ArticlePDF Available

Essential Impact of NF-κB Signaling on the H5N1 Influenza A Virus-Induced Transcriptome

Authors:

Abstract and Figures

Systemic infections of humans and birds with highly pathogenic avian influenza A viruses of the H5N1 subtype are characterized by inner bleedings and a massive overproduction of cytokines known as cytokine storm. Growing evidence supports the role of endothelial cells in these processes. The aim of this study was to elucidate determinants of this strong response in endothelial cells with a focus on the transcription factor NF-kappaB. This factor is known as a major regulator of inflammatory response; however, its role in influenza virus replication and virus-induced immune responses is controversially discussed. By global mRNA profiling of infected cells in the presence or absence of a dominant negative mutant of IkappaB kinase 2 that specifically blocks the pathway, we could show that almost all H5N1 virus-induced genes depend on functional NF-kappaB signaling. In particular, activation of NF-kappaB is a bottleneck for the expression of IFN-beta and thus influences the expression of IFN-dependent genes indirectly in the primary innate immune response against H5N1 influenza virus. Control experiments with a low pathogenic influenza strain revealed a much weaker and less NF-kappaB-dependent host cell response.
Content may be subject to copyright.
Essential Impact of NF-
B Signaling on the H5N1 Influenza A
Virus-Induced Transcriptome
1
Mirco Schmolke,* Dorothee Viemann,
Johannes Roth,
and Stephan Ludwig
2
*
Systemic infections of humans and birds with highly pathogenic avian influenza A viruses of the H5N1 subtype are characterized
by inner bleedings and a massive overproduction of cytokines known as cytokine storm. Growing evidence supports the role of
endothelial cells in these processes. The aim of this study was to elucidate determinants of this strong response in endothelial cells
with a focus on the transcription factor NF-
B. This factor is known as a major regulator of inflammatory response; however, its
role in influenza virus replication and virus-induced immune responses is controversially discussed. By global mRNA profiling of
infected cells in the presence or absence of a dominant negative mutant of I
B kinase 2 that specifically blocks the pathway, we
could show that almost all H5N1 virus-induced genes depend on functional NF-
B signaling. In particular, activation of NF-
B
is a bottleneck for the expression of IFN-
and thus influences the expression of IFN-dependent genes indirectly in the primary
innate immune response against H5N1 influenza virus. Control experiments with a low pathogenic influenza strain revealed a
much weaker and less NF-
B-dependent host cell response. The Journal of Immunology, 2009, 183: 5180 –5189.
Since 1997, when the first cases of directly transmitted
highly pathogenic avian influenza viruses (HPAIV)
3
from
poultry to humans in East Asia were reported (1), a grow-
ing concern arose that these viruses might become the source of a
new influenza pandemic. So far direct transmissions of HPAIV are
rare events (2), need high doses of infectious particles, and did not
lead to any unlimited spread. In poultry, HPAIV of the H5 subtype
show high transmissibility and mortality accompanied by systemic
infection of the host and severe inner bleedings (3). Infections with
HPAIV are characterized by a hyperreaction of the host immune
response accompanied by massive production of cytokines and
chemokines (4, 5). This overproduction of cytokines and chemo-
kines, known as cytokine storm is a cell intrinsic phenomenon (6).
Growing evidence supports the contribution of endothelial cells in
this process. This cell type is not only a major source of cytokines
and chemokines (7) but is also extremely relevant for systemic
viral dissemination, supported by the fact that endotheliotropsim
has been demonstrated for HPAIV in infected birds (8, 9) and
humans (10, 11). In this study, we used primary human endothelial
cells that are highly permissive for and responsive to influenza A
virus infections and show no defects in the innate immune re-
sponse as widely observed for many transformed cell lines.
The innate host cell response is the first line of defense to viral
infections. In the case of influenza virus infections, it is triggered
by accumulation of viral 5-triphosphate vRNA. This pathogen
pattern is mainly sensed by the cytosolic helicase retinoic acid-
induced gene I (RIG-I) (12). Activation of RIG-I leads to phos-
phorylation and nuclear translocation of the constitutively ex-
pressed transcription factor IFN regulatory factor (IRF) 3 via
TANK-binding kinase 1 (13–15) and activation of NF-
B (p50/
p65) via I
B kinase 2 (IKK2)-mediated phosphorylation/degrada-
tion of the inhibitor of
B(I
B
) (16). NF-
B and IRF3 are
known to regulate the expression of many cytokines and chemo-
kines including IFN-
, a major mediator of the innate antiviral
response. Both factors bind to adjacent promoter regions and as
part of the so-called IFN-
enhanceosome (17, 18). IFN-
is se-
creted by the infected cell and can act in an autocrine or paracrine
fashion by binding to the type 1 IFN receptor (IFNAR). Subse-
quently, activation of the JAK/STAT pathway initiates the forma-
tion of the IFN-stimulated gene factor 3 (ISGF3), consisting of
STAT1, STAT2, and IRF9. ISGF3 binds to IFN-sensitive regula-
tory elements (ISRE) and regulates the expression of IFN-stimu-
lated genes (ISG), including IRF7, myxovirus resistance A (MxA),
2–5-oligoadenylate synthetase 1, RIG-I, and IFN-
itself (19 –23),
which can directly or indirectly interfere with the replication of
viruses.
The function of IRF3 and AP-1 in enhancing type I IFN pro-
duction upon influenza A virus infection are undeniable (24).
However, the role of NF-
B in influenza virus infected cells is the
subject of ongoing discussions. Recent reports either claim no in-
volvement of NF-
B in the antiviral gene expression profile in-
duced by influenza viruses (25) or even present data that support
an inhibitory function of NF-
B in this context (26). This is in
contrast to earlier studies that clearly underline the necessity of
initial NF-
B binding to the IFN-
enhanceosome to allow effi-
cient production of IFN-
and thus promote expression of ISGs
(27).
*Institute of Molecular Virology, Center of Molecular Biology of Inflammation and
Interdisciplinary Center of Medical Research and
Institute of Immunology and In-
terdisciplinary Center of Medical Research, Universitaetsklinikum Muenster, Muen-
ster, Germany
Received for publication December 16, 2008. Accepted for publication August
2, 2009.
The costs of publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked advertisement in accordance
with 18 U.S.C. Section 1734 solely to indicate this fact.
1
This work was supported by grants from the Deutsche Forschungsgemeinschaft
(Lu477-11-2, SFB293 A17), the Interdisciplinary Clinical Research Centre of the
University of Muenster (Grant Lud2/032/06), the “FluResearchNet” funded by the
German Ministry of Education and Research, the European Commission funded Spe-
cific Targeted Research Project EUROFLU, and the VIRGIL European Network of
Excellence on Antiviral Drug Resistance.
2
Address correspondence and reprint requests to Dr. Stephan Ludwig, Institute of
Molecular Virology, von Esmarch Strasse 56, 48149 Muenster, Germany. E-mail
address: ludwigs@uni-muenster.de
3
Abbreviations used in this paper: HPAIV, highly pathogenic avian influenza virus;
IRF, IFN regulatory factor; IKK2, I
B kinase 2; ISRE, IFN-sensitive regulatory el-
ement; ISGF3, IFN-stimulated gene factor 3; ISG, IFN-stimulated gene; MxA, myx-
ovirus resistance A; RIG-I, retinoic acid-induced gene I; EGFP, enhanced GFP; MOI,
multiplicity of infection; GO, Gene Ontology; IP-10, IFN-
-inducible protein 10;
NS1, nonstructural protein 1; HMEC-1, human microvascular endothelial cell 1; p.i.,
postinfection.
Copyright © 2009 by The American Association of Immunologists, Inc. 0022-1767/09/$2.00
The Journal of Immunology
www.jimmunol.org/cgi/doi/10.4049/jimmunol.0804198
To dissect the function of NF-
B signaling in the network of
activated pathways triggering the antiviral response in systemic
infections with H5N1 HPAIV, we specifically interfered with
NF-
B signaling by means of expression of a dominant negative
form of IKK2. This approach has been successfully used previ-
ously to efficiently blunt NF-
B activation (28 –31). By global
gene expression profiling, we demonstrate the essential impact of
NF-
B signaling on the antiviral gene response against HPAIV of
the H5N1 subtype.
Materials and Methods
Reagents and plasmids
Recombinant human IFN-
was purchased from PBL. BAY11-7085 was
purchased from Sigma-Aldrich. Cells were preincubated with 5
Mofthe
incubator for 30 min before infection. The retroviral expression plasmids
pCFG5-IEGZ HA and pCFG5-IEGZ IKK2KD were described previously
(28 –31).
Viruses and cells
The HPAIV strain A/Thailand/KAN-1/2004 (H5N1) isolated form a fatal
human case was used with permission from Dr. P. Puthavathana (Bangkok,
Thailand). The low pathogenic human influenza A virus strain A/WSN/33
was taken from the virus collection of the Institute of Molecular Virology
(Muenster, Germany). Viruses were propagated on Madin-Darby canine
kidney (MDCKII) cells. MDCKII were cultured in MEM (PAA) contain-
ing 10% v/v FCS and 100 U/ml penicillin/0.1 mg/ml streptomycin (1
penicillin/streptomycin; Life Technologies). Primary HUVEC (Cambrex)
were cultured as described previously and used in passages five to seven.
Human microvascular endothelial cell 1 (HMEC-1) were cultured in
MCDB-131 containing 10% FCS, 1
g/ml hydrocortisone, 10 ng/ml re-
combinant human epidermal growth factor (R&D Systems), 10 mM L-
glutamine (Life Technologies), and 50 mg/ml gentamicin. Phoenix pack-
aging cells (Orbigen) and HEK293T (American Type Culture Collection)
were cultured in DMEM (PAA) containing 10% v/v FCS and 1
penicillin/streptomycin.
Plaque assay
PFU of a given virus suspension were determined as described earlier (32).
Retroviral gene transfer
Fifteen micrograms of the empty retroviral pCFG5-IEGZ HA vector or
pCFG5-IEGZ IKK2KD expressing dominant negative IKK2 (29) was
transfected into 3 10
6
Phoenix packaging cells with polyethyleneimine
as described previously for HEK293 (33) and selected with 250
g/ml
Zeocin (Invitrogen) for 2 wk to gain stable producer cells. Three 10
6
stable producer cells were seeded on 100-mm dishes 48 h before transduc-
tion. Twenty-four hours before transduction, medium was changed to
HUVEC culture medium (6 ml/dish) and retrovirus-containing superna-
tants were harvested 24 h later. HUVEC and HMEC-1 were infected with
retroviral supernatants as described previously (29). Retrovirally trans-
duced HMEC-1 were selected with 250
g/ml Zeocin (Invitrogen) for 2 wk
to gain stable cell lines.
The efficiency of retroviral gene transfer of HUVEC and HMEC-1 was
measured by flow cytometric detection of recombinant enhanced GFP
(EGFP) that was coexpressed with the gene of interest from a bicistronic
mRNA with a FACSCalibur cytometer (BD Biosciences) 48 h after trans-
duction. Transduction rates ranged from 90 to 100%. Expression of trans-
genes was measured by Western blot.
Western blot and flow cytometry
Cells were lysed in radioimmunoprecipitation assay buffer containing pro-
tease and phosphatase inhibitors (34). Radioimmunoprecipitation assay
protein lysates were mixed with 5Laemmli buffer, separated by SDS-
PAGE, and blotted onto nitrocellulose membranes. Antisera directed
against ERK2, I
B
, and IKK2 were purchased form Santa Cruz Biotech-
nology and Abs against STAT1 and phospho-STAT1 Tyr
701
were obtained
from BD Biosciences. A murine mAb against influenza A virus M1 was
purchased from Serotec. A murine mAb against influenza A virus non-
structural protein 1 (NS1) was generated by V. Wixler in the Institute of
Molecular Virology.
Flow cytometric measurement of intracellular cytokines was performed
as described earlier (28). Mouse mAb against human IL-8 and MCP-1 were
purchased from BD Biosciences. Goat anti-mouse Cy5-labeled secondary
Abs was a gift from V. Wixler (Institute of Molecular Virology). At least
10
4
cells were analyzed in a FACSCalibur cytometer (BD Biosciences)
using CellQuest Pro analysis software.
RNA isolation, cDNA synthesis, and real-time PCR
Total RNA was isolated from 100-mm dishes with a 80 –90% confluency-
grown HUVEC monolayer (1.5–2 10
6
cells) using the RNeasy Kit
(Qiagen) according to the manufacturer’s instruction. One microgram of
total RNA was reverse transcribed with MBI Revert AID Reverse Tran-
scriptase (MBI Fermentas) and oligo(dT) primer according to the manu-
facturer’s instruction. The cDNA was diluted 1/10 and 0.5
l was used for
real-time PCR analysis. Primer sequences are included in the supplemen-
tary material (supplemental Table S1).
4
Real-time PCR was performed
with a Stratagene MX3005P cycler and Brilliant SYBR Green Mastermix
(Stratagene) according to the manufacturer’s instruction. Relative changes
in expression level (n-fold) are calculated according to the 2
⫺⌬⌬CT
method
(35).
DNA microarray and statistical data analysis
Total cellular RNA was isolated from three independent experiments with
wild-type HUVEC that were infected for 5 h with a multiplicity of infec-
tion (MOI) of 5 of the H5N1 virus using a RNeasy kit (Qiagen). Three
additional independent H5N1 infections were performed for microarray
analyses with empty retroviral expression vector- or IKK2KD-expressing
HUVEC. Samples were processed for microarray hybridization using Af-
fymetrix Human Genome 133 Plus 2.0 Gene Arrays according to the man-
ufacturer’s instructions. Fluorescent signals were detected by the GeneChip
Scanner 3000 and recorded and computed by GeneChip Operating Soft-
ware version 1.4 (Affymetrix). Microarray data have been deposited in the
National Center for Biotechnology Information Gene Expression Omnibus
(http://www.ncbi.nlm.nih.gov/geo/Access code: GSE13637).
For a more sophisticated data analysis, we used the Expressionist Suite
software package from GeneData (Basel, Switzerland) as described else-
where (29). Only genes with a fold change of 2.0 or 2.0 and a p0.05
(paired ttest) in multiple independent experiments were considered as
regulated genes. “On/off”-regulated genes were evaluated as described pre-
viously (29). We considered genes with on:off ratios of 0:3, 0:2, 1:3, 3:0,
2:0, and 3:1, respectively. From this group of on/off-regulated genes, we
only included regulations with a high fold change of 5andap0.05 in
the list of regulated genes to differentiate on/off phenomenas occurring
around the background threshold from significant on/off phenomenas. We
applied principal component analyses to reduce mathematically the dimen-
sionality of the entire spectrums of gene expression values of a microarray
experiment to three components (36).
To identify functional categories of genes that are overrepresented in the
data sets of regulated genes, we first assigned Gene Ontology (GO) anno-
tations to every probe set spotted on the Affymetrix 133 Plus 2.0 Array and
compared it with the distribution of GO annotations in the gene group of
interest applying Fisher’s exact test. In the case of genes that are repre-
sented by two or more probe sets, only one transcript was taken into ac-
count to avoid potential bias.
Interaction networks showing direct and indirect relations of given gene
products were generated with the Ingenuity Pathway Analysis 7.1 software
(Ingenuity Systems). The significance of the generated networks was de-
termined by comparison of the interaction of a similar number of random
gene products. The resulting score is a numerical value used to rank net-
works according to their degree of relevance to the network eligible mol-
ecules in your data set. The score takes into account the number of net-
work-eligible molecules in the network and its size, as well as the total
number of network-legible molecules analyzed and the total number of
molecules in Ingenuity’s knowledge base that could potentially be included
in networks. The score is based on the hypergeometric distribution and is
calculated with the right-tailed Fisher’s exact Test. The score is the neg-
ative log of this pvalue. It must be noted that the score is not an indication
of the quality or biological relevance of the network; it simply calculates
the approximate “fit” between each network and your network-eligible
molecules. Networks showing the involvement of a given group of mRNAs
in certain biological processes were generated with Gene Spring GX 10.0
software (Agilent).
Promoter analysis
For the promoter analyses, we took advantage of a computational method
for transcriptional regulatory network inference, CARRIE (Computational
4
The online version of this article contains supplemental material.
5181The Journal of Immunology
Ascertainment of Regulatory Relationships Inferred from Expression) de-
scribed by Haverty et al. (37). Briefly, microarray data and promoter se-
quence data derived from the TRANSFAC database 7.0 are conflated. The
promoter regions of the group of unregulated genes are compared with
the promoters of regulated genes to compute the relative overabundance
of cis-elements in the group of regulated genes. Thereby transcription
factor binding sites are detected which potentially contribute to the
regulation of a given group of genes according to their significance of
overrepresentation.
Results
To gain a global overview on the H5N1 virus-induced gene ex-
pression program in endothelial cells, HUVEC were infected with
a MOI of 5 (MOI 5) to achieve infection rates of 80%, as
detected by immune fluorescence staining of the viral nuclear pro-
tein 8 h after infection (data not shown). HUVEC support efficient
and productive replication of influenza A viruses as shown by
plaque assay of supernatants from infected endothelial cell cultures
in comparison to A549, human lung epithelial cells, which are a
standard cell culture model for influenza A virus infection (sup-
plemental Fig. 1). Before initiation of the microarray analysis, we
tested by real-time RT-PCR at which time point postinfection tran-
scription of mRNAs of prototype genes, such as IFN-
and dif-
ferent ISGs, could be detected. Fig. 1 shows the n-fold mRNA
expression levels normalized to uninfected cells. IFN-
mRNA
expression reaches a maximum after 2 h while prototype ISGs,
such as IFN-
-inducible protein 10 (IP-10), OAS1, or MxA, are
expressed in slightly delayed kinetics as expected (Fig. 1). Proin-
flammatory marker mRNAs like IL-8 or ICAM are induced in a
similar kinetic (data not shown) although to a lesser extent. To
minimize the effects of secondary infections, we decided to use a
time point of 5 h after infection to harvest mRNA, which is well
within the primary replication cycle of influenza viruses. We also
decided to consider only those mRNAs to be regulated that showed
at least a significant ( p0,05) 2-fold change in expression levels
compared with the control group. In former studies, this threshold
guaranteed reliable results and avoided too much false positive
results due to low signal-to-noise ratio (29, 30).
Analysis of the gene expression profile reveals that within 5 h
after infection of HUVEC with the human H5N1 isolate A/Thai-
land/KAN-1/2004, 111 mRNAs were up-regulated and 38 genes
are switched on compared with mock-infected control cells. Inter-
estingly, 600 mRNAs were down-regulated or switched off upon
H5N1 virus infection (data not shown). Although we cannot rule
out a specific down-regulation of individual host mRNAs, it is
more likely that the majority of mRNAs are down-regulated in
consequence of unspecific 5cap snatching (38) or interference
with processing of cellular RNAs by the viral NS1 (39).
To further characterize the genes induced upon infection with
the H5N1 isolate, we grouped these genes into different functional
clusters according to GO classification (www.geneontology.org/).
About one-third of the depicted mRNAs belong to the GO clusters
of inflammatory/immune response and chemotaxis (Fig. 2). This
pattern matches perfectly with recent findings on infections of hu-
mans with HPAIV of the H5N1 subtype, which are accompanied
by a massive systemic production of cytokines and chemokines,
the so-called cytokine storm. Supplemental Table II shows a rep-
resentative subset of mRNAs induced by H5N1 virus infection
(noncharacterized gene products were not listed). Hits were
grouped according to their GO annotated function and ranked ac-
cording to their fold change expression. Consistently with the GO
clustering, a major subset of H5N1-induced mRNAs are described
to be associated either with the antiviral type 1 IFN response (e.g.,
OAS1, MxA) or inflammatory processes (IL-8, ICAM-1). The cy-
tokine storm is a characteristic feature of systemic infections with
highly pathogenic influenza viruses of the H5 subtype. To test
whether the induction of proinflammatory cytokines and ISGs as
well as the NF-
B dependence of this cellular response is limited
to H5N1 isolates, we analyzed HUVEC infected with a low patho-
genic isolate, WSN. Infection of HUVEC with 5 MOI resulted in
moderate gene induction after 5 h compared with H5N1-infected
HUVEC. Induction of IFN-
was 20 times lower after WSN
infection, which consequently results in lower induction of ISGs
like Mx1 and OAS1, as demonstrated by real-time RT-PCR (sup-
plemental Fig. 2). Likewise, production of proinflammatory
cytokines and chemokines or endothelial surface markers was less
pronounced or missing (e.g., IL-6, IL-8, or E-Selectin) in WSN-
infected HUVEC in comparison to H5N1-infected HUVEC.
NF-
B is a crucial factor for expression of antiviral and immune
modulatory cytokines and chemokines that are known to contrib-
ute to the antiviral status of a tissue and are involved in induction
of cell survival/apoptosis (40, 41). To dissect the function of
NF-
B in the primary host gene response, we introduced a dom-
inant negative mutant of the inhibitor of
B kinase 2 (IKK2KD)
into HUVEC by retroviral gene transfer. This tool was already
successfully used for specific and efficient inhibition of stimu-
lus-induced NF-
B signaling in these cells (29, 30). Successful
transduction of the transgene was verified by FACS measure-
ment of coexpressed eGFP that was translated from the same
bicistronic mRNA. Expression of the IKK2KD transgene (Fig.
3A) was further detected in Western blots from total cell lysates
(Fig. 3B). TNF-
-induced I
B
degradation was markedly in-
hibited although not completely abolished in IKK2KD-express-
ing HUVEC (Fig. 3B).
The functional block of NF-
B signaling was verified by im-
paired expression of the NF-
B-dependent gene products IL-8 and
MCP-1 16 h after stimulation with 2 ng/ml TNF-
(Fig. 3C). To
FIGURE 1. Time course of H5N1-induced genes. HUVEC were in-
fected with 5 MOI of H5N1 influenza A virus for the indicated time points.
Expressional changes of mRNAs of IFN-
(), IP-10 (f), ISG15 (Œ),
MxA (), and OAS1 (E) were detected by real-time RT-PCR and are
depicted as mean n-fold (SD) of three independent experiments normal-
ized to controls.
FIGURE 2. Clustering of H5N1-induced genes according to their GO
annotated function. H5N1 influenza A virus-induced mRNAs (up-regulated
and switched on) were grouped according to their predominant GO anno-
tation. Relative distribution of mRNAs (percent) per GO cluster is indi-
cated. GO clusters with 5 mRNAs per cluster are not depicted.
5182 H5N1 VIRUS-INDUCED NF-
B-DEPENDENT TRANSCRIPTOME
exclude impaired virus propagation by NF-
B blockage in early
stages of the replication cycle, we determined the expression of the
viral NS1 and the matrix protein M1 5 h after infection, respec-
tively. Viral protein levels were not altered in infected IKK2KD-
expressing HUVEC (Fig. 3D), which is in accordance with earlier
findings in other cell types (31, 42).
To determine the impact of NF-
B on the overall gene expres-
sion program induced by H5N1 infection, we performed mRNA
array analysis of HUVEC infected with the H5N1 virus (MOI 5)
in the presence or absence of IKK2KD. As an initial comparative
analysis of global gene expression profiles in wild-type, vector-
transduced, or IKK2KD-expressing cells, we performed principal
component analysis. This is a bioinformatic tool for efficient data
reduction without loss of information (43). For that purpose, each
gene profile is transformed to a vector according to the expression
data of each gene in all experimental conditions used. This ap-
proach helps to illustrate the global variance by which different
data sets correlate. The principle components of each individual
infection experiment are depicted as vector clouds in a three-di-
mensional vector space (Fig. 4A). The clouds of each group of
up-regulated mRNAs (e.g., control or H5N1 virus infected) per-
fectly localized together, indicating reliable reproduction of data
within the triplicates. Moreover, a clear separation of the groups
representing H5N1 virus-infected cells vs uninfected as well as
IKK2KD-transduced cells is visible. The induced gene entities of
vector-transduced cells perfectly match with the data from wild-
type control cells. Comparison of vector, control, and IKK2KD-
expressing HUVEC clearly shows that blocking of NF-
B signal-
ing reduces the variance of H5N1-mediated expression changes
compared with the gene profiles of uninfected control. How-
ever, inhibition only partially reverts the influenza A virus-in-
duced gene spectrum (Fig. 4A), suggesting that either all genes
are only partially dependent on NF-
B or that there is a more
variable degree of NF-
B dependence, implying that there
might be a distinct gene subset with no or only a slight require-
ment for the factor. A closer look on the gene patterns revealed
that almost 90% of H5N1-induced mRNAs were NF-
B depen-
dent (Fig. 4B), however, to a variable extent. Among these genes,
FIGURE 3. Block of NF-
B signaling by expression of a dominant negative IKK2 mutant. A, Retroviral transduction of HUVEC: eGFP reporter gene
expression was measured by cytometric analysis. Fluorescence histograms of control cells, vector-transduced cells, and IKK2KD-expressing cells are
depicted. Representative data are shown. B, Expression of IKK2KD was confirmed in Western Blot analysis (upper panel,lanes 5 and 6). TNF-
(2 ng/ml)
induced degradation of I
B
(middle panel,lanes 2,4, and 6) is blocked in the presence of IKK2KD. Equal loading was confirmed by detection of ERK2
(lower panel). C, Expression of IKK2KD blocks TNF-
-induced expression of NF-
B-dependent gene products. HUVEC were stimulated with 2 ng/ml
TNF-
for 16 h in the presence of 25 mM monensin for 16 h. Accumulation of IL-8 and MCP-1 was detected by specific Abs and quantified by FACS
analysis. Gray curves indicate untreated cells and black curves indicate TNF-
-treated cells. D, Western blot analysis of viral M1 and NS1 5 h p.i. in control
cells (lanes 1 and 4), vector-transduced cells (lanes 2 and 5), and IKK2KD-expressing cells (lanes 3 and 6). IKK2KD expression was confirmed with
specific Abs and equal loading was verified with Abs against ERK1.
5183The Journal of Immunology
46% were strictly NF-
B dependent (switched off in the presence
of IKK2KD), while 40% were partially NF-
B dependent. In con-
trast, the expression of IKK2KD has no significant effect on the
genes that are down-regulated or switched off upon influenza virus
infection (data not shown). Strikingly, IKK2KD expression most
severely influenced those mRNAs that were massively induced or
switched on by infection (supplemental Table S2). According to
the microarray data, IFN-
itself belongs to those mRNAs that
strictly need intact NF-
B signaling for their transcription (sup-
plemental Table S2). To validate the microarray data, we per-
formed real-time RT-PCR analysis for a subset of H5N1-induced
mRNAs in the presence or absence of IKK2KD. All tested H5N1-
induced mRNAs are at least partially down-regulated upon expres-
sion of IKK2KD as shown in supplemental Table S1. In parallel,
we used a chemical inhibitor of the NF-
B pathway to validate the
results obtained with the dominant negative IKK2 mutant in
HUVEC. Like IKK2KD-transduced HUVEC, the BAY11-7085-
treated HUVEC show massively impaired production of proin-
flammatory cytokines and chemokines upon H5N1 infection. Sim-
ilar the expression of IFN-
was reduced 5-fold, which
consequently resulted in impaired ISG levels (supplemental Fig.
3). This further underlined the importance of NF-
B for the pri-
mary gene expression response to influenza virus infection. Al-
though HUVEC are a well-established and widely accepted model
for studies of endothelial cell function, we validated our results in
HMEC-1. In accordance with our findings in HUVEC, infection of
HMEC-1 with H5N1 influenza A virus resulted in massive up-
regulation of proinflammatory cytokines and ISGs (supplemental
Fig. S4, Cand D).
Since we proposed a distinct impact of NF-
B signaling in
H5N1 influenza A virus-infected HUVEC, we were curious about
the effects of IKK2KD expression on the gene profile induced by
the low pathogenic control strain. Interestingly, we observed a re-
duced impact of NF-
B blockade on the already weak response to
WSN infection, either by introduction of the dominant negative
IKK2 or by addition of the pharmacological inhibitor BAY 11-
7085. Although IFN-
mRNA induction was reduced by 5-fold in
the presence of IKK2KD after H5N1 infection, we did not observe
a significant reduction in WSN-infected HUVEC. In HMEC-1, we
saw a similarly reduced dependence of the WSN-induced host re-
sponse on NF-
B signaling (supplemental Figs. S2, Aand B, S4,
Cand D, and S5B). A more direct comparison including statistical
analysis of the two cell models and the NF-
B dependence of the
host responses to the high pathogenic and low pathogenic isolate,
respectively, are shown in supplemental Table IV). By Western
blot, we could show that the low pathogenic WSN seems to rep-
licate slower in HUVEC, which might explain the overall reduced
host response (supplemental Fig. S5A). Taken together, these find-
ings point to specific dependence of the H5N1-induced antiviral
response on NF-
B signaling.
The observation that IFN-
is strongly dependent on NF-
Bin
H5N1-infected HUVEC prompted us to ask whether a portion of
the identified NF-
B-regulated genes may be rather indirectly con-
trolled by the factor via induction through the IFN-
signaling
pathway. STAT1 phosphorylation is an immediate hallmark re-
sponse to type I IFN. To test whether the IFN-
or other STAT1-
activating cytokines are actually released as proteins from infected
cells and reduced in IKK2KD-expressing cells, we performed con-
ditioned medium experiments. Wild-type HUVEC or cells that
were vector transduced or IKK2KD transduced were infected for 5
or 8 h, respectively. Supernatants were sterile filtered and trans-
ferred to untreated HUVEC for 15 min. Fig. 5 shows that the
STAT1 phosphorylation at Tyr
701
can be readily observed in
HUVEC after addition of the conditioned supernatants from
H5N1-infected wild-type cells (Fig. 5, lanes 2,4, and 6,upper
panel). Conditioned supernatants from uninfected HUVEC had no
effect on STAT1 phosphorylation (Fig. 5, lanes 1,3, and 5,upper
panel). The specific inhibition of NF-
B signaling clearly results
in reduced levels of phosphorylated STAT1 (Fig. 5, lane 6), which
suggests a reduced production of type I IFNs in these cells. Thus,
the secondary, autocrine type I IFN response is indirectly influ-
enced by blocking NF-
B signaling.
FIGURE 4. Influence of NF-
B block on H5N1 influenza virus-induced mRNAs: A, Principle component analysis of at least three individual experi-
ments per settings. Vector clouds of control HUVEC, vector-transduced HUVEC, H5N1-infected control HUVEC, H5N1-infected vector-transduced
HUVEC, and H5N1-infected IKK2KD-expressing HUVEC are depicted as indicated. Each vector cloud represents the up-regulated/switched on mRNAs
of an independent experiment. B, Relative distribution of H5N1-induced mRNAs (percent) according to the NF-
B dependence determined by microarray
analysis. Strictly NF-
B-dependent mRNAs show expression levels below 2-fold in presence of IKK2KD. A more detailed analysis is listed in supplemental
Table S1.
FIGURE 5. Influence of NF-
B block on H5N1 influenza virus-induced
IFN expression: Western blot analysis of total lysates of HUVEC treated
with conditioned medium from mock-infected control cells (lanes 1,3, and
5) and H5N1-infected cells (5 MOI, 5 h) (lanes 2,4, and 6). Donor cells
were left untreated (lanes 1 and 2), transduced with empty vector (lanes 3
and 4), or pCFG5-IEGZ IKK2KD (lanes 5 and 6). STAT1 Tyr
701
phos-
phorylation was detected 15 min after treatment with conditioned medium
(upper panel). Equal loading was verified by detection of total STAT1
(lower panel).
5184 H5N1 VIRUS-INDUCED NF-
B-DEPENDENT TRANSCRIPTOME
Next, we addressed the question whether the observed effects of
IKK2KD on ISG expression are exclusively due to indirect mech-
anisms via regulating IFN-
or additionally controlled by direct
coregulatory functions of NF-
B in ISGF3- mediated transcrip-
tion. In a time course of infection, transcriptional induction of
IFN-
and ISGs is decreased in IKK2KD-expressing cells at any
given time point (Fig. 6). This was a first indication that an initial
block of IFN-
production is the main NF-
B-controlled event.
To further test whether NF-
B mainly functions as the inducer
of IFN-
in concert with IRF3 or directly influences the expression
of ISGs, we stimulated IKK2KD-expressing endothelial cells with
IFN-
for 3 h. This time point was chosen to mimic the peak
activity of virus-induced IFN-
production. IFN-
-induced ex-
pression of MxA and OAS1 mRNA, as prototype ISGs, are not
influenced in the presence of the dominant negative IKK2 mutant,
which means that the induction of distinct genes generally known
as ISGs via IFNAR1 and JAK/STAT signaling does not necessar-
ily depend on NF-
B (Fig. 7A). This experiment however does not
mimic the overall spectrum of cytokines and chemokines that are
released in parallel to IFN-
during an influenza infection and does
not fully exclude a coregulatory role of NF-
B. To simulate a more
realistic stimulation with the complete set of influenza virus-in-
duced cytokines and chemokines, we again performed conditioned
medium experiments. This time the untreated donor cells were
infected for 3 h and conditioned supernatants of mock-treated or
H5N1 influenza virus-infected cells were transferred to the vector-
or IKK2KD-expressing reporter cells for 2 h. A potential transfer
of infectious viruses was monitored by highly sensitive real-time
PCR (44), which did not detect any viral matrix gene vRNA or
mRNA in the reporter cells after 2 h (data not shown). Three hours
after infection, the donor cells clearly produced IFN-
(Fig. 7B)on
mRNA and protein level as shown by real-time PCR and detection
of phosphorylated STAT1 in a Western blot (Fig. 7C) and indi-
cated by the synthesis of ISGs, such as MxA and IP-10 (Fig. 7B).
Interestingly, the reporter cells did not show a drastic increase in
IFN-
mRNA levels (mean n-fold below 2-fold) although ISGs
were readily transcribed. ISGs like Mx1 or OAS1 were induced by
FIGURE 6. Time course of H5N1-induced mRNAs in the presence of
IKK2KD. Induction of (A) IFN-
(and ), (B) MxA (and ), and
ISG15 (Œand ) mRNAs were measured after H5N1 infection in a time-
dependent manner by highly specific real-time RT-PCR. Filled symbols/
constant lines indicate vector-transduced cells and open symbols/dashed
lines indicate IKK2KD-expressing cells. Mean n-fold expression (SD) of
three independent experiments normalized to controls is depicted.
FIGURE 7. Influence of NF-
B block on IFN-
and H5N1-conditioned medium-mediated signaling. A, Vector-transduced HUVEC (columns 1 and 2)
and IKK2KD-expressing HUVEC (columns 3 and 4) were treated for 3 h with 100 U/ml recombinant human IFN-
(f) or left untreated (). Expression
of MxA and OAS1 mRNA was analyzed by real-time RT-PCR. Mean n-fold expression (SD) of three independent experiments normalized to controls
are depicted. B, Donor cells for production of conditioned medium were mock infected () or infected with 5 MOI of H5N1 for3h(f). Levels of IFN-
,
MxA, and IP-10 mRNA were detected by real-time PCR. Mean n-fold expression (SD) of three independent experiments normalized to controls is
depicted. C, STAT1 Tyr
701
phosphorylation was detected from total lysates of mock-infected donor cells (lane 1,upper panel) or H5N1-infected donor cells
(lane 2, upper panel). Equal loading was verified by detection of total STAT1 (lower panel). D, HUVEC were transduced with empty vector (columns 1
and 2of each panel) or pCFG5-IEGZ IKK2KD (columns 3 and 4of each panel) and treated with conditioned medium of mock-infected cells ()or
H5N1-infected cells (f) for 2 h. Levels of IFN-
, MxA, and OAS mRNAs were detected by real-time RT-PCR. Mean n-fold expression (SD) of three
independent experiments normalized to controls is depicted.
5185The Journal of Immunology
H5N1 virus-conditioned medium (Fig. 7D). However, this induc-
tion was much weaker compared with ISG induction 5 h postin-
fection (p.i.) with influenza A virus. After transfer of H5N1 virus-
conditioned medium, the block of NF-
B signaling did not
significantly alter the expression of OAS1 and Mx1. This under-
lines that NF-
B has no coregulatory function in induced expres-
sion of these prototype ISGs by IFN-
or other STAT-activating
cytokines.
Pathway analysis of the complete H5N1-induced gene profile
(supplemental Fig. 6) using Ingenuity Pathway Analysis 7.1 soft-
ware generated significant interaction networks with pvalues up to
10
59
(only the most significant network is shown). It strength-
ened our initial findings that NF-
B and IFN-
harbor central po-
sitions in the H5N1 influenza virus-induced gene profile. Next, we
wanted to define biological processes that might be differentially
controlled by strictly or partially NF-
B-dependent gene products
induced by H5N1 influenza virus. As shown in supplemental Fig.
7, the strictly NF-
B subset of mRNAs belongs to signaling path-
ways that open up to biological processes like IFN type I produc-
tion (IFN-
and IFN-
) and activation of leukocytes and mono-
cytes. In contrast, the remaining partially dependent mRNAs
control biological processes like T cell activation. However, there
is some overlap of biological functions belonging to the innate
antiviral response in which both groups participate (supplemental
Fig. 7). This is not surprising, since the initial antiviral response is
a highly cross-linked process, with several self-enhancing signal-
ing loops (45, 46).
Finally, we asked which transcription factors are additionally
involved in regulation of H5N1-induced genes that were only par-
tially blocked by IKK2KD expression. The promoter regions of
up-regulated genes were examined by CARRIE analysis as de-
scribed in Materials and Methods. Supplemental Table S3 depicts
only those transcription factor binding sites which were found in at
least four of six analyses with independent control groups of pro-
moter regions of unregulated genes. Strikingly, the transcription
factors that regulate type I IFN gene expression and that are sub-
sequently involved in expression of ISGs (IRF1, IRF7, IRF8 (also
known as ICSBP) and IRF9 (component of the ISRE-binding com-
plex)), NF-
B, and to a lesser extent AP-1 ( p0.03) are over-
represented in this set. It should be noted that IRF3 itself is con-
sidered within the group of the general IRF binding site, because
the consensus site is overlapping to that of other IRF members.
Although this in silico approach does not clarify which of these
binding sites are functionally active, it provides a good overview
of factors that may principally be involved.
The expression of several influenza virus-induced mRNAs was
only partially blocked by expression of IKK2KD. This gave rise to
the question which other factors besides NF-
B are relevant for the
H5N1 virus-induced gene spectrum. AP-1, IRF3, and IRF7 also
have been shown to be components of the IFN-
enhanceosome
(18). IRF3 is constitutively present in the cytosol and migrates to
the nucleus upon TBK/IKK-dependent activation (15) while
IRF7 is induced in response to IFN1R activation via the JAK/
STAT pathway (47). Computational analysis of the promoter re-
gions of partially NF-
B-dependent genes unraveled that binding
sites for members of the IRF family, namely, IRF1, 7, 8 (ICSBP),
9 (enclosed in the ISRE element), and a general IRF binding site
covering IRF3 are highly overrepresented (supplemental Table
S2). This indicates that IRFs may drive the residual virus-induced
gene expression observed upon blockade of NF-
B activity.
Discussion
Systemic infections with HPAIV of the H5N1 subtype are asso-
ciated with a massive production of cytokines and chemokines.
Endothelial cells have been shown to contribute to this hyperacti-
vation of the immune system in human cases (11). It is still un-
certain whether this so-called cytokine storm contributes to the
pathogenesis of the virus (48) and so far it remains unsolved why
H5N1 influenza A viruses allow the production of such a broad
spectrum of antiviral-acting ISGs despite their efficient antagoniz-
ing strategies. Moreover, it is still an open question which signal-
ing pathways are involved in the regulation of the H5N1 influenza
virus-induced overexpression of cytokine and chemokine genes.
By global mRNA profiling, we could show here that infection of
primary endothelial cells with highly pathogenic influenza A vi-
ruses of the H5N1 subtype has a severe impact on the transcrip-
tome of the host cell. This is to our knowledge the first report on
H5N1-induced gene expression in a human primary cell culture
model. Moreover, it is the first study of highly pathogenic influ-
enza A virus-infected endothelial cells. A recent study could show
virus replication with a H3N2 influenza A virus isolate in HUVEC
(49). Since endothelial cells line all blood vessels in mammalian
organisms and are a major source of cytokines and chemokines, it
is highly likely that these cells are involved in a systemic infection
with HPAIV and contribute to the global host response. Case re-
ports from human and animal infections with HPAIV strengthen
this hypothesis (11, 50). These cells replicate the virus efficiently
in cell culture and show a fast and profound host response upon
infection. About 10 times more mRNAs were down than up-reg-
ulated. It is highly likely that the massive down-regulation of
mRNAs is at least partially initiated by 5cap snatching and sub-
sequent degradation of cellular mRNAs (38). Although further
studies have to be performed to examine whether there is a specific
down-regulation of antiviral-acting genes by H5N1 viruses, in this
study we exclusively focused on the up-regulated mRNAs. The
induced profile indeed shows a broad overlap to mRNA profiles
from IFN-
-stimulated HUVEC (51). However, there are certain
published ISGs which were not induced upon IFN type I stimula-
tion in the study by Indraccolo et al., (51) but are expressed in the
case of influenza A virus infection (e.g., IP-10, IP9, Mig, RAN-
TES). A recent study underlined that certain genes, classified as
ISGs, are expressed independently of, although often in parallel to
type I IFNs (52).
In comparison to a previously performed transcriptional pro-
file of an influenza A virus-infected lung epithelial cell line
(53), we detect a much broader spectrum of induced mRNAs in
endothelial cells including a broad spectrum of cytokines and
chemokines. Interestingly, Geiss et al (53) did not detect IFN-
mRNA itself among the up-regulated mRNAs at 4 or 8 h after
infection. This implicates a cell-type-specific response to influ-
enza virus infection. Infection of HUVEC with a low patho-
genic control isolate (WSN) resulted in a less pronounced an-
tiviral and proinflammatory response.
It is well known that influenza virus infection induces activation
of NF-
B signaling via different mechanisms (54, 55), although
the function of NF-
B activation in the signaling network is dis-
cussed highly controversially. We and others could show that in-
fluenza A virus infection is associated with expression of NF-
B-
dependent gene products such as proinflammatory cytokines in cell
culture (6, 53, 56 –58) and in vivo models (59 61). Work from our
group and others revealed that influenza A virus replication in-
duces and depends on activation of NF-
B (31, 42, 62, 63), most
presumably by subsequent activation of caspases that enhance the
passive release of viral ribonucleoproteins from the nucleus in late
phases of viral replication. During preparation of this manuscript,
Kumar et al. (64) additionally claimed that NF-
B signaling is
important for the synthesis of viral RNA in influenza virus-in-
fected cells. In contrast, studies in NF-
B knockout cells in recent
5186 H5N1 VIRUS-INDUCED NF-
B-DEPENDENT TRANSCRIPTOME
years showed that activation of NF-
B is either obsolete for pro-
duction of ISGs upon virus infection (25) or even negatively reg-
ulates production of ISGs and suppresses viral replication (26). In
contrast, the importance of NF-
B for the formation of the IFN-
enhanceosome is undeniable (24). These contradictory findings
prompted us to examine which role NF-
B signaling plays in in-
fluenza A virus-induced cellular gene expression responses, with a
focus on the induction by HPAIV of the H5N1 subtype, that are
characterized by massive induction of cytokines and chemokines
in vivo.
To block NF-
B signaling, we used a dominant negative mutant
of IKK2. This approach has been successfully used previously to
efficiently block NF-
B signaling for gene expression analysis
(28 –30). Selective inhibition of NF-
B signaling by the dominant
negative mutant IKK2KD reverts at least partially the H5N1 virus-
induced gene expression profile. The expression of 46% of the
H5N1-induced mRNAs was completely blocked in the presence of
the dominant negative IKK2 mutant, among them IFN-
mRNA
itself. Among all microarray studies performed so far in this field,
to our knowledge our analysis shows for the first time directly the
obligatory dependence of influenza A virus-induced IFN-
pro-
duction on functional NF-
B signaling. Approximately 43% of
H5N1-induced mRNAs were partially regulated by NF-
B.
Surprisingly, the antiviral host response induced by the low
pathogenic influenza isolate WSN did not show a similar depen-
dence on NF-
B. This points to a distinct function of NF-
B sig-
naling in the antiviral response against highly pathogenic influenza
strains. Ongoing studies address the question if this could be a
molecular reason for the observed cytokine storm in vivo.
In case of the H5N1-induced gene profile, almost all genes that
were previously described to be ISGs were found among NF-
B-
regulated genes. This suggests, that the inhibitory function of
IKK2KD on the H5N1 virus-induced gene profile is to a large
extent mediated indirectly via reduced expression levels of IFN-
.
By analyzing the nature of the regulatory binding sites within the
promoter regions of virus-induced genes, we could show that
NF-
B is less significantly overrepresented than different IRFs. In
combination with the drastic effect of IKK2KD expression exhib-
ited on the overall gene expression profile, it is highly likely that
NF-
B is essential for the direct regulation of a few crucial sig-
naling molecules regulating the antiviral gene program. We could
strengthen this hypothesis by analyzing the time course of IFN-
and ISG induction in the presence of IKK2KD. At any time point
of the kinetic up to8hofinfection, the H5N1 virus-induced pro-
duction of IFN-
was reduced in cells expressing the dominant
negative IKK2. In consequence, the production of ISGs was de-
layed in a similar fashion. Moreover, stimulation of IKK2KD-ex-
pressing HUVEC with IFN-
or treatment with virus-conditioned
medium did not reveal a direct NF-
B dependence in ISG expres-
sion. Thus, functional NF-
B signaling in HUVEC is a critical
bottleneck for an efficient antiviral response against H5N1 influ-
enza viruses and indirectly affects ISG expression by initially con-
trolling expression of IFN-
. This most likely occurs by enhancing
the accessibility of the IFN-
promoter for IRF3 and AP-1 as the
initial step for formation of the enhanceosome and subsequent
IFN-
transcription (65). In support of that hypothesis, we could
demonstrate by conditioned medium transfer experiments that the
initial gene products induced by virus infection are capable of
mediating an IKK2-independent ISG response. Furthermore, re-
lease of the factors appeared to be time dependent and to boost the
antiviral response in cells infected for 5 h, in comparison to cells
only treated for 2 h with conditioned medium (harvested 3 h p.i.
from infected cells).
In contrast to our findings, studies in cells of NF-
B-deficient
mice revealed an inhibitory function of NF-
B in IFN-mediated
signaling (26). This would implicate that ISGs are expressed to a
higher extent in IKK2KD-expressing cells compared with vector-
transduced cells upon stimulation. With our approach, we could
neither confirm this for IFN-
-induced ISGs nor for influenza A
virus-induced ISGs. One reason for this discrepancy might be that
there are compensatory effects in the knockout situation of mouse
cells. In knockout cells, NF-
B signaling is completely blunted,
while in our approach it is not. We manipulated the NF-
B path-
way by introduction of a dominant negative isoform of IKK2,
which allows residual signaling via the remaining endogenous
IKK2 to a certain extent.
In contrast to NF-
B, IRFs were shown to be essential and suf-
ficient for activation of the IFN-
promoter. Previous studies by
Grandvoux et al. (66) used a constitutively active IRF3 mutant
(IRF3–5D) to induce certain ISGs, such as OAS1 or ISG15. It
would be of great interest to see to what extent the coexpression of
a dominant negative IKK2 mutant reverts the effects. Our results
would imply that indeed a basal activity of NF-
B is essential for
IFN-
expression and consequently for the IRF3–5D-induced gene
profile. This cooperation of NF-
B and IRF3 was described in a
virus-independent model by Schafer et al. (67).
Recently, Lee et al. (6) showed in a macrophage model that p38
MAPK signaling is of major importance for the hypercytokinemia
in H5N1 influenza virus-infected individuals (6). The p38 MAPK
signaling pathway triggers a broad spectrum of transcription fac-
tors, among them activating transcription factor 2. Together with
c-Jun, this transcription factor forms the AP-1 that binds to the
IFN-
enhanceosome. By computational analysis, we found that
AP-1 binding sites in H5N1-induced genes are much less promi-
nent compared with the representation of IRF or NF-
B binding
sites within the total spectrum of H5N1-activated genes. More-
over, the presence of AP-1 binding sites is enhanced in partially
NF-
B-dependent mRNAs, indicating an NF-
B-independent reg-
ulatory function of AP-1.
Experimental approaches have to reveal whether p38 MAPK
signaling like NF-
B is a bottleneck for the expression of essential
regulatory molecules like AP-1 in the primary innate immune re-
sponse against highly pathogenic influenza viruses, which would
mean that NF-
B and AP-1 are essential and equivalent partners in
backing up the IRF-mediated production of IFN-
.
In conclusion we could demonstrate here by pathway specific
transcriptome analysis that the gene profile induced by highly
pathogenic influenza A viruses largely depends on functional
NF-
B signaling. Future studies will have to reveal whether this
dependence on NF-
B signaling is the molecular reason for the
massive production of cytokines and chemokines in vivo.
Acknowledgments
We thank Ulla Nordhues for excellent technical assistance in sample prep-
aration for the microarray experiments and Edith Willscher from the Inte-
grated Functional Genomic (Muenster, Germany) for the bioinformatics
support and Desiree Spiering for critically reading this manuscript.
Disclosures
The authors have no financial conflict of interest.
References
1. Pollack, C. V., Jr., C. W. Kam, and Y. K. Mak. 1998. Update: isolation of avian
influenza A(H5N1) viruses from human beings–Hong Kong, 1997–1998. Ann.
Emerg. Med. 31: 647– 649.
2. Uyeki, T. M. 2008. Global epidemiology of human infections with highly patho-
genic avian influenza A (H5N1) viruses. Respirology 13(Suppl. 1): S2–S9.
3. Hsieh, Y. C., T. Z. Wu, D. P. Liu, P. L. Shao, L. Y. Chang, C. Y. Lu, C. Y. Lee,
F. Y. Huang, and L. M. Huang. 2006. Influenza pandemics: past, present and
future. J. Formos. Med. Assoc. 105: 1– 6.
5187The Journal of Immunology
4. Cheung, C. Y., L. L. Poon, A. S. Lau, W. Luk, Y. L. Lau, K. F. Shortridge,
S. Gordon, Y. Guan, and J. S. Peiris. 2002. Induction of proinflammatory cyto-
kines in human macrophages by influenza A (H5N1) viruses: a mechanism for the
unusual severity of human disease? Lancet 360: 1831–1837.
5. Tran, T. H., T. L. Nguyen, T. D. Nguyen, T. S. Luong, P. M. Pham,
V. C. Nguyen, T. S. Pham, C. D. Vo, T. Q. Le, T. T. Ngo, et al. 2004. Avian
influenza A (H5N1) in 10 patients in Vietnam. N. Engl. J. Med. 350: 1179 –1188.
6. Lee, D. C., C. Y. Cheung, A. H. Law, C. K. Mok, M. Peiris, and A. S. Lau. 2005.
p38 mitogen-activated protein kinase-dependent hyperinduction of tumor necro-
sis factor
expression in response to avian influenza virus H5N1. J. Virol. 79:
10147–10154.
7. Danese, S., E. Dejana, and C. Fiocchi. 2007. Immune regulation by microvascular
endothelial cells: directing innate and adaptive immunity, coagulation, and in-
flammation. J. Immunol. 178: 6017– 6022.
8. Feldmann, A., M. K. Schafer, W. Garten, and H. D. Klenk. 2000. Targeted in-
fection of endothelial cells by avian influenza virus A/FPV/Rostock/34 (H7N1) in
chicken embryos. J. Virol. 74: 8018 – 8027.
9. Klenk, H. D. 2005. Infection of the endothelium by influenza viruses. Thromb.
Haemost. 94: 262–265.
10. Claas, E. C., A. D. Osterhaus, R. van Beek, J. C. De Jong, G. F. Rimmelzwaan,
D. A. Senne, S. Krauss, K. F. Shortridge, and R. G. Webster. 1998. Human
influenza A H5N1 virus related to a highly pathogenic avian influenza virus.
Lancet 351: 472– 477.
11. Deng, R., M. Lu, C. Korteweg, Z. Gao, M. A. McNutt, J. Ye, T. Zhang, and J. Gu.
2008. Distinctly different expression of cytokines and chemokines in the lungs of
two H5N1 avian influenza patients. J. Pathol. 216: 328 –336.
12. Pichlmair, A., O. Schulz, C. P. Tan, T. I. Naslund, P. Liljestrom, F. Weber, and
C. Reis e Sousa. 2006. RIG-I-mediated antiviral responses to single-stranded
RNA bearing 5-phosphates. Science 314: 997–1001.
13. Fitzgerald, K. A., S. M. McWhirter, K. L. Faia, D. C. Rowe, E. Latz,
D. T. Golenbock, A. J. Coyle, S. M. Liao, and T. Maniatis. 2003. IKKand
TBK1 are essential components of the IRF3 signaling pathway. Nat. Immunol. 4:
491– 496.
14. Hiscott, J. 2007. Triggering the innate antiviral response through IRF-3 activa-
tion. J. Biol. Chem. 282: 15325–15329.
15. Sharma, S., B. R. tenOever, N. Grandvaux, G. P. Zhou, R. Lin, and J. Hiscott.
2003. Triggering the interferon antiviral response through an IKK-related path-
way. Science 300: 1148 –1151.
16. DiDonato, J. A., M. Hayakawa, D. M. Rothwarf, E. Zandi, and M. Karin. 1997.
A cytokine-responsive I
B kinase that activates the transcription factor NF-
B.
Nature 388: 548 –554.
17. Panne, D., T. Maniatis, and S. C. Harrison. 2007. An atomic model of the inter-
feron-
enhanceosome. Cell 129: 1111–1123.
18. Wathelet, M. G., C. H. Lin, B. S. Parekh, L. V. Ronco, P. M. Howley, and
T. Maniatis. 1998. Virus infection induces the assembly of coordinately activated
transcription factors on the IFN-
enhancer in vivo. Mol. Cell 1: 507–518.
19. Imaizumi, T., M. Hatakeyama, K. Yamashita, A. Ishikawa, H. Yoshida, K. Satoh,
K. Taima, F. Mori, and K. Wakabayashi. 2005. Double-stranded RNA induces
the synthesis of retinoic acid-inducible gene-I in vascular endothelial cells. En-
dothelium 12: 133–137.
20. Levy, D. E., I. Marie, E. Smith, and A. Prakash. 2002. Enhancement and diver-
sification of IFN induction by IRF-7-mediated positive feedback. J. Interferon
Cytokine Res. 22: 87–93.
21. Rebouillat, D., and A. G. Hovanessian. 1999. The human 2,5-oligoadenylate
synthetase family: interferon-induced proteins with unique enzymatic properties.
J. Interferon Cytokine Res. 19: 295–308.
22. Takaoka, A., and H. Yanai. 2006. Interferon signalling network in innate defence.
Cell Microbiol. 8: 907–922.
23. Turan, K., M. Mibayashi, K. Sugiyama, S. Saito, A. Numajiri, and K. Nagata.
2004. Nuclear MxA proteins form a complex with influenza virus NP and inhibit
the transcription of the engineered influenza virus genome. Nucleic Acids Res. 32:
643– 652.
24. Kim, T. K., and T. Maniatis. 1997. The mechanism of transcriptional synergy of
an in vitro assembled interferon-
enhanceosome. Mol. Cell 1: 119 –129.
25. Wang, X., S. Hussain, E. J. Wang, X. Wang, M. O. Li, A. Garcia-Sastre, and
A. A. Beg. 2007. Lack of essential role of NF-
B p50, RelA, and cRel subunits
in virus-induced type 1 IFN expression. J. Immunol. 178: 6770 – 6776.
26. Wei, L., M. R. Sandbulte, P. G. Thomas, R. J. Webby, R. Homayouni, and
L. M. Pfeffer. 2006. NF
B negatively regulates interferon-induced gene expres-
sion and anti-influenza activity. J. Biol. Chem. 281: 11678 –11684.
27. Hiscott, J. 2007. Convergence of the NF-
B and IRF pathways in the regulation
of the innate antiviral response. Cytokine Growth Factor Rev. 18: 483– 490.
28. Denk, A., M. Goebeler, S. Schmid, I. Berberich, O. Ritz, D. Lindemann,
S. Ludwig, and T. Wirth. 2001. Activation of NF-
B via the I
B kinase complex
is both essential and sufficient for proinflammatory gene expression in primary
endothelial cells. J. Biol. Chem. 276: 28451–28458.
29. Viemann, D., M. Goebeler, S. Schmid, K. Klimmek, C. Sorg, S. Ludwig, and
J. Roth. 2004. Transcriptional profiling of IKK2/NF-
B- and p38 MAP kinase-
dependent gene expression in TNF-
-stimulated primary human endothelial
cells. Blood 103: 3365–3373.
30. Viemann, D., M. Schmidt, K. Tenbrock, S. Schmid, V. Muller, K. Klimmek,
S. Ludwig, J. Roth, and M. Goebeler. 2007. The contact allergen nickel triggers
a unique inflammatory and proangiogenic gene expression pattern via activation
of NF-
B and hypoxia-inducible factor-1
.J. Immunol. 178: 3198 –3207.
31. Wurzer, W. J., C. Ehrhardt, S. Pleschka, F. Berberich-Siebelt, T. Wolff,
H. Walczak, O. Planz, and S. Ludwig. 2004. NF-
B-dependent induction of
tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and Fas/FasL is
crucial for efficient influenza virus propagation. J. Biol. Chem. 279:
30931–30937.
32. Olschlager, V., S. Pleschka, T. Fischer, H. J. Rziha, W. Wurzer, L. Stitz,
U. R. Rapp, S. Ludwig, and O. Planz. 2004. Lung-specific expression of active
Raf kinase results in increased mortality of influenza A virus-infected mice. On-
cogene 23: 6639 – 6646.
33. Ehrhardt, C., M. Schmolke, A. Matzke, A. Knoblauch, C. Will, V. Wixler, and
S. Ludwig. 2006. Polyethyleneimine, a cost-effective transfection reagent. Signal
Transduction 6: 179 –184.
34. Ehrhardt, C., T. Wolff, S. Pleschka, O. Planz, W. Beermann, J. G. Bode,
M. Schmolke, and S. Ludwig. 2007. Influenza A virus NS1 protein activates the
PI3K/Akt pathway to mediate antiapoptotic signaling responses. J. Virol. 81:
3058 –3067.
35. Livak, K. J., and T. D. Schmittgen. 2001. Analysis of relative gene expression
data using real-time quantitative PCR and the 2(⫺⌬⌬C
T
) method. Methods 25:
402– 408.
36. Alter, O., P. O. Brown, and D. Botstein. 2000. Singular value decomposition for
genome-wide expression data processing and modeling. Proc. Natl. Acad. Sci.
USA 97: 10101–10106.
37. Haverty, P. M., M. C. Frith, and Z. Weng. 2004. CARRIE web service: auto-
mated transcriptional regulatory network inference and interactive analysis. Nu-
cleic Acids Res. 32: W213–216.
38. Katze, M. G., and R. M. Krug. 1984. Metabolism and expression of RNA poly-
merase II transcripts in influenza virus-infected cells. Mol. Cell. Biol. 4:
2198 –2206.
39. Qiu, Y., and R. M. Krug. 1994. The influenza virus NS1 protein is a poly(A)-
binding protein that inhibits nuclear export of mRNAs containing poly(A). J. Vi-
rol. 68: 2425–2432.
40. Barkett, M., and T. D. Gilmore. 1999. Control of apoptosis by Rel/NF-
B tran-
scription factors. Oncogene 18: 6910 – 6924.
41. DeMeester, S. L., Y. Qiu, T. G. Buchman, R. S. Hotchkiss, K. Dunnigan,
I. E. Karl, and J. P. Cobb. 1998. Nitric oxide inhibits stress-induced endothelial
cell apoptosis. Crit. Care Med. 26: 1500 –1509.
42. Mazur, I., W. J. Wurzer, C. Ehrhardt, S. Pleschka, P. Puthavathana, T. Silberzahn,
T. Wolff, O. Planz, and S. Ludwig. 2007. Acetylsalicylic acid (ASA) blocks
influenza virus propagation via its NF-
B-inhibiting activity. Cell Microbiol. 9:
1683–1694.
43. de Haan, J. R., R. Wehrens, S. Bauerschmidt, E. Piek, R. C. van Schaik, and
L. M. Buydens. 2007. Interpretation of ANOVA models for microarray data
using PCA. Bioinformatics 23: 184 –190.
44. Spackman, E., D. A. Senne, T. J. Myers, L. L. Bulaga, L. P. Garber, M. L. Perdue,
K. Lohman, L. T. Daum, and D. L. Suarez. 2002. Development of a real-time
reverse transcriptase PCR assay for type A influenza virus and the avian H5 and
H7 hemagglutinin subtypes. J. Clin. Microbiol. 40: 3256 –3260.
45. Hu, X., S. D. Chakravarty, and L. B. Ivashkiv. 2008. Regulation of interferon and
Toll-like receptor signaling during macrophage activation by opposing feedfor-
ward and feedback inhibition mechanisms. Immunol. Rev. 226: 41–56.
46. Malmgaard, L. 2004. Induction and regulation of IFNs during viral infections.
J. Interferon Cytokine Res. 24: 439 – 454.
47. Sato, M., N. Hata, M. Asagiri, T. Nakaya, T. Taniguchi, and N. Tanaka. 1998.
Positive feedback regulation of type I IFN genes by the IFN-inducible transcrip-
tion factor IRF-7. FEBS Lett. 441: 106 –110.
48. Salomon, R., E. Hoffmann, and R. G. Webster. 2007. Inhibition of the cytokine
response does not protect against lethal H5N1 influenza infection. Proc. Natl.
Acad. Sci. USA 104: 12479 –12481.
49. Sumikoshi, M., K. Hashimoto, Y. Kawasaki, H. Sakuma, T. Suzutani, H. Suzuki,
and M. Hosoya. 2008. Human influenza virus infection and apoptosis induction
in human vascular endothelial cells. J. Med. Virol. 80: 1072–1078.
50. Rimmelzwaan, G. F., D. van Riel, M. Baars, T. M. Bestebroer, G. van Ameron-
gen, R. A. Fouchier, A. D. Osterhaus, and T. Kuiken. 2006. Influenza A virus
(H5N1) infection in cats causes systemic disease with potential novel routes of
virus spread within and between hosts. Am. J. Pathol. 168: 176 –183, quiz 364.
51. Indraccolo, S., U. Pfeffer, S. Minuzzo, G. Esposito, V. Roni, S. Mandruzzato,
N. Ferrari, L. Anfosso, R. Dell’Eva, D. M. Noonan, et al. 2007. Identification of
genes selectively regulated by IFNs in endothelial cells. J. Immunol. 178:
1122–1135.
52. Melchjorsen, J., H. Kristiansen, R. Christiansen, J. Rintahaka, S. Matikainen,
S. R. Paludan, and R. Hartmann. 2009. Differential regulation of the OASL and
OAS1 genes in response to viral infections. J. Interferon Cytokine Res.
53. Geiss, G. K., M. C. An, R. E. Bumgarner, E. Hammersmark, D. Cunningham, and
M. G. Katze. 2001. Global impact of influenza virus on cellular pathways is
mediated by both replication-dependent and -independent events. J. Virol. 75:
4321– 4331.
54. Flory, E., M. Kunz, C. Scheller, C. Jassoy, R. Stauber, U. R. Rapp, and
S. Ludwig. 2000. Influenza virus-induced NF-
B-dependent gene expression is
mediated by overexpression of viral proteins and involves oxidative radicals and
activation of I
B kinase. J. Biol. Chem. 275: 8307– 8314.
55. Pahl, H. L., and P. A. Baeuerle. 1995. Expression of influenza virus hemagglu-
tinin activates transcription factor NF-
B. J. Virol. 69: 1480 –1484.
56. Chan, M. C., C. Y. Cheung, W. H. Chui, S. W. Tsao, J. M. Nicholls, Y. O. Chan,
R. W. Chan, H. T. Long, L. L. Poon, Y. Guan, and J. S. Peiris. 2005. Proinflam-
matory cytokine responses induced by influenza A (H5N1) viruses in primary
human alveolar and bronchial epithelial cells. Respir. Res. 6: 135.
57. Keeler, C. L., Jr., T. W. Bliss, M. Lavric, and M. N. Maughan. 2007. A functional
genomics approach to the study of avian innate immunity. Cytogenet. Genome
Res. 117: 139 –145.
5188 H5N1 VIRUS-INDUCED NF-
B-DEPENDENT TRANSCRIPTOME
58. Wang, G., J. Zhang, W. Li, G. Xin, Y. Su, Y. Gao, H. Zhang, G. Lin, X. Jiao, and
K. Li. 2008. Apoptosis and proinflammatory cytokine responses of primary
mouse microglia and astrocytes induced by human H1N1 and avian H5N1 in-
fluenza viruses. Cell Mol. Immunol. 5: 113–120.
59. Cameron, C. M., M. J. Cameron, J. F. Bermejo-Martin, L. Ran, L. Xu,
P. V. Turner, R. Ran, A. Danesh, Y. Fang, P. K. Chan, et al. 2008. Gene ex-
pression analysis of host innate immune responses during lethal H5N1 infection
in ferrets. J. Virol. 82: 11308 –11317.
60. Kash, J. C., C. F. Basler, A. Garcia-Sastre, V. Carter, R. Billharz, D. E. Swayne,
R. M. Przygodzki, J. K. Taubenberger, M. G. Katze, and T. M. Tumpey. 2004.
Global host immune response: pathogenesis and transcriptional profiling of type
A influenza viruses expressing the hemagglutinin and neuraminidase genes from
the 1918 pandemic virus. J. Virol. 78: 9499 –9511.
61. Kobasa, D., S. M. Jones, K. Shinya, J. C. Kash, J. Copps, H. Ebihara, Y. Hatta,
J. H. Kim, P. Halfmann, M. Hatta, et al. 2007. Aberrant innate immune response
in lethal infection of macaques with the 1918 influenza virus. Nature 445:
319 –323.
62. Nimmerjahn, F., D. Dudziak, U. Dirmeier, G. Hobom, A. Riedel, M. Schlee,
L. M. Staudt, A. Rosenwald, U. Behrends, G. W. Bornkamm, and J. Mautner.
2004. Active NF-
B signalling is a prerequisite for influenza virus infection.
J. Gen. Virol. 85: 2347–2356.
63. Wurzer, W. J., O. Planz, C. Ehrhardt, M. Giner, T. Silberzahn, S. Pleschka, and
S. Ludwig. 2003. Caspase 3 activation is essential for efficient influenza virus
propagation. EMBO J. 22: 2717–2728.
64. Kumar, N., Z. Xin, Y. Liang, H. Ly, and Y. Liang. 2008. NF-
B signaling
differentially regulates influenza viral RNA synthesis. J. Virol. 82: 9880 –9889.
65. Thanos, D., and T. Maniatis. 1995. Virus induction of human IFN
gene ex-
pression requires the assembly of an enhanceosome. Cell 83: 1091–1100.
66. Grandvaux, N., M. J. Servant, B. tenOever, G. C. Sen, S. Balachandran,
G. N. Barber, R. Lin, and J. Hiscott. 2002. Transcriptional profiling of interferon
regulatory factor 3 target genes: direct involvement in the regulation of interfer-
on-stimulated genes. J. Virol. 76: 5532–5539.
67. Schafer, S. L., R. Lin, P. A. Moore, J. Hiscott, and P. M. Pitha. 1998. Regulation
of type I interferon gene expression by interferon regulatory factor-3. J. Biol.
Chem. 273: 2714 –2720.
5189The Journal of Immunology
... However, NF-κB has been identified as a critical promoter of influenza-related inflammation and efficient viral replication [12]. It has been found that massive inflammation caused by highly pathogenic avian H5N1 viruses is highly dependent on NF-κB [13]. Experiments in vitro and in vivo have proven that inhibition of NF-κB protected against IV-mediated lethal lung injury through inflammation attenuation [14]. ...
... The activation of host cell signaling cascades is reported to be involved in IV replication and excessive inflammatory mediator production [13,22]. We set out to investigate the effects of 5-MF on H1N1 virus-elicited host cell signaling transduction in host cells. ...
... NF-κB and P38 MAPK are critical regulators involved in the regulation of proinflammatory cytokines, which can be activated by viral infection or several cytokines (e.g., TNF-α and IL-1β) [10]. It has been reported that both NF-κB and P38 kinase signaling cascades play pivotal roles in robust cytokine production during low or high pathogenic IV infection [13,21]. Blocking NF-κB or P38 kinase signaling has been found to alleviate excessive proinflammatory cytokine secretion and improve lethal IV-triggered lung injury [37]. ...
Article
Full-text available
Influenza-related acute lung injury (ALI) is a life-threatening condition that results mostly from uncontrolled replication of influenza virus (IV) and severe proinflammatory responses. The methoxy flavonoid compound 5-methoxyflavone (5-MF) is believed to have superior biological activity in the treatment of cancer. However, the effects and underlying mechanism of 5-MF on IV-mediated ALI are still unclear. Here, we showed that 5-MF significantly improved the survival of mice with lethal IV infection and ameliorated IV-mediated lung edema, lung histological changes, and inflammatory cell lung recruitment. We found that 5-MF has antiviral activity against influenza A virus (IAV), which was probably associated with increased expression of radical S -adenosyl methionine domain containing 2 (RSAD2) and suppression of endosomal acidification. Moreover, IV-infected A549 cells with 5-MF treatment markedly reduced proinflammatory mediator expression (IL-6, CXCL8, TNF-α, CXCL10, CCL2, CCL3, CCL4, GM-CSF, COX-2, and PGE 2 ) and prevented P-IKBα, P-P65, and P-P38 activation. Interestingly, we demonstrated that 5-MF treatment could trigger activation of AMP-activated protein kinase (AMPK)α in IV-infected A549 cells, as evidenced by activation of the AMPKα downstream molecule P53. Importantly, the addition of AMPKα blocker compound C dramatically abolished 5-MF-mediated increased levels of RSAD2, the inhibitory effects on H1N1 virus-elicited endosomal acidification, and the suppression expression of proinflammatory mediators (IL-6, TNF-α, CXCL10, COX-2 and PGE 2 ), as well as the inactivation of P-IKBα, P-P65, and P-P38 MAPK signaling pathways. Furthermore, inhibition of AMPKα abrogated the protective effects of 5-MF on H1N1 virus-mediated lung injury and excessive inflammation in vivo. Taken together, these results indicate that 5-MF alleviated IV-mediated ALI and suppressed excessive inflammatory responses through activation of AMPKα signaling.
... There may be other repression mechanisms for the regulation of DNMT1 besides the miRNA pathway. The complex interactions between IAV and the host during infection are often accompanied by changes to signaling pathways, such as the NF-κB/IκB pathway, the MAPK-related signaling pathway, and the PI3K/AKT signaling pathway [35,36]. The PI3K/AKT signaling pathway is involved in the regulation of DNA methylation [37,38]. ...
Article
Full-text available
Influenza A virus (IAV) is a leading cause of human respiratory infections and poses a major public health concern. IAV replication can affect the expression of DNA methyltransferases (DNMTs), and the subsequent changes in DNA methylation regulate gene expression and may lead to abnormal gene transcription and translation, yet the underlying mechanisms of virus-induced epigenetic changes from DNA methylation and its role in virus–host interactions remain elusive. Here in this paper, we showed that DNMT1 expression could be suppressed following the inhibition of miR-142-5p or the PI3K/AKT signaling pathway during IAV infection, resulting in demethylation of the promotor region of the 2′-5′-oligoadenylate synthetase-like (OASL) protein and promotion of its expression in A549 cells. OASL expression enhanced RIG-I-mediated interferon induction and then suppressed replication of IAV. Our study elucidated an innate immunity mechanism by which up-regulation of OASL contributes to host antiviral responses via epigenetic modifications in IAV infection, which could provide important insights into the understanding of viral pathogenesis and host antiviral defense.
... In line with the findings described in Section 3.1 and the identified effectors (see Section 3.4), IL-3, IL-5, and GM-CSF are known to activate the JAK/STAT, Ras/MAPK, and PI3K pathways [102]. The presence of several ILs at the site of inflammation characterizes pathological cytokine and chemokine overproduction, commonly referred to as a "cytokine storm" [93,103]. However, since our data set did not capture the early response and due to the limited gene expression at 1 dpi or 3 dpi, there is currently no conclusive evidence of cytokine overexpression in chickens. ...
Article
Full-text available
Avian influenza is a severe viral infection that has the potential to cause human pandemics. In particular, chickens are susceptible to many highly pathogenic strains of the virus, resulting in significant losses. In contrast, ducks have been reported to exhibit rapid and effective innate immune responses to most avian influenza virus (AIV) infections. To explore the distinct genetic programs that potentially distinguish the susceptibility/resistance of both species to AIV, the investigation of coincident SNPs (coSNPs) and their differing causal effects on gene functions in both species is important to gain novel insight into the varying immune-related responses of chickens and ducks. By conducting a pairwise genome alignment between these species, we identified coSNPs and their respective effect on AIV-related differentially expressed genes (DEGs) in this study. The examination of these genes (e.g., CD74, RUBCN, and SHTN1 for chickens and ABCA3, MAP2K6, and VIPR2 for ducks) reveals their high relevance to AIV. Further analysis of these genes provides promising effector molecules (such as IκBα, STAT1/STAT3, GSK-3β, or p53) and related key signaling pathways (such as NF-κB, JAK/STAT, or Wnt) to elucidate the complex mechanisms of immune responses to AIV infections in both chickens and ducks.
... The observed proviral effects of RIPK4 and CHUK suggest that certain components of NF-kB signaling are beneficial for SARS-CoV-2 infection and may be actively regulated by the virus to promote a viral replicative niche. It has been previously observed that NF-kB pathways can play both proviral and antiviral roles which can be actively regulated and rerouted by other coronaviruses and influenza A (Poppe et al., 2017;Schmolke et al., 2009). ...
Preprint
SARS-CoV-2 can cause a range of symptoms in infected individuals, from mild respiratory illness to acute respiratory distress syndrome. A systematic understanding of the host factors mediating viral infection or restriction is critical to elucidate SARS-CoV-2 host-pathogen interactions and the progression of COVID-19. To this end, we conducted genome-wide CRISPR knockout and activation screens in human lung epithelial cells with endogenous expression of the SARS-CoV-2 entry factors ACE2 and TMPRSS2. These screens uncovered proviral and antiviral host factors across highly interconnected host pathways, including components implicated in clathrin transport, inflammatory signaling, cell cycle regulation, and transcriptional and epigenetic regulation. We further identified mucins, a family of high-molecular weight glycoproteins, as a prominent viral restriction network. We demonstrate that multiple membrane-anchored mucins are critical inhibitors of SARS-CoV-2 entry and are upregulated in response to viral infection. This functional landscape of SARS-CoV-2 host factors provides a physiologically relevant starting point for new host-directed therapeutics and suggests interactions between SARS-CoV-2 and airway mucins of COVID-19 patients as a host defense mechanism.
... On the other hand, the NFκB signaling pathway is responsible for the transcription of pro-inflammatory cytokines, where unbalanced expression can lead to lung damage and a severe disease response [12]. It has been shown that inhibition of NFκB impaired IAV replication and cytokine expression such as IL-8, MCP-1, IL-6, RNATES, and IFN-α/β [7,13,14]. Thus, NFκB could be a potential therapeutic target for IAV infection, as it has much higher potential compared to single target approaches to simultaneously inhibit cascades of pro-inflammatory cytokines and chemokines. ...
Article
Full-text available
Background: We examined associations between NFκB1 polymorphisms and influenza A (H1N1) clinical outcomes in Canadian. Methods: A total of thirty-six Caucasian patients admitted to the intensive care unit (ICU) in hospitals in Canada were recruited during the 2009 H1N1 pandemic. Genomic DNA was extracted from the whole blood samples. The NFkB1 gene was targeted for genotyping using next-generation sequencing technology-Roche 454. Results: A total of 136 single nucleotide polymorphisms (SNPs) were discovered within the NFκB1 gene. Among them, 63 SNPs were significantly enriched in patients admitted in the ICU (p < 0.05) compared with the British Caucasian population in the 1000 Genomes study. These enriched SNPs are mainly intron variants, and only two are exon SNPs from the non-transcribing portion of the NFκB1 gene. Conclusions: Genetic variations in the NFκB1 gene could influence clinical outcomes of pandemic H1N1 infections. Our findings showed that sequence variations of the NFκB1 gene might influence patient response to influenza infection.
... The observed proviral effects of RIPK4 and CHUK suggest that certain components of NF-kB signaling are beneficial for SARS-CoV-2 infection and may be actively regulated by the virus to promote a viral replicative niche. It has been previously observed that NF-kB pathways can have both proviral and antiviral roles, and can be actively regulated and rerouted by other coronaviruses and influenza A 30,40 . ...
Article
Full-text available
Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) causes a range of symptoms in infected individuals, from mild respiratory illness to acute respiratory distress syndrome. A systematic understanding of host factors influencing viral infection is critical to elucidate SARS-CoV-2–host interactions and the progression of Coronavirus disease 2019 (COVID-19). Here, we conducted genome-wide CRISPR knockout and activation screens in human lung epithelial cells with endogenous expression of the SARS-CoV-2 entry factors ACE2 and TMPRSS2. We uncovered proviral and antiviral factors across highly interconnected host pathways, including clathrin transport, inflammatory signaling, cell-cycle regulation, and transcriptional and epigenetic regulation. We further identified mucins, a family of high molecular weight glycoproteins, as a prominent viral restriction network that inhibits SARS-CoV-2 infection in vitro and in murine models. These mucins also inhibit infection of diverse respiratory viruses. This functional landscape of SARS-CoV-2 host factors provides a physiologically relevant starting point for new host-directed therapeutics and highlights airway mucins as a host defense mechanism. Genome-wide CRISPR knockout and activation screens in human lung epithelial cells with endogenous expression of the SARS-CoV-2 entry factors ACE2 and TMPRSS2 identify mucins as key host factors restricting viral infection.
... H5N1 virus infection dramatically enriched NF-κB, which is known as a major regulator of inflammatory response and necroptosis pathway. Our results were consistent with the previous research that low pathogenic influenza strains revealed a much weaker and less NF-κB-dependent host cell response compared with H5N1 virus [28]. However, more than that, necroptosis pathways were newly discovered. ...
Article
Full-text available
Mast cells, widely residing in connective tissues and on mucosal surfaces, play significant roles in battling against influenza A viruses. To gain further insights into the host cellular responses of mouse mast cells with influenza A virus infection, such as the highly pathogenic avian influenza A virus H5N1 and the human pandemic influenza A H1N1, we employed high-throughput RNA sequencing to identify differentially expressed genes (DEGs) and related signaling pathways. Our data revealed that H1N1-infected mouse mast P815 cells presented more up- and down-regulated genes compared with H5N1-infected cells. Gene ontology analysis showed that the up-regulated genes in H1N1 infection were enriched for more degranulation-related cellular component terms and immune recognition-related molecular functions terms, while the up-regulated genes in H5N1 infection were enriched for more immune-response-related biological processes. Network enrichment of the KEGG pathway analysis showed that DEGs in H1N1 infection were specifically enriched for the FoxO and autophagy pathways. In contrast, DEGs in H5N1 infection were specifically enriched for the NF-κB and necroptosis pathways. Interestingly, we found that Nbeal2 could be preferentially activated in H5N1-infected P815 cells, where the level of Nbeal2 increased dramatically but decreased in HIN1-infected P815 cells. Nbeal2 knockdown facilitated inflammatory cytokine release in both H1N1- and H5N1-infected P815 cells and aggravated the apoptosis of pulmonary epithelial cells. In summary, our data described a transcriptomic profile and bioinformatic characterization of H1N-1 or H5N1-infected mast cells and, for the first time, established the crucial role of Nbeal2 during influenza A virus infection.
... In RSV, H1N1 or SARS-CoV-2 infection, clusters of NFκB target genes are expressed, driving the main part of pathological changes in tissues [64,[110][111][112][113][114][115]. Respiratory viral infections result in inflammation and oxidative injury, as well as feedback-mediated enhancement of the expression of inflammatory genes. ...
Article
Full-text available
The bronchial vascular endothelial network plays important roles in pulmonary pathology during respiratory viral infections, including respiratory syncytial virus (RSV), influenza A(H1N1) and importantly SARS-Cov-2. All of these infections can be severe and even lethal in patients with underlying risk factors.A major obstacle in disease prevention is the lack of appropriate efficacious vaccine(s) due to continuous changes in the encoding capacity of the viral genome, exuberant responsiveness of the host immune system and lack of effective antiviral drugs. Current management of these severe respiratory viral infections is limited to supportive clinical care. The primary cause of morbidity and mortality is respiratory failure, partially due to endothelial pulmonary complications, including edema. The latter is induced by the loss of alveolar epithelium integrity and by pathological changes in the endothelial vascular network that regulates blood flow, blood fluidity, exchange of fluids, electrolytes, various macromolecules and responses to signals triggered by oxygenation, and controls trafficking of leukocyte immune cells. This overview outlines the latest understanding of the implications of pulmonary vascular endothelium involvement in respiratory distress syndrome secondary to viral infections. In addition, the roles of infection-induced cytokines, growth factors, and epigenetic reprogramming in endothelial permeability, as well as emerging treatment options to decrease disease burden, are discussed.
... Furthermore, by combining transcription factor analysis and western blotting validation, we identified NF-κB and MAPK signaling pathways as targets of BBD intervention. This is not surprising, as all of these signaling pathways are involved in triggering or controlling inflammatory processes [38,39]. Additionally, based on the results of a recent study by Tan et al. indicating that MyD88-TRIF pathways are coordinated with ERK phosphorylation for excessive pro-inflammatory cytokine production in macrophages [17], it is also likely that BBD abolishes the coordination that occurs among these signaling pathways. ...
Article
Full-text available
It has become evident that the actions of pro-inflammatory cytokines and/or the development of a cytokine storm are responsible for the occurrence of severe COVID-19 during SARS-CoV-2 infection. Although immunomodulatory mechanisms vary among viruses, the activation of multiple TLRs that occurs primarily through the recruitment of adapter proteins such as MyD88 and TRIF contributes to the induction of a cytokine storm. Based on this, controlling the robust production of pro-inflammatory cytokines by macrophages may be applicable as a cellular approach to investigate potential cytokine-targeted therapies against COVID-19. In the current study, we utilized TLR2/MyD88 and TLR3/TRIF co-activated macrophages and evaluated the anti-cytokine storm effect of the traditional Chinese medicine (TCM) formula Babaodan (BBD). An RNA-seq-based transcriptomic approach was used to determine the molecular mode of action. Additionally, we evaluated the anti-inflammatory activity of BBD in vivo using a mouse model of post-viral bacterial infection-induced pneumonia and seven severely ill COVID-19 patients. Our study reveals the protective role of BBD against excessive immune responses in macrophages, where the underlying mechanisms involve the inhibition of the NF-κB and MAPK signaling pathways. In vivo, BBD significantly inhibited the release of IL-6, thus resulting in increased survival rates in mice. Based on limited data, we demonstrated that severely ill COVID-19 patients benefited from BBD treatment due to a reduction in the overproduction of IL-6. In conclusion, our study indicated that BBD controls excessive immune responses and may thus represent a cytokine-targeted agent that could be considered to treating COVID-19.
... Cytokines also attract leukocytes to the endothelium and these activated neutrophils produce neutrophil extracellular traps, which have been shown to have cytotoxic effects on endothelial cells and contribute to lung damage in influenza-infected mice [48]. Influenza can directly infect endothelial cells and activate NF-κB, causing upregulated cytokine and chemokine production and subsequent vascular leakage [49]. It has also been shown that agonists of S1P1, a receptor expressed in pulmonary endothelial cells, suppresses the cytokine storm and decreases mortality [50]. ...
Article
Full-text available
The current pandemic of coronavirus disease 2019 (COVID-19) is caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). While this respiratory virus only causes mild symptoms in younger healthy individuals, elderly people and those with cardiovascular diseases such as systemic hypertension are susceptible to developing severe conditions that can be fatal. SARS-CoV-2 infection is also associated with an increased incidence of cardiovascular diseases such as myocardial injury, acute coronary syndrome, and thromboembolism. Understanding the mechanisms of the effects of this virus on the cardiovascular system should thus help develop therapeutic strategies to reduce the mortality and morbidity associated with SARS-CoV-2 infection. Since this virus causes severe and fatal conditions in older individuals with cardiovascular comorbidities, effective therapies targeting specific populations will likely contribute to ending this pandemic. In this review article, the effects of various viruses—including other coronaviruses, influenza, dengue, and human immunodeficiency virus—on the cardiovascular system are described to help provide molecular mechanisms of pathologies associated with SARS-CoV-2 infection and COVID-19. The goal is to provide mechanistic information from the biology of other viral infections in relation to cardiovascular pathologies for the purpose of developing improved vaccines and therapeutic agents effective in preventing and/or treating the acute and long-term consequences of SARS-CoV-2 and COVID-19.
Article
Full-text available
The genes of the family of interferon (IFN) regulatory factors (IRF) encode DNA binding transcriptional factors that are involved in modulation of transcription of IFN and interferon-induced genes (ISG). The presence of IRF binding sites in the promoter region of IFNA and IFNB genes indicates that IRF factors recognizing these sites play an important role in the virus-mediated induction of these genes. We have described a novel human gene of this family, IRF-3, that is constitutively expressed in a variety of cell types. IRF-3 binds to the interferon-sensitive response element (ISRE) present in the ISG15 gene promoter and activates its transcriptional activity. In the present study, we examined whether IRF-3 can modulate transcriptional activity of IFNA and IFNB promoter regions. Our results demonstrate that IRF-3 can bind to the IRF-like binding sites present in the virus-inducible region of the IFNA4 promoter and to the PRDIII region of the IFNB promoter but cannot alone stimulate their transcriptional activity in the human cell line, 293. However, the fusion protein generated from the IRF-3 binding domain and the RelA(p65) activation domain effectively activates both IFNA4 and IFNB promoters. Cotransfection of IRF-3 and RelA(p65) expression plasmids activates the IFNBgene promoter but not the promoter of IFNA4 gene that does not contain the NF-kB binding site. Surprisingly, activation of the IFNA4 gene promoter by virus and IRF-1 in these cells was inhibited by IRF-3. These data indicate that in 293 cells IRF-3 does not stimulate expression of IFN genes but can cooperate with RelA(p65) to stimulate the IFNB promoter.
Article
Full-text available
Rapid induction of type I interferon (IFN) expression is a central event in the establishment of the innate immune response against viral infection and requires the activation of multiple transcriptional proteins following engagement and signaling through Toll-like receptor-dependent and -independent pathways. The transcription factor interferon regulatory factor-3 (IRF-3) contributes to a first line of defense against viral infection by inducing the production of IFN-beta that in turn amplifies the IFN response and the development of antiviral activity. In murine knock-out models, the absence of IRF-3 and the closely related IRF-7 ablates IFN production and increases viral pathogenesis, thus supporting a pivotal role for IRF-3/IRF-7 in the development of the host antiviral response.
Article
Full-text available
The 2'-5' oligoadenylate synthetase (OAS) family consist of three genes encoding active OAS enzymes (OAS1-3) and an OAS-Like (OASL) gene encoding an inactive protein. The transcription of all four members of this family is actively induced by interferon (IFN), but so far no attempt to systematically analyze the expression of these genes during viral infection has been made. We analyzed the expression of the human OAS1 and OASL genes in response to infection with Sendai virus or Influenza A virus. Surprisingly, we found a marked difference in the expression pattern of these genes. Our data showed that the OASL gene is rapidly induced in response to viral infection and that this induction is mediated by IFN regulatory factor 3 (IRF-3). In contrast to the OASL gene, the induction of the OAS1 gene by virus infection was lower, and did require a functional type I IFN response. The pronounced difference in gene regulation between the OAS1 and OASL genes agrees with a functional difference between these genes, which must exist as a consequence of the lack of the 2-5A synthetase activity of the OASL protein. Furthermore, the behavior of the OASL gene is consistent with the behavior of an antiviral gene.
Article
We present evidence that transcriptional activation of the human interferon-β (IFNβ) gene requires the assembly of a higher order transcription enhancer complex (enhanceosome). This multicomponent complex includes at least three distinct transcription factors and the high mobility group protein HMG I(Y). Both the in vitro assembly and in vivo transcriptional activity of this complex require a precise helical relationship between individual transcription factor-binding sites. In addition, HMG I(Y), which binds specifically to three sites within the enhancer, promotes cooperative binding of transcription factors in vitro and is required for transcriptional synergy between these factors in vivo. Thus, HMG I(Y) plays an essential role in the assembly and function of the IFNβ gene enhanceosome.
Article
A functional interferon-β gene enhanceosome was assembled in vitro using the purified recombinant transcriptional activator proteins ATF2/c-JUN, IRF1, and p50/p65 of NF-κB. Maximal levels of transcriptional synergy between these activators required the specific interactions with the architectural protein HMG I(Y) and the correct helical phasing of the binding sites of these proteins on the DNA helix. Analyses of the in vitro assembled enhanceosome revealed that the transcriptional synergy is due, at least in part, to the cooperative assembly and stability of the complex. Reconstitution experiments showed that the formation of a stable enhanceosome-dependent preinitiation complex requires cooperative interactions between the enhanceosome; the general transcription factors TFIID, TFIIA, and TFIIB; and the cofactor USA. These studies provide a direct biochemical demonstration of the importance of the structure and function of natural multicomponent transcriptional enhancer complexes in gene regulation.
Article
In search for a cheap and effective transfection reagent we used the positively charged polyplex-forming compound polyethylenimine (PEI). This compound is commercially available from different companies either as a non-modified chemical reagent or with additives as a more cost intensive transfection reagent. Here we used the non-modified PEI reagent to optimize transfection protocols for different cell-lines. With these optimized conditions we were able to transiently transfect a number of cell-lines up to 40–90%.
Article
Activated macrophages and their inflammatory products play a key role in innate immunity and in pathogenesis of autoimmune/inflammatory diseases. Macrophage activation needs to be tightly regulated to rapidly mount responses to infectious challenges but to avoid toxicity associated with excessive activation. Rapid and potent macrophage activation is driven by cytokine-mediated feedforward loops, while excessive activation is prevented by feedback inhibition. Here we discuss feedforward mechanisms that augment macrophage responses to Toll-like receptor (TLR) ligands and cytokines that are mediated by signal transducer and activator of transcription 1 (STAT1) and induced by interferon-gamma (IFN-gamma). IFN-gamma also drives full macrophage activation by inactivating feedback inhibitory mechanisms, such as those mediated by interleukin-10 (IL-10), and STAT3. Priming of macrophages with IFN-gamma reprograms cellular responses to other cytokines, such as type I IFNs and IL-10, with a shift toward pro-inflammatory STAT1-dominated responses. Similar but partially distinct priming effects are induced by other cytokines that activate STAT1, including type I IFNs and IL-27. We propose a model whereby opposing feedforward and feedback inhibition loops crossregulate each other to fine tune macrophage activation. In addition, we discuss how dysregulation of the balance between feedforward and feedback inhibitory mechanisms can contribute to the pathogenesis of autoimmune and inflammatory diseases, such as rheumatoid arthritis and systemic lupus erythematosus.
Article
The pathogenesis of human H5N1 influenza remains poorly understood and controversial. 'Cytokine storm' has been hypothesized to be the main cause of the severity of this disease. However, the significance of this hypothesis has been called into question by a recent report, which demonstrates that inhibition of the cytokine response did not protect against lethal H5N1 influenza infection in mice. Here we showed discrepant findings in two adult H5N1 autopsies and a fetus obtained at autopsy which also raise doubt about this hypothesis. Antigens of 10 cytokines/chemokines which were found to be significantly elevated in previous H5N1-infected patients and in vitro experiments, and mRNA of eight of these, were absent from the lungs of a pregnant woman and her fetus. In contrast, antigens of seven cytokines/chemokines and mRNA of six of these were found to be increased in the lungs of a male autopsy. The cells expressing these cytokines and chemokines were identified as type II pneumocytes, bronchial epithelial cells, macrophages and vascular endothelial cells. Levels of cytokines and chemokines in the serum of the male case were also significantly higher than those of infectious (infection other than by H5N1) and non-infectious controls. In comparison with results from our previous study, it appeared that the male case had increased cytokine/chemokine expression but reduced viral load, while the pregnant female had diminished cytokine/chemokine expression but a significantly increased viral load in the lungs. These disparate findings in these two cases suggest that 'cytokine storm' alone could not be a sufficient explanation for the severe lung injury of this newly emerging disease.