parameters (e.g., S1/S2 cleavage efficiency, plaque size, pseudoviral infectivity, pseudovirus entry efficiency, and viral replication kinetics) of eleven previous variants of concern (VOCs) and variants of interest (VOIs) correlating with S protein fusogenicity. Results and discussion S protein fusogenicity was found to be strongly correlated with S1/S2 cleavage efficiency and plaque size formed by clinical isolates. However, S protein fusogenicity was less associated with pseudoviral infectivity, pseudovirus entry efficiency, and viral replication kinetics. Taken together, our results suggest that S1/S2 cleavage efficiency and plaque size could be potential indicators to predict the intrinsic pathogenicity and S protein fusogenicity of newly emerged SARS-CoV-2 variants.
Article
Full-text available
The COVID-19 pandemic has highlighted the importance in the understanding of the biology of SARS-CoV-2. After more than two years since the first report of COVID-19, it remains crucial to continue studying how SARS-CoV-2 proteins interact with the host metabolism to cause COVID-19. In this review, we summarize the findings regarding the functions of the 16 non-structural, 6 accessory and 4 structural SARS-CoV-2 proteins. We place less emphasis on the spike protein, which has been the subject of several recent reviews. Furthermore, comprehensive reviews about COVID-19 therapeutic have been also published. Therefore, we do not delve into details on these topics; instead we direct the readers to those other reviews. To avoid confusions with what we know about proteins from other coronaviruses, we exclusively report findings that have been experimentally confirmed in SARS-CoV-2. We have identified host mechanisms that appear to be the primary targets of SARS-CoV-2 proteins, including gene expression and immune response pathways such as ribosome translation, JAK/STAT, RIG-1/MDA5 and NF-kβ pathways. Additionally, we emphasize the multiple functions exhibited by SARS-CoV-2 proteins, along with the limited information available for some of these proteins. Our aim with this review is to assist researchers and contribute to the ongoing comprehension of SARS-CoV-2’s pathogenesis.
Article
In late 2021, the US Centers for Disease Control and Prevention (CDC) posted an interim final rule (86 FR 64075) to the federal register regulating the possession, use, and transfer of SARS-CoV-1/SARS-CoV-2 chimeric viruses. In doing so, the CDC provided the reasoning that viral chimeras combining the transmissibility of SARS-CoV-2 with the pathogenicity and lethality of SARS-CoV-1 pose a significant risk to public health and should thus be placed on the select agents and toxins list. However, 86 FR 64075 lacked clarity in its definitions and scope, some of which the CDC addressed in response to public comments in the final rule, 88 FR 13322, in early 2023. To evaluate these regulatory actions, we reviewed the existing select agent regulations to understand the landscape of chimeric virus regulation. Based on our findings, we first present clear definitions for the terms "chimeric virus," "viral chimera," and "virulence factor" and provide a list of SARS-CoV-1 virulence factors in an effort to aid researchers and federal rulemaking for these agents moving forward. We then provide suggestions for a combination of similarity and functional characteristic cutoffs that the government could use to enable researchers to distinguish between regulated and nonregulated chimeras. Finally, we discuss current select agent regulations and their overlaps with 86 FR 64075 and 88 FR 13322 and make suggestions for how to address chimera concerns within and/or without these regulations. Collectively, we believe that our findings fill important gaps in current federal regulations and provide forward-looking philosophical and practical analysis that can guide future decisionmaking.
Preprint
Full-text available
SARS-CoV-2 initiates infection in the conducting airways, which rely on mucocilliary clearance (MCC) to minimize pathogen penetration. However, it is unclear how MCC impacts SARS-CoV-2 spread after infection is established. To understand viral spread at this site, we performed live imaging of SARS-CoV-2 infected differentiated primary human bronchial epithelium cultures for up to 9 days. Fluorescent markers for cilia and mucus allowed longitudinal monitoring of MCC, ciliary motion, and infection. The number of infected cells peaked at 4 days post-infection in characteristic foci that followed mucus movement. Inhibition of MCC using physical and genetic perturbations limited foci. Later in infection, MCC was diminished despite relatively subtle ciliary function defects. Resumption of MCC and infection spread after mucus removal suggests that mucus secretion mediates this effect. We show that MCC facilitates SARS-CoV-2 spread early in infection while later decreases in MCC inhibit spread, suggesting a complex interplay between SARS-CoV-2 and MCC.
Article
Full-text available
Significance We report that SARS-CoV-2 utilizes its ORF8 protein as a unique mechanism to alter the expression of surface MHC-Ι expression to evade immune surveillance. Our study is significant for providing an understanding of the pathogenesis of SARS-CoV-2 and will provide additional perspective to the intensive ongoing investigation into the mechanism and function of T cell antiviral immunity in COVID-19.
Article
Full-text available
Type I interferons (IFNs) are our first line of defence against virus infection. Recent studies have suggested the ability of SARS-CoV-2 proteins to inhibit IFN responses. Emerging data also suggest that timing and extent of IFN production is associated with manifestation of COVID-19 severity. In spite of progress in understanding how SARS-CoV-2 activates antiviral responses, mechanistic studies into wildtype SARS-CoV-2-mediated induction and inhibition of human type I IFN responses are scarce. Here we demonstrate that SARS-CoV-2 infection induces a type I IFN response in vitro and in moderate cases of COVID-19. In vitro stimulation of type I IFN expression and signaling in human airway epithelial cells is associated with activation of canonical transcriptions factors, and SARS-CoV-2 is unable to inhibit exogenous induction of these responses. Furthermore, we show that physiological levels of IFNα detected in patients with moderate COVID-19 is sufficient to suppress SARS-CoV-2 replication in human airway cells.
Article
Full-text available
The risk posed by Severe Acute Respiratory Syndrome Coronavirus -2 (SARS-CoV-2) dictates that live-virus research is conducted in a biosafety level 3 (BSL3) facility. Working with SARS-CoV-2 at lower biosafety levels can expedite research yet requires the virus to be fully inactivated. In this study, we validated and compared two protocols for inactivating SARS-CoV-2: heat treatment and ultraviolet irradiation. The two methods were optimized to render the virus completely incapable of infection while limiting the destructive effects of inactivation. We observed that 15 min of incubation at 65 °C completely inactivates high titer viral stocks. Complete inactivation was also achieved with minimal amounts of UV power (70,000 µJ/cm2), which is 100-fold less power than comparable studies. Once validated, the two methods were then compared for viral RNA quantification, virion purification, and antibody detection assays. We observed that UV irradiation resulted in a 2-log reduction of detectable genomes compared to heat inactivation. Protein yield following virion enrichment was equivalent for all inactivation conditions, but the quality of resulting viral proteins and virions were differentially impacted depending on inactivation method and time. Here, we outline the strengths and weaknesses of each method so that investigators might choose the one which best meets their research goals.
Article
Full-text available
The established risk factors of coronavirus disease 2019 (COVID-19) are advanced age, male sex and comorbidities, but they do not fully explain the wide spectrum of disease manifestations. Genetic factors implicated in the host antiviral response provide for novel insights into its pathogenesis. We performed an in-depth genetic analysis of chromosome 21 exploiting the genome-wide association study data, including 6,406 individuals hospitalized for COVID-19 and 902,088 controls with European genetic ancestry from the COVID-19 Host Genetics Initiative. We found that five single nucleotide polymorphisms within TMPRSS2 and near MX1 gene show associations with severe COVID-19. The minor alleles of the five SNPs correlated with a reduced risk of developing severe COVID-19 and high level of MX1 expression in blood. Our findings demonstrate that host genetic factors can influence the different clinical presentations of COVID-19 and that MX1 could be a potential therapeutic target.
Article
Full-text available
UK variant transmission Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has the capacity to generate variants with major genomic changes. The UK variant B.1.1.7 (also known as VOC 202012/01) has many mutations that alter virus attachment and entry into human cells. Using a variety of statistical and dynamic modeling approaches, Davies et al. characterized the spread of the B.1.1.7 variant in the United Kingdom. The authors found that the variant is 43 to 90% more transmissible than the predecessor lineage but saw no clear evidence for a change in disease severity, although enhanced transmission will lead to higher incidence and more hospital admissions. Large resurgences of the virus are likely to occur after the easing of control measures, and it may be necessary to greatly accelerate vaccine roll-out to control the epidemic. Science , this issue p. eabg3055
Article
Full-text available
Dysregulated immune cell responses have been linked to the severity of coronavirus disease 2019 (COVID-19), but the specific viral factors of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) were currently unknown. Herein, we reveal that the Ig-like fold ectodomain of the viral protein SARS-CoV-2 ORF7a interacts with high efficiency to CD14⁺ monocytes in human peripheral blood, compared to pathogenic protein SARS-CoV ORF7a. The crystal structure of SARS-CoV-2 ORF7a at 2.2 Å resolution reveals three remarkable changes on the amphipathic side of the four-stranded β-sheet, implying a potential functional interface of the viral protein. Importantly, SARS-CoV-2 ORF7a coincubation with CD14⁺ monocytes ex vivo triggered a decrease in HLA-DR/DP/DQ expression levels and upregulated significant production of proinflammatory cytokines, including IL-6, IL-1β, IL-8, and TNF-α. Our work demonstrates that SARS-CoV-2 ORF7a is an immunomodulating factor for immune cell binding and triggers dramatic inflammatory responses, providing promising therapeutic drug targets for pandemic COVID-19.
Preprint
Full-text available
The SARS-CoV-2 lineage B.1.1.7, now designated Variant of Concern 202012/01 (VOC) by Public Health England, originated in the UK in late Summer to early Autumn 2020. We examine epidemiological evidence for this VOC having a transmission advantage from several perspectives. First, whole genome sequence data collected from community-based diagnostic testing provides an indication of changing prevalence of different genetic variants through time. Phylodynamic modelling additionally indicates that genetic diversity of this lineage has changed in a manner consistent with exponential growth. Second, we find that changes in VOC frequency inferred from genetic data correspond closely to changes inferred by S-gene target failures (SGTF) in community-based diagnostic PCR testing. Third, we examine growth trends in SGTF and non-SGTF case numbers at local area level across England, and show that the VOC has higher transmissibility than non-VOC lineages, even if the VOC has a different latent period or generation time. Available SGTF data indicate a shift in the age composition of reported cases, with a larger share of under 20 year olds among reported VOC than non-VOC cases. Fourth, we assess the association of VOC frequency with independent estimates of the overall SARS-CoV-2 reproduction number through time. Finally, we fit a semi-mechanistic model directly to local VOC and non-VOC case incidence to estimate the reproduction numbers over time for each. There is a consensus among all analyses that the VOC has a substantial transmission advantage, with the estimated difference in reproduction numbers between VOC and non-VOC ranging between 0.4 and 0.7, and the ratio of reproduction numbers varying between 1.4 and 1.8. We note that these estimates of transmission advantage apply to a period where high levels of social distancing were in place in England; extrapolation to other transmission contexts therefore requires caution.
Preprint
Full-text available
The COVID-19 disease, caused by the SARS-Cov-2, presents a heterogeneous clinical spectrum. The risk factors do not fully explain the wide spectrum of disease manifestations, so it is possible that genetic factors could account for novel insights into its pathogenesis. In our previous study, we hypothesized that common variants on chromosome 21, near TMPRSS2 and MX1 genes, may be genetic risk factors associated to the different clinical manifestations of COVID-19. Here, we performed an in-depth genetic analysis of chromosome 21 exploiting the genome-wide association study data including 6,406 individuals hospitalized for COVID-19 and 902,088 controls with European genetic ancestry from COVID-19 Host Genetics Initiative. We found that five single nucleotide polymorphisms (SNPs) within TMPRSS2 and near MX1 gene show suggestive associations (P≤1×10 ⁻⁵ ) with severe COVID-19. All five SNPs replicated the association in two independent cohorts of Asian subjects while two and one out of the 5 SNPs replicated in African and Italian populations, respectively (P≤0.05). The minor alleles of these five SNPs correlated with a reduced risk of developing severe COVID-19 and increased level of MX1 expression in blood. Our findings provide further evidence that host genetic factors can contribute to determine the different clinical presentations of COVID-19 and that MX1, an antiviral effector of type I and III interferon pathway, may be a potential therapeutic target.
Preprint
Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) ORF7a, the ortholog of SARS-CoV ORF7a, is a type I transmembrane protein and plays an important role in virus-host interactions. Deletion variants in ORF7a may influence virulence, but only a few such isolates have been reported. Here, we report 8 unique ORF7a deletion variants of 6 to 96 nucleotides in length identified from 2,726 clinical specimens collected in March of 2020.