ArticlePDF AvailableLiterature Review

Role of p38 mitogen‐activated protein kinase signalling in virus replication and potential for developing broad spectrum antiviral drugs

Authors:

Abstract

Mitogen‐activated protein kinases (MAPKs) play a key role in complex cellular processes such as proliferation, development, differentiation, transformation and apoptosis. Mammals express at least four distinctly regulated groups of MAPKs which include extracellular signal‐related kinases (ERK)‐1/2, p38 proteins, Jun amino‐terminal kinases (JNK1/2/3) and ERK5. p38 MAPK is activated by a wide range of cellular stresses and modulates activity of several downstream kinases and transcription factors which are involved in regulating cytoskeleton remodeling, cell cycle modulation, inflammation, antiviral response and apoptosis. In viral infections, activation of cell signalling pathways is part of the cellular defense mechanism with the basic aim of inducing an antiviral state. However, viruses can exploit enhanced cell signalling activities to support various stages of their replication cycles. Kinase activity can be inhibited by small molecule chemical inhibitors, so one strategy to develop antiviral drugs is to target these cellular signalling pathways. In this review, we provide an overview on the current understanding of various cellular and viral events regulated by the p38 signalling pathway, with a special emphasis on targeting these events for antiviral drug development which might identify candidates with broad spectrum activity.
Received: 29 October 2020
-
Revised: 21 December 2020
-
Accepted: 23 December 2020
DOI: 10.1002/rmv.2217
REVIEW
Role of p38 mitogenactivated protein kinase signalling in
virus replication and potential for developing broad
spectrum antiviral drugs
Yogesh Chander
1,2
|Ram Kumar
1,3
|Nitin Khandelwal
1,4
|Namita Singh
2
|
Brij Nandan Shringi
3
|Sanjay Barua
1
|Naveen Kumar
1
1
National Centre for Veterinary Type Cultures,
ICARNational Research Centre on Equines,
Hisar, Haryana, India
2
Department of Bio and Nano Technology,
Guru Jambeshwar University of Science and
Technology, Hisar, Haryana, India
3
Department of Veterinary Microbiology and
Biotechnology, Rajasthan University of
Veterinary and Animal Sciences, Bikaner, India
4
Department of Biotechnology, GLA
University, Mathura, India
Correspondence
Naveen Kumar, National Centre for
Veterinary Type Cultures, ICARNational
Research Centre on Equines, Hisar, Haryana
125001, India.
Email: naveenkumar.icar@gmail.com
Funding information
Science and Engineering Research Board,
Grant/Award Numbers: CRG/2018/004747,
CRG/2019/000829, CVD/2020/000103
Summary
Mitogenactivated protein kinases (MAPKs) play a key role in complex cellular
processes such as proliferation, development, differentiation, transformation and
apoptosis. Mammals express at least four distinctly regulated groups of MAPKs
which include extracellular signalrelated kinases (ERK)1/2, p38 proteins, Jun
aminoterminal kinases (JNK1/2/3) and ERK5. p38 MAPK is activated by a wide
range of cellular stresses and modulates activity of several downstream kinases and
transcription factors which are involved in regulating cytoskeleton remodeling, cell
cycle modulation, inflammation, antiviral response and apoptosis. In viral infections,
activation of cell signalling pathways is part of the cellular defense mechanism with
the basic aim of inducing an antiviral state. However, viruses can exploit enhanced
cell signalling activities to support various stages of their replication cycles. Kinase
activity can be inhibited by small molecule chemical inhibitors, so one strategy to
develop antiviral drugs is to target these cellular signalling pathways. In this review,
we provide an overview on the current understanding of various cellular and viral
events regulated by the p38 signalling pathway, with a special emphasis on targeting
these events for antiviral drug development which might identify candidates with
broad spectrum activity.
KEYWORDS
MAPK, p38, signalling pathway, virus infection
1
|
INTRODUCTION
As per the report of the International Committee on Taxonomy of
Viruses (ICTV, 2019), 4958 viral species are listed across 14 orders,
143 families and 846 genera.
1
Some emerging viruses are
problematic adding to the list of the major disease outbreaks
that have occurred throughout history. Viral epidemics still
occur and the world is currently facing an outbreak of a new
coronavirus disease (COVID19). Whereas the most successful
approach to control virus infections is to use vaccines, this
strategy is not an option for many infections. A potentially more
effective general approach to control virus infections is to develop
effective antivirals.
2
The majority of antiviral drugs approved
by Food and Drug Administration (FDA), act by directly targeting
viral encoded factors.
3
However, successful directlyacting drugs
may eventually fail due to the emergence of drugresistant
Yogesh Chander and Ram Kumar contributed equally to this study.
Rev Med Virol. 2021;116. wileyonlinelibrary.com/journal/rmv © 2021 John Wiley & Sons Ltd.
-
1
mutants.
3
Therefore, alternative antiviral approach needs to be
explored.
Mammalian cells respond to various biotic and abiotic stresses
through signal transduction pathways. Kinases and phosphatases
play a key role in activation/deactivation of various molecules
involved in the signal transduction pathways.
4
Completion of the
human genome project in the beginning of 21st century identified
518 kinases (the so called kinome).
5
Kinases are implicated in various
physiological processes to maintain homeostasis and, their dysregu-
lation could result in pathology. Infections by pathogens, including
viral infections, are also associated with perturbation of the kinome.
Each step of the virus replication cycle is believed to be regulated by
multiple kinases.
6,7
The family members of mitogenactivated protein kinase (MAPK)
are the key kinases involved in most signal transduction pathways.
8
Mammalian cells express at least four distinctly regulated groups
of MAPKs, p38 proteins, extracellular signalrelated kinases1/2
(ERK1/2), Jun aminoterminal kinases (JNK1/2/3) and ERK5. These
MAPKs are activated by specific MAPKKs (MAPK kinase): MKK3/6
for p38, MEK1/2 for ERK1/2, MEK5 for ERK5 and MKK4/7 (JNKK1/
2) for the JNKs (Figure 1). Each MAPKK can be activated by multiple
MAPKKK (MAPK kinase), thereby increasing the complexity and di-
versity of MAPK signalling. Each MAPKKK is believed to confer
responsiveness to distinct stimuli. The activation of classical MAPK
signalling pathway begins at the host cell plasma membrane where
various cell surface receptors phosphorylate MAPKKKs. The
MAPKKKs then phosphorylate MAPKKs which subsequently activate
MAPKs. The activated MAPKs eventually modulate the transcription
factors that drive contextspecific gene expression. Although, in
Figure 1,the MAPK signalling pathway is depicted as simple linear,
unidirectional groups of protein kinases, in the real sense it is quite
complex wherein a high degree of crosstalk exists between
the MAPK cascade and the other signalling pathways such as nuclear
factorkappa B (NF‐κB), phosphatidylinositol3kinase (PI3K)/pro-
tein kinase B (Akt), and janus kinase/signal transducer and activator
of transcription (JAK–STAT) pathways.
The host cell activates the signalling pathway with the basic aim
of inducing an antiviral state, including production of antiviral cyto-
kines. However, viruses can misuse the activated pathways for
completion of various stages of their replication cycle. Depending on
the nature of interacting proteins, the net outcome of the activation
of p38 signalling pathway may be in favor (proviral) or detrimental to
virus replication. Cellular factors and signalling pathways, which are
critical for completion of various steps in the virus life cycle but
are dispensable for the host, could serve as potential drug targets.
9
Since the genetic variability of the host is quite low as compared to
the viruses, hosttargeting antiviral agents are considered to mini-
mize drug resistance.
10
Certain acute viral infections such as severe
acute respiratory syndrome coronavirus 2 (SARSCoV2),
11
Ebolavi-
rus (EBOV),
12
dengue virus (DENV)
13
H5N1 influenza virus
14
lead to
hyperactivation of the immune response (cytokine storm), mediated
via intracellular cell signalling pathways and eventually resulting in
immunopathology.
15
Therefore, besides restricting virus replication,
inhibiting the activity of cell signalling pathways may also minimize
immunopathology.
15
Additionally, virus replication can be controlled
by potentiating innate immune response which includes administra-
tion of ligand that triggers innate immune receptors and blocking
immunoinhibitory interactions and cytokine therapy.
15
This review
discusses the various upstream regulators of p38, its downstream
substrates and their role in virus replication and pathogenesis,
with the basic aim of targeting these virus–host interaction events
for antiviral drug development.
2
|
p38 SIGNALLING PATHWAY
p38 is usually activated in response to stress therefore, it is also
considered as a stressactivated MAPK.
8
p38 has four splice vari-
ants (isoforms) which include p38α(MAPK14), p38β(MAPK11),
p38γ(MAPK12) and p38δ(MAPK13).
16
All p38 isoforms are known
to share approximately 60% sequence similarity within the p38
MAPK group and 40%–45% with the other MAPK family mem-
bers.
17
p38 signalling can be activated by abiotic and biotic stress
including viral infections (Figure 1). The activation of p38 signalling
is mediated via a variety of cell surface
18,19
as well as intracellular
receptors
20
(Figure 1). In most instances, the activated receptor/
adaptor proteins recruit membrane proximal upstream regulator
MAPKKK which activates p38MAPK, mediated via intermediate
kinases MAPKK. MAPKKs such as MKK3 and MKK6 catalyzes
the phosphorylation of a conserved Thr–Gly–Tyr motif of p38.
21
Additionally, MKK4, an upstream kinase of JNK, can also activate
p38.
22
MAPKKindependent activation of p38 has also been
described in T cells.
21
Likewise, in endothelial cells, Gprotein
coupled receptors (GPCRs) directly recruit transforming growth
factor‐βactivated protein kinase 1binding protein 1 (TAB1) to
induce p38 activation.
23
However, the eventual biological response
of p38 pathway is based on the substrate selectivity of upstream
kinases. Activated p38 further induces activation of several tran-
scription factors/effector proteins (Figure 1) which eventually reg-
ulates DNA replication and repair (mediated via growth arrest and
DNA damageinducible protein GADD45 alpha [GAdd45‐α], heat
shock protein 90 [Hsp90] and casein kinase 2 [CK2]), transcription
(mediated via NF‐κB, interferon regulatory factor 3 [IRF3], IRF7,
activating transcription factor [ATF], cAMPresponse element
binding protein [CREB], cMyc, activator protein 1 [AP1], STAT
and nuclear factor of activated Tcells [NFAT]), and translation
(mediated via MAPK interacting kinase 1 [MNK1]/eukaryotic
translation initiation factor4E [eIF4E], Tcell intracellular antigen
[TIA1], TIA1related protein [TIAR], IRESspecific cellular trans-
acting factors [ITAF], polypyrimidine tractbinding protein [PTB],
eukaryotic initiation factors [eIFs] and far upstream elementbinding
protein 1 [FUBP1]) of more than 100 cytoplasmic and nuclear
substrates. Activated p38 also regulates several other cellular pro-
cesses such as apoptosis (mediated via p53, caspases and CCAAT
enhancerbinding protein homologous protein [CHOP]), cellular
trafficking and cytoskeleton remodelling (mediated via Rabenosyn5
2
-
CHANDER
ET AL.
A[Rab5A] and early endosome antigen 1 [EEA1]; Figure 1). In
addition, p38mediated activation of transcription factors such as
NF‐κB, IRF3, IRF7, ATF, CREB, cMyc, AP1, STAT and NFAT results
in activation of several proinflamamtory cytokine genes (Figure 1).
However, hyperactivation of the signalling pathway(s) may lead to
cytokine storm
15
which has been observed in some acute viral
infections such as SARSCoV2,
11
EBOV,
12
DENV
13,14
and H5N1
14
and leads to tissue damaging immunopathology.
Like other kinases, p38 activity is regulated by phosphorylation
and dephosphorylation.
24
A number of phosphatases are involved in
termination of p38 kinase catalytic activity. The major group includes
the family members of dualspecificity phosphatases.
25
Besides,
FIGURE 1 p38 Signalling pathway: p38 is activated by a wide range of cellular stresses as well as in response to inflammatory cytokines.
Upon activation, these receptors transduce signals to downstream kinases by utilizing number of adaptor and accessory proteins such as
cAMP, DAG, GRB2, SOS, Rac1, Cdc42, FADD, TRADD, RIP, TRAF, MyD88, TIRAP, TRIF, TRAM, IRAK, TRAF6, 1RAcP, MyD88, IRAK, TAB1,
TRAF6, TRAF6. Both cell surface (such as GPCRs, RTKs, TNFRs, TLRs, IL1R and TGFβR) and intracellular (such as RIGI, MDA5) receptors
can activate phosphorylation (activation) of p38, which is mediated via MAPKKKs (TAK1, ASK1/2, DLK 1, TPL2, MEKK 14, MLK2/3, and
TAO1/2) and MAPKKs (MKK3/4/6). In addition, physiological stress can directly activate Rac1 and Cdc42 (receptorindependent)mediated
induction of p38 signalling cascade. Besides, ZAP70 can also activate p38 in MAPKK/MAPKKKindependent manner. The activation of p38
can be inhibited by phosphatases, scaffold proteins and miRNAs. Activated p38 further stimulate several transcription factors/effector
proteins which eventually regulates DNA replication and repair (mediated via GAdd45‐α, Hsp90 and Ck2), transcription (mediated via NF‐κB,
IRF3, IRF7, ATF, CREB, cMyc, AP1, STAT, NFAT) and translation (mediated via MNK1/eIF4E, TIA1, ITAF, PTB, eIFs, FUGP1) of more than
100 cytoplasmic and nuclear substrates. p38 activation also involves regulation of several other cellular processes such as apoptosis (mediated
via p53, caspases, CHOP), cellular trafficking and cytoskeleton remodeling (mediated via Rab5A and EEA1). In addition, p38mediated
activation of transcription factors such as NF‐κB, IRF3, IRF7, ATF, CREB, cMyc, AP1, STAT and NFAT results in activation of several
proinflammatory cytokine genes. Like cellular processes, p38mediated regulation of transcription factors and effector proteins also modulate
several viral processes such as replication, transcription and translation of viral genome, virus induced apoptosis, trafficking of viral proteins
and assembly of viral replication complexes. Although p38 activation leads to induction of proinflammatory cytokines genes which limit the
virus replication but its hyperactivation may lead to cytokine storm (immunopathology). Since p38 supports several steps of virus replication
cycle and may involve in inducing cytokine storm, inhibitors targeting p38 signalling pathways may restrict virus replication, besides
controlling cytokine storm and hence these may serve as novel therapeutic agents
CHANDER ET AL.
-
3
scaffolding proteins which simultaneously interact with multi
components
26
as well as microRNAs (miRNAs),
27,28
may also regulate
p38 activity.
3
|
ROLE OF p38 IN VIRUS REPLICATION
A wide variety of viruses are known to directly interact with p38 or
its substrates. While some of the interactions are proviral others are
inhibitory to virus replication. On the contrary, some viruses may
subvert p38 functions to effectively replicate inside cells. The cellular
events that involve virusp38 interactions are discussed.
3.1
|
p38Dependent regulation of inflammatory
response during viral infections
The major outcome of the activation of MAPK pathway is the pro-
duction of proinflammatory cytokines which eventually produce an
antiviral state. The p38 regulates virusinduced inflammatory re-
sponses by regulating expression of proinflammatory cytokines,
arachidonic acid metabolite, chemokines and cell adhesion molecules.
Suboptimal production of these mediators leads to immune defi-
ciency whereas aberrant expression may result in cytokine storm
(immunopathology). p38 regulates the production of these inflam-
matory mediators (cytokines) at the level of messenger RNA (mRNA)
transcription, translation and processing (splicing, stability, decay and
translocation).
3.1.1
|
Modulation of transcription factors
p38mediated regulation of cytokine gene expression involves mod-
ulation of both transcriptional and epigenetic factors.
29
Out of the
several upstream regulators, predominantly TAK (transforming
growth factor‐β‐activated protein1)mediated p38 activation results
in inflammatory response. It involves activation of a number of
transcriptional factors such as NF‐κB, ATF, CREB, cMyc, AP1, NFAT
and IRFs.
30
Furthermore, p38 may also directly modulate the func-
tions of transcription factors or it may be mediated via its substrates
such as MAPKactivated protein kinase 2/3 (MK2/3), mitogenand
stressactivated kinase 1 (MSK1/2) and MK5 (also known as p38
regulated/activated protein kinase, PRAK).
31
p38mediated MK2
activation results in an inflammatory response (involving transcrip-
tion factors such as NFAT, AP1 and Myc) whereas p38mediated
MSK1 activation results in the suppression of inflammatory response
which is mediated via CREB and ATF1.
3235
p38 upstream regulator TAK1 may also directly activate certain
transcription factors which predominantly involve NF‐κB and IRFs.
30
In resting (inactive) stage, NF‐κB resides in the cytoplasm where
inhibitory‐κB (IκB ) prevents its localization into the nucleus by
binding with it. TAK1mediated degradation of IκB, which is mediated
via inhibitory‐κB kinase‐α (IKKα) and IKKβ, allows the release of
NF‐κB (called NF‐κB activation). The activated NF‐κB then enters
into the nucleus to activate the transcription of several proin-
flammatory cytokine genes. Like TAK1, some viruses such as Sendai
virus (SeV) and Rous sarcoma virus (RSV) may also induce the
activation of NF‐κB by directly phosphorylating IKKs.
36,37
In the
nucleus, p38 induces acetylation of NF‐κB by recruiting p300 and
CREBbinding protein (CBP), which eventually triggers cytokine in-
duction.
38
Additionally, TAK1 also leads to IRF3/IRF7 phosphoryla-
tion, mediated via TBK1 and kinase I kappa B kinase i (IKKi).
39
Upon
activation, IRF3 and IRF7 induce transcriptional activation of in-
terferons (IFN‐α, IFN‐β and IFN‐γ). The activation of IRF3 leads to
type I IFN induction (INF‐α, INF‐β) as is observed in influenza A vi-
ruses (IAV) infection.
40
Similarly, chikungunya virus leads to IRF7
phosphorylation which upregulates the IFN‐γ and TNF‐α production.
Like several other kinases, TAK1 also plays a dual role in the
viral life cycle. On the one hand, it facilitates virus replication
whereas on the other hand it participates in the induction of
antiviral immune response. For example, human immunodeficiency
virus 1 (HIV1) Vpr protein activates NF‐κB and AP1, mediated via
TAK1. By interacting with the viral promoters, NF‐κB and AP1
eventually facilitate viral genome synthesis.
41
On the other hand,
TAK1p38 signalling activates NF‐κB to induce antiviral innate
immune response.
42
In contrast, for effective propagation, several
viruses are known to downregulate TAK1mediated antiviral innate
immune responses. For instance, transporter associated with anti-
gen processing; a virusinducible endoplasmic reticulum (ER)asso-
ciated protein inhibits TAK1 to suppress antiviral immune
response.
42
Similarly, during coxsackie virus A 16 infection, phos-
pholipase β2; GPCRassociated protein inhibits virusinduced TAK1
phosphorylation which eventually suppresses the activation of
antiviral responses.
43
However, the over activation of antiviral re-
sponses (cytokine storm) may lead to host damage (immunopa-
thology) as commonly seen in H5N1 influenza virus, SARSCoV2,
EBOV, DENV and feline infectious peritonitis virus infection.
15
In
such instances, cytokine storm may be therapeutically managed by
administration of small molecule chemical inhibitors that suppress
activation of p38 and/or other related signal transduction pathways
(Table 1).
In addition to NF‐κB, viruses may also induce an inflammatory
response by activating another transcription factor such as AP1
(a p38 substrate). Virusinduced AP1 activation results in the pro-
duction of antiviral cytokines, besides inducing apoptosis. These
cellular events are detrimental to the virus replication.
81
Pharma-
cological inhibition of p38 activation during SeV and RSV infection
suppresses AP1 that results in decreased cytokine production (IFN‐β,
C–C motif chemokine ligand 2, chemokine ligand8 and interleukin6
[IL6])
81
thereby facilitating virus replication.
The ATF/CREB family is another group of transcription factors
which regulate cell proliferation and apoptosis by regulating the
production of IL32, an antiviral cytokine.
82
For example, CREB
induced IL32 induction inhibits IAV replication.
83
Likewise, p38
mediated phosphorylation of STAT1 antagonizes virus infection via
activation of IFN pathway.
84
For example, STAT1 serves as a natural
4
-
CHANDER
ET AL.
host restriction factor against hepatitis C virus (HCV),
85
herpes
simplex virus1 (HSV1),
86
hepatitis E virus (HEV)
87
and IAV.
88
p38 also regulates cyclooxygenase2 (COX2)/prostaglandin E2
signalling pathway to produce several inflammatory mediators. COX
2 potentiates the host immune response against IAV by upregulating
IL3mediated STAT1/2/3 phosphorylation and protein kinase R (PKR)
activation.
89
In contrast, COX2 also facilitates transcription and
translation of viral genes, for example, in cytomegalovirus (CMV),
70
HCV, DENV,
90
mouse hepatitis coronavirus (MHV),
91
feline
calicivirus (FCV), murine norovirus
92
and sapovirus.
93
3.1.2
|
Modulation of RNAbinding proteins
p38 recruits RNAbinding proteins (RBPs) at AUrich elements
(AREs) of viral/cellular mRNA at its 30untranslated region (30UTR).
For example, p38mediated recruitment of human antigen R and
KHtype splicing regulatory protein at cytokine mRNA prevents its
degradation.
94,95
p38 itself can bind with AREs of several cytokines
(IL8, IL6, IL3, IL2 and IL1) and cellular enzymes (COX2) to
determine the stability of their mRNA.
96
3.1.3
|
Modulation of translation
p38 participates both in capdependent and internal ribosomal entry
sites (IRES)mediated translation of cellular/viral mRNA. p38 directly
phosphorylates MAP kinaseinteracting kinases 1 (MNK1), which
supports the initiation of mRNA translation via phosphorylation of
eIF4E and recruiting eIF4G at translation initiation complex.
97
The
eIF4E recognizes and binds with 7methylguanosinecontaining 50
cap of mRNA during translation initiation. By facilitating translation
of cytokines and other inflammatory mediators, p38/MNK1 pathway
plays a key role in inducing inflammatory response following viral
infections.
97,98
Alternatively, p38dependent but MNK1indepen-
dent, induction of type I IFNs and IL12 mRNA translation has also
been implicated in SeV infection.
99
3.2
|
p38Dependent regulation of immune
response in viral infections
p38 is known to modulate both innate and acquired antiviral immune
responses. This involves regulation of immune cell proliferation, cell
TABLE 1p38 signalling as target for antiviral drug development
Association with p38 Cellular target Target virus Inhibitors References
Upstream p38
regulators
ASK1 PCV2 Thioredoxin
44
MKK3/6 ARV Dorsomorphin
45
p38 p38 EMCV, HPAIV, HBV, HSV,
CVB3, NDV, RV,
EBV, MHV, PyV and EV71
SB203580, SB202190, NJK14047,
BCT197, BX795,
SB239063, p38kinhIII, wogonin,
oxymatrine, vemurafenib
4656
p38regulated kinases MSK1/2 RSV, EBOV, HIV1 and KSHV H89, SB747651A
57,58
Casein kinase II HCV, HIV, HDV, HSV,
BTV and HPV
DRB (5,6dichloro1‐β‐ d
ribofuranosylbenzimidazole),
chrysin, benzothiophenes, hypericin,
DMAT (MBS384443), TBCA
(tetrabromocinnamic acid), TBB
(4,5,6,7tetrabromobenzotriazole)
and CX4945
5965
p38dependent cell
cycle regulator
CDC25B AIV NSC95397
66
eEF2K/eEF2 CVB3 Emodin
67
p38associated enzyme COX1/2 DENV2, PSaV, HCMV,
MHV and FCV
NS398, indomethacin celecoxib, tolfenamic
acid, curcumin and SC560
6872
HDAC HIV1 Valproic acid (VPA)
73
ADAM17 HIV1 TAPI2
74
p38associated
translational
factors
MNK1/eIF4E BPXV CGP57380
75
eIF4E/eIF4 G CoV and BPXV 4E2RCat, 4EGI1, apigenin
76
p38dependent
transcriptional
factors
cMyc Adenovirus, HSV1 and AIV CB839
77
CREB VZV XX65023
78
cFOS KSHV Cycloheximide (CHX)
79
p53 PRV PFT‐α
80
CHANDER ET AL.
-
5
differentiation, antigen presentation and cell migration.
100
p38 pro-
motes proliferation and differentiation of immune cells by triggering
the production of chemokines/cytokines such as granulocyte–
macrophage colonystimulating factor, erythropoietin and cluster of
differentiation 40.
101104
p38 Substrate myocyte enhancer factor2
(MEF2) regulates myocyte differentiation which plays essential roles
in Tcell proliferation and transformation during human Tlympho-
tropic virus 1 (HTLV1) infection.
105
p38Dependent expression of
cluster of differentiation 1day (CD1d) molecules (structurally similar
to major histocompatibility complex [MHC] class I) also promotes
vaccinia virus, vesicular stomatitis virus and lymphocytic choriome-
ningitis virus antigen presentation to natural killer T (NKT)
cells.
106,107
Once recognizing viral antigen in the context of CD1d
molecules, the NKT cells rapidly produce cytokines and activate cells,
regulating both innate and adaptive immune responses. Similarly,
chemokine ligand 14 suppresses human papillomavirusassociated
head and neck cancer through antigenspecific CD8+Tcell re-
sponses by upregulating MHCI expression.
108
In addition, MK2/3
activates lymphocytespecific protein 1 (LSP1) which acts as an
intracellular Factin binding protein on immune cells and regulates
neutrophil motility, adhesion to matrix protein and transendothelial
migration. LSP1 facilitates HIV1 transport into the proteasome.
109
3.3
|
p38Dependent transcription and translation
of viral genome
Viruses are heavily dependent on the host machinery for transcrip-
tion and translation of their genome. Efficient replication of viruses
involves supportive functions of over a thousand different cellular
proteins. p38 has also been extensively studied for its role in tran-
scription and translation of cellular as well as viral genome.
3.3.1
|
Regulation of replication and transcription of
viral genome
p38 modulates over 30 different transcription factors which regulate
the expression of several cellular and viral genes. Some of the tran-
scriptional factors directly bind with the viral promoters and facili-
tate expression of the viral genes. For example, by binding with
the long terminal repeats (LTRs) of HIV1, NFAT and NF‐κB facilitate
the expression of viral genes.
110112
p38mediated regulation of virus replication also involves
epigenetic modulation of host/viral genome. For example, HTLV1
Tax protein triggers p38/MEF2/CREB signalling axis which results in
recruitment of histone acetyl transferases (HATs) such as p300, CBP,
and PCAF (p300/CBPassociated factor) at viral LTRs. This facilitates
the disassembly of nucleosome (opening double helix of DNA) and
hence the expression of viral genes).
113
Similarly, MEF2mediated
acetylation of Epstein–Barr virus (EBV) protein BamHI Z fragment
leftward open reading frame 1 (BZLF1) results in its activation which
eventually initiates lytic viral replication.
114
Likewise, activation of
p38/MSK1/CREB1 axis facilitates Kaposi's sarcomaassociated
herpesvirus,
115
VaricellaZoster virus (VZV) and HBV
116
replication
via recruitment of HATs. Besides regulation of the expression of early
genes, p38MEF2 may also trigger p38MAPK dependent phos-
phorylation of ATF2/cjun which ensures gene expression in late
(lytic) phase of virus replication.
117
In contrast, p38dependent transcriptional repression of viral
genes may aid in the maintenance of viral latency. For example, p38
downstream substrates cMyc and Sp1 (specificity protein 1) recruits
histone deacetylase1 (HDAC1) which leads to the repression of HIV
1 gene expression and hence maintenance of latency.
73
Likewise,
during EBV latency, p38regulated class II HDAC binds with MEF2
and induces its proteosomal degradation. This leads to inhibition of
the nuclear translocation of MEF2, which is required for acetylation
of BZLF1 promoter and hence its expression, eventually resulting in
maintenance of viral latency.
114,118
3.3.2
|
p38Dependent translation of viral proteins
p38 regulates both capdependent and IRESmediated translation of
viral mRNA. A large number of viruses exploit the capdependent
mechanism of mRNA translation which is mediated via activation of
p38/MNK1/eIF4E2 axis.
10
For example, buffalopox virus exploits the
cellular machinery of translation by recruiting MNK1/eIF4E.
119
p38
also recruits several RBPs in order to initiate IRESmediated trans-
lation initiation of cellular/viral genome. For example, in picornavirus
infection, p38 facilitates recruitment/activation of several RBPs such
as PTB, eIFs, and ITAFs for efficient translation of viral mRNA.
120,121
Similarly, p38regulated TIA1 and TIAR interact with 50UTR of
enterovirus71 (EV71) genome and 30stemloop of flaviviruses (West
Nile virus [WNV] and DENV) to support translation initiation of viral
proteins.
122124
Likewise, the depletion of TIA1 or TIAR results in
reduced viral mRNA and protein synthesis along with reduced virus
yields during Newcastle disease virus (NDV) infection.
125
FUBP1,
another p38 downstream protein, binds with IRES of EV71 mRNA
and supports its translation.
126
In contrast, FUBP1 inhibits replica-
tion of Japanese encephalitis virus (JEV) via interacting with
viral UTRs.
127
Some p38 downstream substrates such as MK2/3 and MK5 also
regulate viral mRNA translation by either phosphorylating HSP27 or
via inhibiting PKR. The activated HSP27 is known to facilitate IRES
mediated EV71 replication.
128
In contrast, HSP27 also negatively
regulates several RNA virus replications which is mediated via
potentiating antiviral immunity. For example, HIV1,
129
classical
swine fever virus (CSFV)
130
and Zika virus infections.
131
PKR is
activated via double stranded RNA, an intermediate in the synthesis
of several RNA viruses. Activated PKR phosphorylates eIF2αwhich
inhibits global protein translation. This may also limit viral replication.
Therefore, in order to efficiently replicate, viruses such as IAV an-
tagonizes the function of activated PKR by phosphorylating MK2/3, a
p38 substrate. The activated MK2/3 interacts with p88 to inhibit
PKR thereby selectively restoring the viral mRNA translation.
132
6
-
CHANDER
ET AL.
3.4
|
Modulation of apoptosis during viral infections
3.4.1
|
Role of p38 in virusinduced apoptosis
Induction of apoptosis includes both extrinsic and intrinsic signals
(Figure 2). Four overlapping mechanisms are employed by the viruses
to induce p38mediated induction of apoptosis. The first one involves
transcriptional regulation of apoptotic proteins. For example, SARS
CoV 3a protein induces p38dependent upregulation of p53 which
triggers the release of cytochromec and activates caspase9 to
induce apoptosis. Similarly, HIV1 glycoprotein 120 (gp120) induces
p38 which results in NF‐κB and p53mediated apoptosis.
133
The
second mechanism involves p38dependent enhanced expression of
TNF‐α and Fas ligand. For example, DENV induces apoptosis by
upregulating p38dependent TNF‐α expression in mice. Third
mechanism involves p38dependent induction of cysteine proteases
(caspases 3/8/9). For example, p38mediated caspases 3 activation
during coxsackievirus B3 (CVB3) infection leads to death of car-
diomyocytes. Similarly, caspase 3/8/9 activation is also associated
with DENV and IAVinduced apoptosis. Fourth, accumulation of
unfolded/misfolded proteins or overloading of viral proteins in ER
eventually leads to ER stress which may also induce apoptosis. For
example, IAV recruits inositol requiring enzyme 1 in ER which in-
teracts with TNF receptorassociated factor 2 and apoptosis signal
regulating kinase 1 (ASK1) to induce p38mediated apoptosis. Like-
wise, some RNA viruses activate PERK (RNAdependent protein ki-
nase [PKR]like ER kinase) which induces eIF2α/ATF4mediated
activation of CHOP, a substrate of p38 (Figure 2).
134
The activated
CHOP induces apoptosis via caspase cascades (caspase 11).
3.4.2
|
Regulation of p38induced apoptosis by
viruses
The early induction of apoptosis may result in nonproductive virus
replication. Therefore, in order to effectively replicate inside the host,
a virus must subvert the cellular apoptotic machinery to inhibit
apoptosis. HIV1 Nef protein inhibits ASK1dependent p38 activa-
tion that enhances cell survivability and hence increases virus
yield.
135
In addition, HIV1 Nef protein also interacts and competi-
tively inhibits p53 functions to inhibit apoptosis.
135
Likewise, IAV
NS1 protein downregulates AP1 to inhibit apoptosis.
136
Several vi-
ruses inhibit p38mediated apoptosis by activating PI3K/Akt signal-
ling pathways.
137
In addition, some viruses are known to exploit
STAT3 for inhibiting p38dependent apoptosis.
138,139
This is medi-
ated via suppression of p53 activity or induction of antiapoptotic
proteins such as Bcl2 and BclxL. Some viruses encode viral FADD
like interleukin 1 beta converting enzyme inhibitory proteins (vFLIP)
that have cellular deatheffector domain (DED), like those present in
caspases8 and Fasassociated protein with death domain (FADD),
the extrinsic apoptotic receptor. The vFLIPs encoded by molluscum
contagiousum virus (MC159L) and equine herpesvirus 2 (EHV2)
(MC159L and E8) bind with the DED(s) of caspase 8 or FADD to
inhibit apoptosis and cell death.
140
The vFLIP, encoded by human
herpesvirus 8 (EHV2; ORF K13) inhibits NF‐κBinduced apoptosis.
141
FUBP1 (p38 substrate) also antagonizes apoptosis by suppressing
transcriptional activity of p53. For example, the E1A protein of
adenovirus 5 stabilizes the p53FUBP1 interaction and antagonizes
p53dependent apoptotic mechanism.
142
However, induction of apoptosis during late phase of viral life
cycle may be beneficial for dissemination of the infection to the
adjacent cells. p38 has two major downstream substratesp53 and
STAT3. p53 induces apoptosis whereas STAT3 antagonizes p53
function. During the resting stage, STAT3 is located in the cytoplasm.
Upon p38mediated phosphorylation, it translocates into the nucleus
where it inhibits p53 phosphorylation thereby leading to inhibition of
apoptosis (cell survival). To ensure effective apoptosis, SARSCoV
protein 3a not only induces p38mediated activation of p53 but it
also interacts with and inhibits nuclear translocation of STAT3.
143
3.5
|
Modulation of cell cycle and virus growth
p38 modulates activity of several transcription factors and other
kinases/metabolic enzymes which eventually regulates cell cycle
progression as well as apoptosis. However, whether the end effect
would result in cell survival or apoptosis is determined by the nature
of downstream kinase/transcription factors involved. p38mediated
cell survival involves modulating the activity of cyclins and cyclin
dependent kinase (CDK) inhibitors which eventually regulate cell
cycle arrest at G1/S and G2/M to facilitate cell repair/differentiation.
The p38 performs cell survival function by either inducing cell
cycle arrest (allows sufficient time for DNA repair) by activating
Gadd45‐α (growth arrest and DNA damageinducible protein
GADD45 alpha) and inhibiting cell cycle control protein 2 (CDC2).
Gadd45‐α is known to induce G2/M arrest which facilitates the
repair of cellular DNA and cell survival.
144
In HIV1 infected cells,
GADD45‐α represses the transcription of HIV1 genes
145
thereby
facilitating the induction of latent viral infection which is a major
barrier to HIV1 eradication.
145
p38 also inhibits CDC2 activity,
mediated via inhibiting CDC25B/C (also known as Mphase inducer
phosphatase 2/3) that eventually results in the delay of G2 phase.
Some viruses can replicate only in dividing cells and therefore they
are known to activate CDC2. For example, VZV glycoprotein gI,
146
EBV protein EBNALP,
147
HEV ORF3
148
and HSV1 ICP0
149
activate
CDC2. The activated CDC2 facilitates cellular and viral mRNA
translation, mediated via recruitment of eukaryotic elongation factor
2 (eEF2).
150
However, viruses such as CVB3
67
and HIV1
151
exploits
eEF2 for translating their protein, besides inducing CDC2dependent
cell progression which is essential for their replication. In contrast,
some viruses require nondividing cells for their replication and
have the capacity to inactivate CDC2. For instance, reoviral S1
geneencoded ς1s nonstructural protein,
152
papilloma virus E2
153
and HIV1 Vpr protein.
154
Besides inducing cell cycle arrest, virusp38 interaction is also
shown to reprogram cellular apoptosis so as to keep the cells in
CHANDER ET AL.
-
7
dividing phase for their optimal replication. The reprogramming of
apoptosis by p38 is mediated via phosphorylation of CK2 and Hsp90/
cell division cycle 37 (Hsp90/Cdc37). CK2 inhibits caspasemediated
apoptosis and triggers DNA damage response which eventually fa-
cilitates cell cycle progression from G1 to S phase and G2 to M phase.
Hsp90Cdc37 is a cochaperone that binds and stabilizes several
kinases (CDK4, CDK6 and eIF2α) required for cell cycle progression.
Viruses are known to exploit both of these p38 substrates (CK2 and
Hsp90Cdc37) to ensure their efficient replication. For example, CK2
phosphorylates HCV (NS5A) and bluetongue virus (NS2) pro-
teins
59,155
which are required for viral RNA synthesis and assembly.
Similarly, HSP90Cdc37 supports viral assembly and genome
synthesis of duck hepatitis B virus (hepadnavirus).
156
3.6
|
Modulation of cell cytoskeleton and virus
growth
Besides regulating several intracellular cell signalling cascades, p38
also regulates remodeling of cell cytoskeleton. Viruses employ
several distinct mechanisms to exploit p38dependent endocytic
remodeling (vesicular transport) to facilitate intracellular transport of
virus/viral proteins
157
and celltocell transmission (viral synapses).
p38mediated activation of EEA1 and Rab5A
158
are required for
early endosomal fusion and endosomal trafficking
159
which assist in
endocytosis of RSV and IAV.
160
Similarly, Rab5A facilitates endocy-
tosis (entry) of DENV and WNV.
161,162
Besides endocytosis, Rab5A
and EEA1 also modulate virus replication complexes (assem-
bly).
163,164
For instance, CSFV NS4B is known to interact with
Rab5
163
whereas HCV NS4B interacts with EEA1 and Rab5A for the
optimal assembly of the virion components.
164
Zetachainassociated protein kinase 70 (ZAP70; an upstream
regulator of p38) modulates actin cytoskeleton and polarization of
microtubuleorganizing center.
165
HIV1 exploits this remodelled
cytoskeleton activity for efficient celltocell transmission by forming
virological synapses.
166
This is due to the fact that ZAP70 defective
lymphoid cells exhibit restricted celltocell HIV1 transmission
capability.
166
TypeI transmembrane protease ADAM17 (disinterring and
metalloproteinase domaincontaining protein 17) is required for
cleavage and release of a soluble ectodomain (exosomes) from
membranebound proproteins. This is required for the release of
FIGURE 2 Role of p38 signalling in virusinduced apoptosis: p38 participates in regulating both intrinsic and extrinsic pathways of
apoptosis. Endoplasmic stress response which is activated upon accumulation of viral proteins in the endoplasmic reticulum recruits IRE1
which activates p38dependent apoptosis. p38 can also be induced by activation of PRRs such as RIG1, MDM5 and TLRs. Activated p38
regulates apoptosis at various levels which includes (i) stabilization of p53 that induce cytochromec release form mitochondrial wall by up
regulating apoptotic protein (BaX) that eventually activates caspases 3/9. (ii) Direct activation of caspases cascade. (iii) modulating CHOP
activity, which is mediated via PERK/eIF2α/ATF4 signalling cascade. (iv) inducing cytokine stress (upregulate expression of TNFα) and
(v) upregulating expression of FasL
8
-
CHANDER
ET AL.
diverse groups of membraneanchored cytokines, cell adhesion
molecules, receptors, ligands, and enzymes. The p38mediated acti-
vation of ADAM17 leads to epidermal growth factor receptor
dependent cell proliferation. Several viral families are known to
exploit these transmembrane proteases for optimum infectivity. For
example, HIV1 Nef protein facilitates the release of ADAM17con-
taining exosomes. It activates virus supportive intracellular signalling
via cleavage of proTNF‐α into mature form. This supports the
infectivity of incoming HIV1 in quiescent CD4
+
T lymphocytes.
167
Likewise, ADAM17 also facilitates the entry of papillomavirus by
induction of growth factors which triggers the formation of an
endocytic entry platform.
168
By regulating expression of the cytoskeleton proteins such as
Rho family GTPasesRhoA, (Ras homology family member A), Racl
(Rasrelated C3 botulinum toxin substrate 1), and CDC4, p38asso-
ciated transcriptional factor serum response factor (SRF) maintains
the cytoskeleton integrity of cardiac cells.
169
The CVB3encoded 2A
protease cleaves SRF to dissociate its transactivation domain
from the DNAbinding domain. This leads to downregulation of SRF
mediated gene expression which ultimately results in the disruption
of cytoskeleton integrity, hence resulting in cardiac pathology.
170
4
|
TARGETING p38 SIGNALLING FOR ANTIVIRAL
DRUG DEVELOPMENT
Over 250 antiviral drugs have been approved by the FDA
3
but the
majority of these act by directly targeting viral encoded factors.
3
In
fact, successful drugs (directly acting) may eventually fail because of
the emergence of drugresistant mutants.
171,172
Indeed, drug resis-
tance has been observed against most of the virusdirected antiviral
drugs approved so far. Therefore, alternative approaches towards
antiviral drug development need to be explored.
More than 80 kinase inhibitors have been evaluated which are
in some stage of clinical trial and over 35 distinct kinases have been
developed to the level of phase I clinical trial.
173
Most of the
currently known kinase inhibitors target at the kinase activation
loop in the ATP binding site
173
and have been initially investigated
for the treatment of cancer.
174
Increasing evidence also suggests
kinases as potential drug targets against viral infections.
175
In
viral infections, p38 signalling pathway may be targeted to fight at
least at two fronts, restricting virus replication and minimizing
immunopathology induced due to the over activation of immune
response.
4.1
|
Inhibiting virus production
As indicated above, viruses may exploit enhanced cell signalling ac-
tivity for completion of its replication cycle. The viral replication cycle
is a multistep process that includes attachment, entry, genome syn-
thesis, assembly of newly synthesized virion particles and budding.
Each step of the viral life cycle involves host cell kinases with a single
kinase regulating single or multiple steps of the viral life
cycle.
79,136,176,177
p38mediated regulation of transcription factors
and effector proteins support several viral processes such as repli-
cation, transcription, translation, virusinduced apoptosis, trafficking
of viral proteins and assembly of viral replication complexes
(Figure 1). Therefore, targeting these hostdependency factors may
also serve as a novel strategy for antiviral drug development
(Figure 1). For instance, p38 signalling pathway is needed for efficient
synthesis of SARSCoV2 subgenomic mRNA.
1
Likewise, NJK14047
(p 38 inhibitor) inhibitor blocks synthesis of HBV pregenomic RNA
and covalently closed circular DNA
6
and BX795 inhibits p38medi-
ated expression of HSV1 genes.49
9
In consequence, it is likely that
targeting the signalling pathway(s) which regulate multiple steps of
viral replication cycle could be a most effective strategy to develop
antiviral drugs.
175
Kinase requirement is usually common among the members of a
particular virus family.
75
Nevertheless, p38 is also known to support
replication of multiple viral families which includes IAV, RSV, HBV,
HSV1, coxsackievirus, NDV, rotavirus, encephalomyocarditis, coro-
navirus and enterovirus).
4656
In this context, p38 inhibitor
SB203580 which is known to block SARSCoV2, human coronavirus
229E and DENV
178181
replication, may serve as a broadspectrum
antiviral agent.
4.2
|
Blocking cytokine storm
Cell signalling pathways are activated with the basic aim of inducing
an antiviral state. p38mediated activation of transcription factors
such as NF‐κB, IRF3, IRF7, ATF, CREB, cMyc, AP1, STAT and NFAT
activate several proinflamamtory cytokine genes, which usually limit
virus replication (Figure 1). However, certain acute viral infections
such as SARSCoV2,
11
EBOV,
12
DENV
13,14
and H5N1
14
lead to the
hyperactivation of signalling pathways, producing a cytokine storm
(Figure 1). More than 150 cytokines are induced during cytokine
storms but those primarily involved include TNF‐α, IL6 and IFNs.
15
The cytokine storm leads to tissue damaging immunopathology.
Under such instances, p38 signalling pathway may be therapeutically
managed in minimizing the immunopathology induced by the cyto-
kine storms (Table 1). For instance, p38specific inhibitor SB203580
inhibits cytokine storm induced in response to SARSCoV2, EBOV
and DENV infection.
178181
Likewise, PD169316
182
and SB 239063
52
which also target p38 signalling pathway restrict EV71and rhino-
virusinduced cytokine storms.
Considering the p38 requirement for optimal virus replication
and its role in inducing cytokine storm, it is tempting to speculate
that p38 suppression may have dual effects, first restricting viral
production in the target cells and secondly, blocking the cytokine
storm. For example, p38 inhibitor SB203580 restricts SARSCoV2
replication by targeting the viral genome synthesis, besides sup-
pressing p38mediated cytokine storm.
178
Similarly, SB203580 and
SB202190 were also shown to suppress EBOV replication (at entry
level) and the associated cytokine storm.
183
Likewise, oxymatrine
CHANDER ET AL.
-
9
was shown to restrict IAV replication and virusinduced
immunopathology.
184
Besides directly inhibiting p38 phosphorylation (activation),
regulating p38 upstream (such as ASK1, MKK3/6) or downstream
(such asMSK1/2, casein kinase II, CDC25B, COX1/2, HDAC,
ADAM17, MNK1/eIF4E) kinases also serve as targets for antiviral
drug development (Table 1). However, most of these studies have
been conducted on a laboratory scale while only few of them have
entered clinical trials.
5
|
CHALLENGES OF p38DIRECTED ANTIVIRAL
THERAPY
A major criticism of the hostdirected antiviral therapy is the side
effects which may emerge due to suppression of certain cellular
functions. Some kinase inhibitors have been shown to induce toxicity
and offtarget effects which could result in hypertension, hypothy-
roidism, skin reactions, cardiotoxicity and proteinuria.
185,186
How-
ever, hostdirected therapeutic agents which are in clinical use
against cardiovascular and inflammatory diseases or cancers have
minimal or no adverse side effects.
187
Nevertheless, p38 inhibitor SB
681323 has also been shown to be well tolerated in treating
neuropathic pain.
188
The major advantage of hostdirected antiviral therapy is
considered to be due its potential in minimizing the drug resistance,
as compared to the directlyacting antiviral agents. However, drug
resistance against some hostdirected agents can, in fact, occur under
certain circumstances. For instance, longterm selection pressure of a
hostdirected antiviral agent allows the virus the opportunity to
adapt to use an alternate host factor.
189
Likewise, under longterm
selection pressure, virus can alter its affinity towards the target that
confers resistance.
190
In addition, longterm kinase inhibition may
induce pressure on the host cells to acquire resistance against the
targeted agent through kinase mutations.
191194
6
|
CONCLUSIONS
Classically, antiviral drugs are developed by targeting certain viral
proteins but, under selective pressure, viruses quickly acquire drug
resistance. The human genome encodes 518 kinases that have been
implicated in the regulation of various physiological processes to
maintain homeostasis. Like several other biotic and abiotic stresses,
viral infections are also associated with perturbation of the “kinome.”
Some of the cellular kinases are dispensable for host cell viability
but might be needed during virus infection and thus serve as novel
targets for antiviral drug development. Some kinases are required for
multiple viral family members and therefore may be targeted for the
development of broadspectrum antiviral drugs. Since the virus
cannot easily regain the missing cellular events by mutations, host
directed antiviral therapies should have a reduced risk of generating
drugresistant viruses. We have provided insights into how viruses
interact with p38 signalling pathway (p38/associated MAPKS/sub-
strates) and how these virushost interaction events can be thera-
peutically managed in restricting virus replication and minimizing
immunopathology (cytokine storm). Most of these p38 inhibitors are
under preclinical development. Further studies on cytotoxicity and
antiviral efficacy of these inhibitors in clinical trials are essential.
ACKNOWLEDGEMENTS
This work was supported by the Science and Engineering Research
Board (Grant number CRG/2018/004747, CRG/2019/000829
and CVD/2020/000103), Department of Science and Technology,
Government of India.
AUTHORS CONTRIBUTIONS
Naveen Kumar, Yogesh Chander and Ram Kumar wrote the original
draft. Naveen Kumar, Ram Kumar, Nitin Khandelwal, Namita Singh,
Sanjay Barua, and Brij Nandan Shringi edited the manuscript. Naveen
Kumar and Sanjay Barua received the funding.
ORCID
Naveen Kumar
https://orcid.org/0000-0003-3974-9409
REFERENCES
1. Walker PJ, Siddell SG, Lefkowitz EJ, et al. Changes to virus tax-
onomy and the International Code of Virus Classification and
nomenclature ratified by the International Committee on Taxon-
omy of Viruses (2019). Arch Virol. 2019;164:24172429. https://
doi.org/10.1007/s00705-019-04306-w.
2. Goris N, Vandenbussche F, De Clercq K. Potential of antiviral
therapy and prophylaxis for controlling RNA viral infections of
livestock. Antiviral Res. 2008;78:170178. https://doi.org/10.1016/
j.antiviral.2007.10.003.
3. Chaudhuri S, Symons JA, Deval J. Innovation and trends in the
development and approval of antiviral medicines: 19872017 and
beyond. Antiviral Res. 2018;155:7688. https://doi.org/10.1016/j.
antiviral.2018.05.005.
4. Ardito F, Giuliani M, Perrone D, Troiano G, Lo Muzio L. The crucial
role of protein phosphorylation in cell signaling and its use as
targeted therapy (Review). Int J Mol Med. 2017;40:271280.
https://doi.org/10.3892/ijmm.2017.3036.
5. Milanesi L, Petrillo M, Sepe L, et al. Systematic analysis of human
kinase genes: a large number of genes and alternative splicing
events result in functional and structural diversity. BMC Bioinfor-
matics 2005; 6 Suppl 4: S20. https://doi.org/10.1186/1471-2105-6-
S4-S20.
6. Briedis KM, Starr A, Bourne PE. Analysis of the human kinome
using methods including fold recognition reveals two novel kinases.
PLoS One. 2008;3:e1597. https://doi.org/10.1371/journal.pone.
0001597.
7. Kumar N, Barua S, Thachamvally R, Tripathi BN. Systems
perspective of Morbillivirus replication. J Mol Microbiol Biotechnol.
2016;26:389400. https://doi.org/10.1159/000448842.
8. Cargnello M, Roux PP. Activation and function of the MAPKs
and their substrates, the MAPKactivated protein kinases. Micro-
biol Mol Biol Rev 2011;75:5083. https://doi.org/10.1128/MMBR.
00031-10.
9. van de Wakker SI, Fischer MJE, Oosting RS. New drugstrategies to
tackle viralhost interactions for the treatment of influenza virus
infections. Eur J Pharmacol. 2017;809:178190. https://doi.org/
10.1016/j.ejphar.2017.05.038.
10
-
CHANDER
ET AL.
10. Kumar R, Khandelwal N, Thachamvally R, et al. Role of MAPK/
MNK1 signaling in virus replication. Virus Res 2018;253:4861.
https://doi.org/10.1016/j.virusres.2018.05.028.
11. Lukan N. Cytokine storm, not only in COVID19 patients. Mini
review. Immunol Lett. 2020;228:3844. https://doi.org/10.1016/j.
imlet.2020.09.007.
12. Younan P, Iampietro M, Nishida A, et al. Ebola virus binding to Tim
1 on T lymphocytes induces a cytokine storm. mBio. 2017;8.
https://doi.org/10.1128/mBio.00845-17.
13. Srikiatkhachorn A, Mathew A, Rothman AL. Immunemediated
cytokine storm and its role in severe dengue. Semin Immunopathol.
2017;39:563574. https://doi.org/10.1007/s00281-017-0625-1.
14. Us D. [Cytokine storm in avian influenza]. Mikrobiyol Bul.
2008;42:365380.
15. Kumar N, Sharma S, Kumar R, et al. Hostdirected antiviral ther-
apy. Clin Microbiol Rev. 2020;33. https://doi.org/10.1128/CMR.00
168-19.
16. Cohen P. The search for physiological substrates of MAP and SAP
kinases in mammalian cells. Trends Cell Biol. 1997;7:353361.
https://doi.org/10.1016/s0962-8924(97)01105-7.
17. Koul HK, Pal M, Koul S. Role of p38 MAP kinase signal transduction
in solid tumors. Genes Cancer. 2013;4:342359. https://doi.org/
10.1177/1947601913507951.
18. Obata T, Brown G, Yaffe M. MAP kinase pathways activated by
stress: The p38 MAPK pathway. Crit Care Med. 2000;28:N67N77.
https://doi.org/10.1097/00003246-200004001-00008.
19. Grimsey NJ, Lin Y, Narala R, Rada CC, MejiaPena H, Trejo J. G
proteincoupled receptors activate p38 MAPK via a noncanonical
TAB1TAB2and TAB1TAB3dependent pathway in endothelial
cells. J Biol Chem. 2019; 294: 58675878. https://doi.org/10.1074/
jbc.RA119.007495.
20. Yoneyama M, Onomoto K, Jogi M, Akaboshi T, Fujita T. Viral RNA
detection by RIGIlike receptors. Curr Opin Immunol. 2015;32:48
53. https://doi.org/10.1016/j.coi.2014.12.012.
21. Wilson KP, Fitzgibbon MJ, Caron PR, et al. Crystal structure of p38
mitogenactivated protein kinase. J Biol Chem. 1996;271:27696
27700.
22. Jiang Y, Gram H, Zhao M, et al. Characterization of the structure
and function of the fourth member of p38 group mitogenactivated
protein kinases, p38δ.J Biol Chem. 1997;272:3012230128.
23. Ge B, Gram H, Di Padova F, et al. MAPKKindependent activation
of p38αmediated by TAB1dependent autophosphorylation of
p38α.Science. 2002;295:12911294.
24. Sun H, Charles CH, Lau LF, Tonks NK. MKP1 (3CH134), an im-
mediate early gene product, is a dual specificity phosphatase that
dephosphorylates MAP kinase in vivo. Cell. 1993;75:487493.
25. NunesXavier C, RomaMateo C, Rios P, et al. Dualspecificity MAP
kinase phosphatases as targets of cancer treatment. Anticancer
Agents Med Chem. 2011;11:109132.
26. Uhlik MT, Abell AN, Johnson NL, et al. Rac–MEKK3–MKK3 scaf-
folding for p38 MAPK activation during hyperosmotic shock. Nat
Cell Biol. 2003;5:11041110.
27. McCaskill JL, Ressel S, Alber A, et al. Broadspectrum inhibition of
respiratory virus infection by microRNA mimics targeting p38
MAPK signaling. Mol Ther Nucleic Acids. 2017;7:256266.
28. Suomalainen M, Nakano M, Boucke K, Keller S, Greber U. Adeno-
virusactivated PKA and p38/MAPK pathways boost microtubule
mediated nuclear targeting of virus. EMBO J. 2001;20:13101319.
29. Phillips T, Hoopes L. Transcription factors and transcriptional
control in eukaryotic cells. Nature Educ. 2008;1:119.
30. Li Q, Verma IM. NFkappaB regulation in the immune system. Nat
Rev Immunol. 2002;2:725734. https://doi.org/10.1038/nri910.
31. Cargnello M, Roux PP. Activation and function of the MAPKs and
their substrates, the MAPKactivated protein kinases. Microbiol Mol
Biol Rev. 2011;75:5083.
32. Bachstetter AD, Van Eldik LJ. The p38 MAP kinase family as reg-
ulators of proinflammatory cytokine production in degenerative
diseases of the CNS. Aging Dis. 2010;1:199211.
33. Gorska MM, Liang Q, Stafford SJ, et al. MK2 controls the level of
negative feedback in the NFkappaB pathway and is essential for
vascular permeability and airway inflammation. J Exp Med.
2007;204:16371652. https://doi.org/10.1084/jem.20062621.
34. Kim C, Sano Y, Todorova K, et al. The kinase p38 alpha serves cell
typespecific inflammatory functions in skin injury and coordinates
proand antiinflammatory gene expression. Nat Immunol.
2008;9:10191027. https://doi.org/10.1038/ni.1640.
35. Ananieva O, Darragh J, Johansen C, et al. The kinases MSK1 and
MSK2 act as negative regulators of Tolllike receptor signaling. Nat
Immunol. 2008;9:10281036. https://doi.org/10.1038/ni.1644.
36. Fink K, Duval A, Martel A, SoucyFaulkner A, Grandvaux N. Dual
role of NOX2 in respiratory syncytial virusand sendai virus
induced activation of NFkappaB in airway epithelial cells.
J Immunol. 2008;180:69116922. https://doi.org/10.4049/jimmu
nol.180.10.6911.
37. Liu P, Jamaluddin M, Li K, Garofalo RP, Casola A, Brasier AR.
Retinoic acidinducible gene I mediates early antiviral response
and Tolllike receptor 3 expression in respiratory syncytial virus
infected airway epithelial cells. J Virol. 2007;81:14011411. https://
doi.org/10.1128/jvi.01740-06.
38. Saha RN, Jana M, Pahan K. MAPK p38 regulates transcriptional
activity of NFkappaB in primary human astrocytes via acetylation
of p65. J Immunol. 2007;179:71017109. https://doi.org/10.4049/
jimmunol.179.10.7101.
39. Jiang M, O¨sterlund P, Fagerlund R, et al. MAP kinase p38αregu-
lates type III interferon (IFN‐λ1) gene expression in human
monocytederived dendritic cells in response to RNA stimulation.
J Leukocyte Biol. 2015;97:307320. https://doi.org/10.1189/jlb.
2A0114-059RR.
40. Hale BG, Albrecht RA, GarcíaSastre A. Innate immune evasion
strategies of influenza viruses. Future Microbiol. 2010;5:2341.
https://doi.org/10.2217/fmb.09.108.
41. Liu R, Lin Y, Jia R, et al. HIV1 Vpr stimulates NF‐κB and AP1
signaling by activating TAK1. Retrovirology. 2014;11:45. https://doi.
org/10.1186/1742-4690-11-45.
42. Xia Z, Xu G, Yang X, et al. Inducible TAP1 negatively regulates the
antiviral innate immune response by targeting the TAK1 complex.
J Immunol. 2017;198:36903704. https://doi.org/10.4049/jimm
unol.1601588.
43. Wang L, Zhou Y, Chen Z, et al. PLCβ2 negatively regulates the
inflammatory response to virus infection by inhibiting phosphoi-
nositidemediated activation of TAK1. Nat Commun. 2019;10:746.
https://doi.org/10.1038/s41467-019-08524-3.
44. Wei L, Zhu S, Wang J, et al. Regulatory role of ASK1 in porcine
circovirus type 2induced apoptosis. Virology. 2013;447:285291.
45. Ji WT, Lee LH, Lin FL, Wang L, Liu HJ. AMPactivated protein ki-
nase facilitates Avian Reovirus to induce mitogenactivated protein
kinase (MAPK) p38 and MAPK kinase 3/6 signalling that is bene-
ficial for virus replication. J Gen Virol. 2009;90:30023009. https://
doi.org/10.1099/vir.0.013953-0.
46. Hirasawa K, Kim A, Han HS, Han J, Jun HS, Yoon JW. Effect of
p38 mitogenactivated protein kinase on the replication of
encephalomyocarditis virus. J Virol. 2003;77:56495656. https://
doi.org/10.1128/jvi.77.10.5649-5656.2003.
47. Börgeling Y, Schmolke M, Viemann D, Nordhoff C, Roth J, Ludwig S.
Inhibition of p38 mitogenactivated protein kinase impairs influ-
enza virusinduced primary and secondary host gene responses
and protects mice from lethal H5N1 infection. J Biol Chem.
2014;289:1327. https://doi.org/10.1074/jbc.M113.469239.
48. Chang WW, Su IJ, Lai MD, Chang WT, Huang W, Lei HY. Sup-
pression of p38 mitogenactivated protein kinase inhibits hepatitis
CHANDER ET AL.
-
11
B virus replication in human hepatoma cell: the antiviral role of
nitric oxide. J Viral Hepatitis. 2008;15:490497. https://doi.org/
10.1111/j.1365-2893.2008.00968.x.
49. Su Ar, Qiu M, Li Yl, et al. BX795 inhibits HSV1 and HSV2
replication by blocking the JNK/p38 pathways without interfering
with PDK1 activity in host cells. Acta Pharmacol Sin. 2017;38:402
414. https://doi.org/10.1038/aps.2016.160.
50. Si X, Luo H, Morgan A, et al. Stressactivated protein kinases are
involved in coxsackievirus B3 viral progeny release. J Virol.
2005;79:1387513881. https://doi.org/10.1128/jvi.79.22.13875-
13881.2005.
51. Zhan Y, Yu S, Yang S, et al. Newcastle disease virus infection ac-
tivates PI3K/Akt/mTOR and p38 MAPK/Mnk1 pathways to benefit
viral mRNA translation via interaction of the viral NP protein and
host eIF4E. PLoS Pathog. 2020;16:e1008610. https://doi.org/
10.1371/journal.ppat.1008610.
52. Griego SD, Weston CB, Adams JL, TalSinger R, Dillon SB. Role of
p38 mitogenactivated protein kinase in rhinovirusinduced cyto-
kine production by bronchial epithelial cells. J Immunol.
2000;165:52115220.
53. Johnson JC, Martinez O, Honko AN, Hensley LE, Olinger GG,
Basler CF. Pyridinyl imidazole inhibitors of p38 MAP kinase impair
viral entry and reduce cytokine induction by Zaire ebolavirus in
human dendritic cells. Antiviral Res. 2014;107:102109. https://doi.
org/10.1016/j.antiviral.2014.04.014.
54. Banerjee S, Narayanan K, Mizutani T, Makino S. Murine corona-
virus replicationinduced p38 mitogenactivated protein kinase
activation promotes interleukin6 production and virus replication
in cultured cells. J Virol. 2002;76:59375948. https://doi.org/
10.1128/jvi.76.12.5937-5948.2002.
55. Dobson SJ, Anene A, Boyne JR, Mankouri J, Macdonald A,
Whitehouse A. Merkel cell polyomavirus small tumour antigen
activates the p38 MAPK pathway to enhance cellular motility.
Biochem J. 2020;477:27212733.
56. Peng H, Shi M, Zhang L, et al. Activation of JNK1/2 and p38 MAPK
signaling pathways promotes enterovirus 71 infection in immature
dendritic cells. BMC Microbiol. 2014;14:147. https://doi.org/
10.1186/1471-2180-14-147.
57. Jamaluddin M, Tian B, Boldogh I, Garofalo RP, Brasier AR. Respi-
ratory syncytial virus infection induces a reactive oxygen species
MSK1phosphoSer276 RelA pathway required for cytokine
expression. J Virol. 2009;83:1060510615. https://doi.org/
10.1128/JVI.01090-09.
58. Li B, Wan Z, Huang G, et al. Mitogenand stressactivated kinase 1
mediates EpsteinBarr virus latent membrane protein 1promoted
cell transformation in nasopharyngeal carcinoma through its in-
duction of Fra1 and cJun genes. BMC Cancer. 2015;15:390.
https://doi.org/10.1186/s12885-015-1398-3.
59. Kim S, Jin B, Choi SH, Han KH, Ahn SH. Casein kinase II inhibitor
enhances production of infectious genotype 1a hepatitis C virus
(H77S). PLoS ONE. 2014;9:e113938. https://doi.org/10.1371/
journal.pone.0113938.
60. Critchfield JW, Coligan JE, Folks TM, Butera ST. Casein kinase II is
a selective target of HIV1 transcriptional inhibitors. Proc. Natl.
Acad. Sci. U.S.A. 1997;94:61106115. https://doi.org/10.1073/
pnas.94.12.6110.
61. Meruelo D, Lavie G, Lavie D. Therapeutic agents with dramatic
antiretroviral activity and little toxicity at effective doses: aromatic
polycyclic diones hypericin and pseudohypericin. Proc. Natl. Acad.
Sci. U.S.A. 1988;85:52305234. https://doi.org/10.1073/pnas.85.
14.5230.
62. Yeh TS, Lo SJ, Chen PJ, Lee Y. Casein kinase II and protein kinase
C modulate hepatitis delta virus RNA replication but not empty
viral particle assembly. J Virol. 1996;70:61906198.
63. Koffa MD, Kean J, Zachos G, Rice SA, Clements JB. CK2 protein
kinase is stimulated and redistributed by functional herpes simplex
virus ICP27 protein. J Virol. 2003;77:43154325. https://doi.org/
10.1128/jvi.77.7.4315-4325.2003.
64. Mohl BP, Roy P. Cellular casein kinase 2 and protein phosphatase
2A modulate replication site assembly of bluetongue virus. J Biol
Chem. 2016;291:1456614574.
65. Piirsoo A, Piirsoo M, Kala M, et al. Activity of CK2αprotein kinase
is required for efficient replication of some HPV types. PLoS Pathog.
2019;15:e1007788.
66. Perwitasari O, Torrecilhas AC, Yan X, et al. Targeting cell division
cycle 25 homolog B to regulate influenza virus replication. J Virol.
2013;87:1377513784.
67. Zhang HM, Wang F, Qiu Y, et al. Emodin inhibits coxsackievirus B3
replication via multiple signalling cascades leading to suppression
of translation. Biochem J. 2016;473:473485.
68. Lin CK, Tseng CK, Wu YH, et al. Cyclooxygenase2 facilitates
dengue virus replication and serves as a potential target for
developing antiviral agents. Sci Rep. 2017;7:44701.
69. Alfajaro MM, Choi JS, Kim DS, et al. Activation of COX2/PGE2
promotes Sapovirus replication via the inhibition of nitric oxide
production. J Virol. 2017;91.
70. Schröer J, Shenk T. Inhibition of cyclooxygenase activity blocks
celltocell spread of human cytomegalovirus. Proc Natl Acad Sci U.
S. A. 2008;105:1946819473. https://doi.org/10.1073/pnas.0810
740105.
71. Raaben M, Einerhand AW, Taminiau LJ, et al. Cyclooxygenase ac-
tivity is important for efficient replication of mouse hepatitis virus
at an early stage of infection. Virol J. 2007;4:15.
72. Alfajaro MM, Cho EH, Park JG, et al. Feline calicivirusand murine
norovirusinduced COX2/PGE2 signaling pathway has proviral
effects. PloS One. 2018;13:e0200726.
73. Jiang G, Espeseth A, Hazuda DJ, Margolis DM. cMyc and Sp1
contribute to proviral latency by recruiting histone deacetylase 1
to the human immunodeficiency virus type 1 promoter. J Virol.
2007;81:1091410923.
74. Moss ML, Rasmussen FH. Fluorescent substrates for the pro-
teinases ADAM17, ADAM10, ADAM8, and ADAM12 useful
for highthroughput inhibitor screening. Anal Biochem. 2007;366:
144148.
75. Kumar N, Khandelwal N, Kumar R, et al. Inhibitor of sarco/endo-
plasmic reticulum calciumATPase impairs multiple steps of para-
myxovirus replication. Front Microbiol. 2019;10:209. https://doi.
org/10.3389/fmicb.2019.00209.
76. Khandelwal N, Chander Y, Kumar R, et al. Antiviral activity of
Apigenin against Buffalopox: Novel mechanistic insights and drug
resistance considerations. Antivir Res. 2020;181:104870.
77. Thai M, Thaker SK, Feng J, et al. MYCinduced reprogramming of
glutamine catabolism supports optimal virus replication. Nat Com-
mun. 2015;6:19.
78. François S, Sen N, Mitton B, Xiao X, Sakamoto KM, Arvin A. Vari-
cellazoster virus activates CREB, and inhibition of the pCREB
p300/CBP interaction inhibits viral replication in vitro and skin
pathogenesis in vivo. J Virol. 2016;90:86868697.
79. Li X, Du S, Avey D, Li Y, Zhu F, Kuang E. ORF45mediated pro-
longed cFos accumulation accelerates viral transcription during
the late stage of lytic replication of Kaposi's sarcomaassociated
herpesvirus. J Virol. 2015;89:68956906.
80. Li X, Zhang W, Liu Y, Xie J, Hu C, Wang X. Role of p53 in pseu-
dorabies virus replication, pathogenicity, and host immune re-
sponses. Vet Res. 2019;50:9.
81. Robitaille AC, Caron E, Zucchini N, et al. DUSP1 regulates
apoptosis and cell migration, but not the JIP1protected cytokine
response, during respiratory syncytial virus and Sendai virus
12
-
CHANDER
ET AL.
infection. Sci Rep. 2017;7:17388. 17388. https://doi.org/10.1038/
s41598-017-17689-0.
82. Persengiev S, Green M. The role of ATF/CREB family members in
cell growth, survival and apoptosis. Apoptosis. 2003;8:225228.
https://doi.org/10.1023/A:1023633704132.
83. Li W, Sun W, Liu L, et al. IL32: a host proinflammatory factor
against influenza viral replication is upregulated by aberrant
epigenetic modifications during influenza A virus infection.
J Immunol. 2010;185:50565065. https://doi.org/10.4049/jimmun
ol.0902667.
84. Ramsauer K, Sadzak I, Porras A, et al. p38 MAPK enhances STAT1
dependent transcription independently of Ser727 phosphoryla-
tion. Proc. Natl. Acad. Sci. U.S.A. 2002;99:1285912864. https://doi.
org/10.1073/pnas.192264999.
85. Yamauchi S, Takeuchi K, Chihara K, et al. STAT1 is essential for the
inhibition of hepatitis C virus replication by interferon‐λ but not by
interferon‐α.Sci Rep. 2016;6:38336. https://doi.org/10.1038/
srep38336.
86. Mott KR, UnderHill D, Wechsler SL, Town T, Ghiasi H. A role for
the JAKSTAT1 pathway in blocking replication of HSV1 in den-
dritic cells and macrophages. Virol J. 2009;6:56. https://doi.org/
10.1186/1743-422X-6-56.
87. Xu L, Zhou X, Wang W, et al. IFN regulatory factor 1 restricts
hepatitis E virus replication by activating STAT1 to induce antiviral
IFNstimulated genes. FASEB J. 2016;30:33523367. https://doi.
org/10.1096/fj.201600356R.
88. Zhang S, Huo C, Xiao J, et al. pSTAT1 regulates the influenza A
virus replication and inflammatory response in vitro and vivo.
Virology. 2019;537:110120. https://doi.org/10.1016/j.virol.2019.
08.023.
89. Liu L, Cao Z, Chen J, et al. Influenza A virus induces interleukin27
through cyclooxygenase2 and protein kinase A signaling. J Biol
Chem. 2012;287:1189911910. https://doi.org/10.1074/jbc.M111.
308064.
90. Lin CK, Tseng CK, Wu YH, et al. Cyclooxygenase2 facilitates
dengue virus replication and serves as a potential target for
developing antiviral agents. Sci Rep. 2017;7:44701. https://doi.org/
10.1038/srep44701.
91. Raaben M, Einerhand AWC, Taminiau LJA, et al. Cyclooxygenase
activity is important for efficient replication of mouse hepatitis
virus at an early stage of infection. Virol J. 2007;4:55. https://doi.
org/10.1186/1743-422X-4-55.
92. Alfajaro MM, Cho EH, Park JG, et al. Feline calicivirusand mu-
rine norovirusinduced COX2/PGE2 signaling pathway has pro-
viral effects. PLOS ONE. 2018;13:e0200726. https://doi.org/
10.1371/journal.pone.0200726.
93. Alfajaro MM, Choi JS, Kim DS, et al. Activation of COX2/PGE
2
promotes Sapovirus replication via the inhibition of nitric oxide
production. J Virol. 2017;91:e0165601616. https://doi.org/
10.1128/jvi.01656-16.
94. Neininger A, Kontoyiannis D, Kotlyarov A, et al. MK2 targets AU
rich elements and regulates biosynthesis of tumor necrosis factor
and interleukin6 independently at different posttranscriptional
levels. J Biol Chem. 2002;277:30653068. https://doi.org/10.1074/
jbc.C100685200.
95. Herdy B, Karonitsch T, Vladimer GI, et al. The RNAbinding protein
HuR/ELAVL1 regulates IFN‐β mRNA abundance and the type I IFN
response. Eur J Immunol. 2015;45:15001511. https://doi.org/
10.1002/eji.201444979.
96. Amirouche A, Tadesse H, Lunde JA, Bélanger G, Côté J, Jasmin
BJ. Activation of p38 signaling increases utrophin A expression
in skeletal muscle via the RNAbinding protein KSRP and
inhibition of AUrich elementmediated mRNA decay: implica-
tions for novel DMD therapeutics. Hum Mol Genet.
2013;22:30933111.
97. Nagaleekar VK, Sabio G, Aktan I, et al. Translational control of NKT
cell cytokine production by p38 MAPK. J Immunol. 2011;186:4140
4146.
98. Pietersma A, Tilly BC, Gaestel M, et al. p38 mitogen activated
protein kinase regulates endothelial VCAM1 expression at the
posttranscriptional level. Biochem Biophys Res Commun. 1997;
230:4448.
99. Mikkelsen SS, Jensen SB, Chiliveru S, et al. RIGImediated acti-
vation of p38 MAPK is essential for viral induction of interferon
and activation of dendritic cells: dependence on TRAF2 and TAK1.
J Biol Chem. 2009;284:1077410782. https://doi.org/10.1074/jbc.
M807272200.
100. Huang G, Shi LZ, Chi H. Regulation of JNK and p38 MAPK in the
immune system: signal integration, propagation and termination.
Cytokine. 2009;48:161169.
101. Mogensen TH. Pathogen recognition and inflammatory signaling in
innate immune defenses. Clin Microbiol Rev. 2009;22:240273.
https://doi.org/10.1128/CMR.00046-08.
102. Nagata Y, Takahashi N, Davis RJ, Todokoro K. Activation of p38
MAP kinase and JNK but not ERK is required for erythropoietin
induced erythroid differentiation. Blood. 1998;92:18591869.
103. Zhang S, Kaplan MH. The p38 mitogenactivated protein kinase is
required for IL12induced IFN‐γ expression. J Immunol. 2000;
165:13741380.
104. Mathur RK, Awasthi A, Wadhone P, Ramanamurthy B, Saha B.
Reciprocal CD40 signals through p38MAPK and ERK1/2 induce
counteracting immune responses. Nat Med. 2004;10:540544.
105. Jain P, Lavorgna A, Sehgal M, et al. Myocyte enhancer factor
(MEF)2 plays essential roles in Tcell transformation associated
with HTLV1 infection by stabilizing complex between Tax and
CREB. Retrovirology. 2015;12:23. https://doi.org/10.1186/s12977-
015-0140-1.
106. Webb TJR, Litavecz RA, Khan MA, et al. Inhibition of CD1d1
mediated antigen presentation by the vaccinia virus B1R and H5R
molecules. Eur J Immunol. 2006;36:25952600.
107. Renukaradhya GJ, Webb TJR, Khan MA, et al. Virusinduced inhi-
bition of CD1d1mediated antigen presentation: reciprocal regu-
lation by p38 and ERK. J Immunol. 2005;175:43014308.
108. Westrich JA, Vermeer DW, Silva A, et al. CXCL14 suppresses hu-
man papillomavirusassociated head and neck cancer through an-
tigenspecific CD8+Tcell responses by upregulating MHCI
expression. Oncogene. 2019;38:71667180. https://doi.org/
10.1038/s41388-019-0911-6.
109. Smith AL, Ganesh L, Leung K, JongstraBilen J, Jongstra J, Nabel
GJ. Leukocytespecific protein 1 interacts with DCSIGN and me-
diates transport of HIV to the proteasome in dendritic cells. J Exp
Med. 2007;204:421430. https://doi.org/10.1084/jem.20061604.
110. Manley K, O'Hara BA, Gee GV, Simkevich CP, Sedivy JM, Atwood
WJ. NFAT4 is required for JC virus infection of glial cells. J Virol.
2006;80:1207912085. https://doi.org/10.1128/jvi.01456-06.
111. Cicala C, Arthos J, Censoplano N, et al. HIV1 gp120 induces NFAT
nuclear translocation in resting CD4+Tcells. Virology.
2006;345:105114. https://doi.org/10.1016/j.virol.2005.09.052.
112. Nogalski MT, Podduturi JP, DeMeritt IB, Milford LE, Yurochko AD.
The human cytomegalovirus virion possesses an activated casein
kinase II that allows for the rapid phosphorylation of the inhibitor
of NF‐κB, IkappaBalpha. J Virol. 2007;81:53055314. https://doi.
org/10.1128/jvi.02382-06.
113. Gachon F, Thebault S, Peleraux A, Devaux C, Mesnard JM. Mo-
lecular interactions involved in the transactivation of the human T
cell leukemia virus type 1 promoter mediated by Tax and CREB2
(ATF4). Mol Cell Biol. 2000;20:34703481. https://doi.org/
10.1128/mcb.20.10.3470-3481.2000.
114. Gruffat H, Manet E, Sergeant A. MEF2mediated recruitment of
class II HDAC at the EBV immediate early gene BZLF1 links latency
CHANDER ET AL.
-
13
and chromatin remodeling. EMBO Rep. 2002;3:141146. https://
doi.org/10.1093/embo-reports/kvf031.
115. Cheng F, Sawant TV, Lan K, Lu C, Jung JU, Gao SJ. Screening of the
human kinome identifies MSK1/2CREB1 as an essential pathway
mediating Kaposi's sarcomaassociated herpesvirus lytic replica-
tion during primary infection. J Virol. 2015;89:92629280. https://
doi.org/10.1128/jvi.01098-15.
116. François S, Sen N, Mitton B, Xiao X, Sakamoto KM, Arvin A. Vari-
cellaZoster virus activates CREB, and inhibition of the pCREB
p300/CBP interaction inhibits viral replication in vitro and skin
pathogenesis in vivo. J Virol. 2016;90:86868697. https://doi.org/
10.1128/jvi.00920-16.
117. Adamson AL, Darr D, HolleyGuthrie E, et al. EpsteinBarr virus
immediateearly proteins BZLF1 and BRLF1 activate the ATF2
transcription factor by increasing the levels of phosphorylated p38
and cJun Nterminal kinases. J Virol. 2000;74:12241233. https://
doi.org/10.1128/jvi.74.3.1224-1233.2000.
118. Zhao M, New L, Kravchenko VV, et al. Regulation of the MEF2
family of transcription factors by p38. Mol Cell Biol. 1999;19:2130.
https://doi.org/10.1128/mcb.19.1.21.
119. Kumar R, Khandelwal N, Chander Y, et al. MNK1 inhibitor as an
antiviral agent suppresses buffalopox virus protein synthesis.
Antiviral Res. 2018;160:126136. https://doi.org/10.1016/j.
antiviral.2018.10.022.
120. Pichon X, Wilson LA, Stoneley M, et al. RNA binding protein/RNA
element interactions and the control of translation. Curr Protein
Pept Sci. 2012;13:294304. https://doi.org/10.2174/138920312
801619475.
121. Rübsamen D, Blees JS, Schulz K, et al. IRESdependent translation
of egr2 is induced under inflammatory conditions. RNA.
2012;18:19101920. https://doi.org/10.1261/rna.033019.112.
122. Wang X, Huanru W, Li Y, et al. TIA1 and TIAR interact with 5'UTR
of Enterovirus 71 genome and facilitate viral replication. Biochem
Biophys Res Commun. 2015;466. https://doi.org/10.1016/j.bbrc.
2015.09.020.
123. Li W, Li Y, Kedersha N, et al. Cell proteins TIA1 and TIAR interact
with the 30stemloop of the West Nile virus complementary
minusstrand RNA and facilitate virus replication. J Virol.
2002;76:1198912000. https://doi.org/10.1128/jvi.76.23.11989-
12000.2002.
124. Emara MM, Brinton MA. Interaction of TIA1/TIAR with West
Nile and dengue virus products in infected cells interferes with
stress granule formation and processing body assembly. Proc Natl
Acad Sci U. S. A. 2007;104:90419046. https://doi.org/10.1073/
pnas.0703348104.
125. Sun Y, Dong L, Yu S, et al. Newcastle disease virus induces stable
formation of bona fide stress granules to facilitate viral replication
through manipulating host protein translation. FASEB J. 2017;
31:14821493. https://doi.org/10.1096/fj.201600980R.
126. Huang PN, Lin JY, Locker N, et al. Far upstream element binding
protein 1 binds the internal ribosomal entry site of enterovirus 71
and enhances viral translation and viral growth. Nucleic Acids Res.
2011;39:96339648. https://doi.org/10.1093/nar/gkr682.
127. Chien HL, Liao CL, Lin YL. FUSE binding protein 1 interacts with
untranslated regions of Japanese encephalitis virus RNA and
negatively regulates viral replication. J Virol. 2011;85:46984706.
128. Dan X, Wan Q, Yi L, et al. Hsp27 responds to and facilitates
enterovirus A71 replication by enhancing viral internal ribosome
entry sitemediated translation. J Virol. 2019;93:e0232202318.
https://doi.org/10.1128/JVI.02322-18.
129. Liang D, Benko Z, Agbottah E, Bukrinsky M, Zhao RY. AntiVpr
activities of heat shock protein 27. Mol Med. 2007;13:229239.
https://doi.org/10.2119/2007–00004.
130. Ling S, Luo M, Jiang S, et al. Cellular Hsp27 interacts with classical
swine fever virus NS5A protein and negatively regulates viral
replication by the NF‐κB signaling pathway. Virology.
2018;518:202209. https://doi.org/10.1016/j.virol.2018.02.020.
131. Taguwa S, Yeh MT, Rainbolt TK, et al. Zika virus dependence on
host Hsp70 provides a protective strategy against infection and
disease. Cell Rep 2019;26:906920.e903. https://doi.org/10.1016/j.
celrep.2018.12.095.
132. Goodman AG, Smith JA, Balachandran S, et al. The cellular protein
P58IPK regulates influenza virus mRNA translation and replication
through a PKRmediated mechanism. J Virol. 2007;81:22212230.
https://doi.org/10.1128/JVI.02151-06.
133. Trushin SA, AlgecirasSchimnich A, Vlahakis SR, et al. Glycoprotein
120 binding to CXCR4 causes p38dependent primary T cell
death that is facilitated by, but does not require cellassociated
CD4. J Immunol. 2007;178:48464853.
134. Krupkova O, Sadowska A, Kameda T, et al. p38 MaPK Facilitates
crosstalk between endoplasmic reticulum stress and il6 release in
the intervertebral disc. Front Immunol. 2018;9:1706.
135. Geleziunas R, Xu W, Takeda K, Ichijo H, Greene WC. HIV1 Nef
inhibits ASK1dependent death signalling providing a potential
mechanism for protecting the infected host cell. Nature.
2001;410:834838.
136. Ludwig S, Wang X, Ehrhardt C, et al. The influenza A virus NS1
protein inhibits activation of Jun Nterminal kinase and AP1
transcription factors. J virology. 2002;76:1116611171. https://doi.
org/10.1128/jvi.76.21.11166-11171.2002.
137. Wei Y, Jiang Y, Zou F, et al. Activation of PI3K/Akt pathway by
CD133p85 interaction promotes tumorigenic capacity of glioma
stem cells. Proc Natl Acad Sci U. S. A. 2013;110:68296834. https://
doi.org/10.1073/pnas.1217002110.
138. Reddy S, Foreman HC, Sioux T, et al. Ablation of STAT3 in the
B cell compartment restricts gammaherpesvirus latency in vivo.
mBio. 2016;7:e0072300716. https://doi.org/10.1128/mBio.00
723-16.
139. Sen N, Che X, Rajamani J, et al. Signal transducer and activator of
transcription 3 (STAT3) and survivin induction by varicellazoster
virus promote replication and skin pathogenesis. Proc Natl Acad Sci.
U. S. A. 2012;109:600605. https://doi.org/10.1073/pnas.111
4232109.
140. Shisler JL. Viral and cellular FLICEinhibitory proteins: a compari-
son of their roles in regulating intrinsic immune responses. J Virol.
2014;88:65396541. https://doi.org/10.1128/JVI.00276-14.
141. Chaudhary PM, Jasmin A, Eby MT, Hood L. Modulation of the NF
kappa B pathway by virally encoded death effector domainscon-
taining proteins. Oncogene. 1999;18:57385746. https://doi.org/
10.1038/sj.onc.1202976.
142. Frost JR, Mendez M, Soriano AM, et al. Adenovirus 5 E1Amedi-
ated suppression of p53 via FUBP1. J Virol. 2018;92:e00439
00418. https://doi.org/10.1128/jvi.00439-18.
143. Padhan K, Minakshi R, Towheed MAB, Jameel S. Severe acute
respiratory syndrome coronavirus 3a protein activates the mito-
chondrial death pathway through p38 MAP kinase activation. J Gen
Virol. 2008;89:19601969.
144. Zhu N, Shao Y, Xu L, Yu L, Sun L. Gadd45‐α and Gadd45‐γ utilize
p38 and JNK signaling pathways to induce cell cycle G2/M arrest in
HepG2 hepatoma cells. Mol Biol Rep. 2009;36:2075.
145. Liang Z, Liu R, Zhang H, et al. GADD45 proteins inhibit HIV1
replication through specific suppression of HIV1 transcription.
Virology. 2016;493:111. https://doi.org/10.1016/j.virol.2016.02.
014. DOI:
146. Ye M, Duus KM, Peng J, Price DH, Grose C. Varicellazoster virus
Fc receptor component gI is phosphorylated on its endodomain by
a cyclindependent kinase. J Virol. 1999;73:13201330.
147. Kitay MK, Rowe DT. Cell cycle stagespecific phosphorylation of
the EpsteinBarr virus immortalization protein EBNALP. J Virol.
1996;70:78857893.
14
-
CHANDER
ET AL.
148. Zafrullah M, Ozdener MH, Panda SK, Jameel S. The ORF3 protein
of hepatitis E virus is a phosphoprotein that associates with the
cytoskeleton. J Virol. 1997;71:90459053.
149. Advani SJ, Weichselbaum RR, Roizman B. The role of cdc2 in the
expression of herpes simplex virus genes. Proc Natl Acad Sci. U. S. A.
2000;97:1099611001.
150. Smith EM, Proud CG. cdc2–cyclin B regulates eEF2 kinase activity
in a cell cycleand amino aciddependent manner. EMBO J.
2008;27:10051016.
151. ValienteEcheverría F, Melnychuk L, Vyboh K, et al. eEF2 and Ras
GAP SH3 domainbinding protein (G3BP1) modulate stress granule
assembly during HIV1 infection. Nat Commun. 2014;5:4819.
https://doi.org/10.1038/ncomms5819.
152. Poggioli GJ, Dermody TS, Tyler KL. Reovirusinduced ς1sdepen-
dent G
2
/M phase cell cycle arrest is associated with inhibition of
p34
cdc2
.J Virol. 2001;75:74297434. https://doi.org/10.1128/
jvi.75.16.7429-7434.2001.
153. Fournier N, Raj K, Saudan P, et al. Expression of human papillo-
mavirus 16 E2 protein in Schizosaccharomyces pombe delays the
initiation of mitosis. Oncogene. 1999;18:40154021.
154. Re F, Braaten D, Franke EK, Luban J. Human immunodeficiency
virus type 1 Vpr arrests the cell cycle in G2 by inhibiting the
activation of p34cdc2cyclin B. J Virol. 1995;69:68596864.
155. Mohl BP, Roy P. Cellular casein kinase 2 and protein phosphatase
2A modulate replication site assembly of Bluetongue virus. J Biol
Chem. 2016;291:1456614574. https://doi.org/10.1074/jbc.M116.
714766.
156. Wang X, Grammatikakis N, Hu J. Role of p50/CDC37 in hep-
adnavirus assembly and replication. J Biol Chem. 2002;277:24361
24367. https://doi.org/10.1074/jbc.M202198200.
157. Xu K, Nagy PD. Enrichment of phosphatidylethanolamine in viral
replication compartments via Coopting the endosomal Rab5 small
GTPase by a positivestrand RNA virus. PLoS Biol. 2016;14:
e2000128. https://doi.org/10.1371/journal.pbio.2000128.
158. Cavalli V, Vilbois F, Corti M, et al. The stressinduced MAP kinase
p38 regulates endocytic trafficking via the GDI: Rab5 complex. Mol
Cell. 2001;7:421432.
159. Nielsen E, Christoforidis S, UttenweilerJoseph S, et al. Rabenosyn
5, a novel Rab5 effector, is complexed with hVPS45 and recruited
to endosomes through a FYVE finger domain. J Cell Biol.
2000;151:601612.
160. Marchant D, Singhera GK, Utokaparch S, et al. Tolllike receptor 4
mediated activation of p38 mitogenactivated protein kinase is a
determinant of respiratory virus entry and tropism. J Virol.
2010;84:1135911373. https://doi.org/10.1128/jvi.00804-10.
161. Acosta EG, Castilla V, Damonte EB. Differential requirements in
endocytic trafficking for penetration of dengue virus. PloS One.
2012;7:e44835.
162. Krishnan MN, Sukumaran B, Pal U, et al. Rab 5 is required for the
cellular entry of dengue and West Nile viruses. J Virol.
2007;81:48814885.
163. Lin J, Wang C, Zhang L, et al. Rab5 enhances classical swine fever
virus proliferation and interacts with viral NS4B protein to facili-
tate formation of NS4B related complex. Front Microbiol.
2017;8:1468. 1468. https://doi.org/10.3389/fmicb.2017.01468.
164. Stone M, Jia S, Heo WD, Meyer T, Konan KV. Participation of Rab5,
an early endosome protein, in hepatitis C virus RNA replication
machinery. J Virol. 2007;81:45514563. https://doi.org/10.1128/
jvi.01366-06.
165. Meinl E, Lengenfelder D, Blank N, Pirzer R, Barata L, Hivroz C.
Differential requirement of ZAP70 for CD2mediated activation
pathways of mature human T cells. J Immunol. 2000;165:3578
3583.
166. SolFoulon N, Sourisseau M, Porrot F, et al. ZAP70 kinase regu-
lates HIV celltocell spread and virological synapse formation.
EMBO J. 2007;26:516526. https://doi.org/10.1038/sj.emboj.
7601509.
167. Arenaccio C, Chiozzini C, ColumbaCabezas S, et al. Exosomes from
human immunodeficiency virus type 1 (HIV1)infected cells li-
cense quiescent CD4+T lymphocytes to replicate HIV1 through a
Nefand ADAM17dependent mechanism. J Virol. 2014;88:11529
11539. https://doi.org/10.1128/JVI.01712-14.
168. Mikuličić S, Finke J, Boukhallouk F, et al. ADAM17dependent
signaling is required for oncogenic human papillomavirus entry
platform assembly. eLife. 2019;8:e44345. https://doi.org/10.7554/
eLife.44345.
169. Hill CS, Wynne J, Treisman R. The Rho family GTPases RhoA, Racl,
and CDC42Hs regulate transcriptional activation by SRF. Cell.
1995;81:11591170.
170. Wong J, Zhang J, Yanagawa B, et al. Cleavage of serum response
factor mediated by enteroviral protease 2A contributes to
impaired cardiac function. Cell Res. 2011;22:360371. https://doi.
org/10.1038/cr.2011.114.
171. Khandelwal N, Chander Y, Rawat KD, et al. Emetine inhibits
replication of RNA and DNA viruses without generating drug
resistant virus variants. Antiviral Res. 2017;144:196204. https://
doi.org/10.1016/j.antiviral.2017.06.006.
172. Kumar N, Sharma NR, Ly H, Parslow TG, Liang Y. Receptor tyrosine
kinase inhibitors that block replication of influenza a and other
viruses. Antimicrob Agents Chemother. 2011;55:55535559. https://
doi.org/10.1128/AAC.00725-11.
173. Zhang J, Yang PL, Gray NS. Targeting cancer with small molecule
kinase inhibitors. Nat Rev Cancer. 2009;9:2839. https://doi.org/
10.1038/nrc2559.
174. Player MR. Protein kinase inhibitors for the treatment of inflam-
matory disease. Curr Top Med Chem. 2009;9:598.
175. Kumar N, Liang Y, Parslow TG, Liang Y. Receptor tyrosine
kinase inhibitors block multiple steps of influenza a virus replica-
tion. J Virol. 2011;85:28182827. https://doi.org/10.1128/JVI.01
969-10.
176. Sharma S, Sundararajan A, Suryawanshi A, et al. T cell immuno-
globulin and mucin protein3 (Tim3)/Galectin9 interaction
regulates influenza A virusspecific humoral and CD8 Tcell re-
sponses. Proc Natl Acad Sci U. S. A. 2011;108:1900119006. https://
doi.org/10.1073/pnas.1107087108.
177. Sharma S, Mulik S, Kumar N, Suryawanshi A, Rouse BT. An anti
inflammatory role of VEGFR2/Src kinase inhibitor in herpes sim-
plex virus 1induced immunopathology. J Virol. 2011;85:5995
6007. https://doi.org/10.1128/JVI.00034-11.
178. Bouhaddou M, Memon D, Meyer B, et al. The global phosphory-
lation landscape of SARSCoV2 infection. Cell. 2020;182:685712.
e619
179. Kono M, Tatsumi K, Imai AM, Saito K, Kuriyama T, Shirasawa H.
Inhibition of human coronavirus 229E infection in human epithelial
lung cells (L132) by chloroquine: involvement of p38 MAPK and
ERK. Antiviral Res. 2008;77:150152.
180. Sreekanth GP, Chuncharunee A, Sirimontaporn A, et al. SB203580
modulates p38 MAPK signaling and dengue virusinduced liver
injury by reducing MAPKAPK2, HSP27, and ATF2 phosphorylation.
PloS One. 2016;11:e0149486.
181. Fu Y, Yip A, Seah PG, Blasco F, Shi PY, Hervé M. Modulation of
inflammation and pathology during dengue virus infection by p38
MAPK inhibitor SB203580. Antiviral Res. 2014;110:151157.
182. Zhang Z, Wang B, Wu S, et al. PD169316, a specific p38 inhibitor,
shows antiviral activity against enterovirus71. Virology.
2017;508:150158.
183. Johnson JC, Martinez O, Honko AN, Hensley LE, Olinger GG,
Basler CF. Pyridinyl imidazole inhibitors of p38 MAP kinase impair
viral entry and reduce cytokine induction by Zaire ebolavirus in
human dendritic cells. Antiviral Res. 2014;107:102109.
CHANDER ET AL.
-
15
184. Dai JP, Wang QW, Su Y, et al. Oxymatrine inhibits influenza A
virus replication and inflammation via TLR4, p38 MAPK and NF‐κB
pathways. Int J Mol Sci. 2018;19:965.
185. Orphanos GS, Ioannidis GN, Ardavanis AG. Cardiotoxicity induced
by tyrosine kinase inhibitors. Acta Oncol 2009;48:964970. https://
doi.org/10.1080/02841860903229124.
186. Shah DR, Shah RR, Morganroth J. Tyrosine kinase inhibitors:
their ontarget toxicities as potential indicators of efficacy.
Drug Saf 2013;36:413426. https://doi.org/10.1007/s40264-013-
0050-x.
187. Zeisel MB, Lupberger J, Fofana I, Baumert TF. Hosttargeting
agents for prevention and treatment of chronic hepatitis C per-
spectives and challenges. J Hepatol. 2013;58:375384. https://doi.
org/10.1016/j.jhep.2012.09.022.
188. Anand P, Shenoy R, Palmer JE, et al. Clinical trial of the p38 MAP
kinase inhibitor dilmapimod in neuropathic pain following nerve
injury. Eur J Pain. 2011;15:10401048. https://doi.org/10.1016/j.
ejpain.2011.04.005.
189. Hopcraft SE, Evans MJ. Selection of a hepatitis C virus with altered
entry factor requirements reveals a genetic interaction between
the E1 glycoprotein and claudins. Hepatology. 2015;62:10591069.
https://doi.org/10.1002/hep.27815.
190. Puyang X, Poulin DL, Mathy JE, et al. Mechanism of resistance of
hepatitis C virus replicons to structurally distinct cyclophilin
inhibitors. Antimicrob Agents Chemother. 2010;54:19811987.
https://doi.org/10.1128/AAC.01236-09.
191. Wilson TR, Fridlyand J, Yan Y, et al. Widespread potential for
growthfactordriven resistance to anticancer kinase inhibitors.
Nature. 2012;487:505509. https://doi.org/10.1038/nature11249.
192. Sato Y, Tsurumi T. Genome guardian p53 and viral infections. Re-
views in Medical Virology. 2013;23 (4):213–220. http://dx.doi.org/
10.1002/rmv.1738.
193. Turpin E, Luke K, Jones J, et al. Influenza Virus Infection Increases
p53 Activity: Role of p53 in Cell Death and 10.1128/
JVI.79.14.88028811. Viral Replication. Journal of virology 2005,
79: 88028811.
194. Nishitoh H. CHOP is a multifunctional transcription factor in the
ER stress response. Journal of Biochemistry. 2012;151 (3):217–219.
http://dx.doi.org/10.1093/jb/mvr143.
How to cite this article: Chander Y, Kumar R, Khandelwal N,
et al. Role of p38 mitogenactivated protein kinase signalling
in virus replication and potential for developing broad
spectrum antiviral drugs. Rev Med Virol. 2021;116. https://
doi.org/10.1002/rmv.2217
16
-
CHANDER
ET AL.
... Almendros et al. (2022) claim that this mechanism triggers a number of pathogenic signalling pathways that lead to serious cardiovascular diseases. Extracellular signalregulated kinase (ERK1/2), a subfamily of mitogen activated protein kinase (MAPK), has been suggested as a viable target among protein kinases that regulate cellular transcriptional activity, cell proliferation, differentiation, and cell survival ( Chander et al., 2021 ). The ERK1/2 cascade is crucial for myocardial protection against ischemia reperfusion damage, according to reports by Chander et al. (2021) and Ronkina and Gaestel (2022) . ...
... Extracellular signalregulated kinase (ERK1/2), a subfamily of mitogen activated protein kinase (MAPK), has been suggested as a viable target among protein kinases that regulate cellular transcriptional activity, cell proliferation, differentiation, and cell survival ( Chander et al., 2021 ). The ERK1/2 cascade is crucial for myocardial protection against ischemia reperfusion damage, according to reports by Chander et al. (2021) and Ronkina and Gaestel (2022) . Inflammatory responses that are brought on when this enzyme's activities are dysregulated greatly enhance apoptotic signalling and cell death ( Ally et al., 2020 ). ...
Article
The inhibition of p38 mitogen-activated protein kinase (p38-MAPK) by small molecule chemical inhibitors was previously shown to impair severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) replication, however, mechanisms underlying antiviral activity remains unexplored. In this study, reduced growth of SARS-CoV-2 in p38-α knockout Vero cells, together with enhanced viral yield in cells transfected with construct expressing p38α, suggested that p38-MAPK is essential for the propagation of SARS-CoV-2. The SARS-CoV-2 was also shown to induce phosphorylation (activation) of p38, at time when transcription/translational activities are considered to be at the peak levels. Further, we demonstrated that p38 supports viral RNA/protein synthesis without affecting viral attachment, entry, and budding in the target cells. In conclusion, we provide mechanistic insights on the regulation of SARS-CoV-2 replication by p38 MAPK.
Article
In this study, we demonstrated the antiviral efficacy of hesperetin against multiple poxviruses, including buffalopox virus (BPXV), vaccinia virus (VACV), and lumpy skin disease virus (LSDV). The time‐of‐addition and virus step‐specific assays indicated that hesperetin reduces the levels of viral DNA, mRNA, and proteins in the target cells. Further, by immunoprecipitation (IP) of the viral RNA from BPXV‐infected Vero cells and a cell‐free RNA‐IP assay, we demonstrated that hesperetin‐induced reduction in BPXV protein synthesis is also consistent with diminished interaction between eukaryotic translation initiation factor eIF4E and the 5′ cap of viral mRNA. Molecular docking and MD simulation studies were also consistent with the binding of hesperetin to the cap‐binding pocket of eIF4E, adopting a conformation similar to m7GTP binding. Furthermore, in a BPXV egg infection model, hesperetin was shown to suppress the development of pock lesions on the chorioallantoic membrane and associated mortality in the chicken embryos. Most importantly, long‐term culture of BPXV in the presence of hesperetin did not induce the generation of drug‐resistant viral mutants. In conclusion, we, for the first time, demonstrated the antiviral activity of hesperetin against multiple poxviruses, besides providing some insights into its potential mechanisms of action.
Article
Full-text available
Equine herpesvirus type 8 (EHV-8) causes abortion and respiratory disease in horses and donkeys, leading to serious economic losses in the global equine industry. Currently, there is no effective vaccine or drug against EHV-8 infection, underscoring the need for a novel antiviral drug to prevent EHV-8-induced latent infection and decrease the pathogenicity of this virus. The present study demonstrated that hyperoside can exert antiviral effects against EHV-8 infection in RK-13 (rabbit kidney cells), MDBK (Madin–Darby bovine kidney), and NBL-6 cells (E. Derm cells). Mechanistic investigations revealed that hyperoside induces heme oxygenase-1 expression by activating the c-Jun N-terminal kinase/nuclear factor erythroid-2-related factor 2/Kelch-like ECH-associated protein 1 axis, alleviating oxidative stress and triggering a downstream antiviral interferon response. Accordingly, hyperoside inhibits EHV-8 infection. Meanwhile, hyperoside can also mitigate EHV-8-induced injury in the lungs of infected mice. These results indicate that hyperoside may serve as a novel antiviral agent against EHV-8 infection. IMPORTANCE Hyperoside has been reported to suppress viral infections, including herpesvirus, hepatitis B virus, infectious bronchitis virus, and severe acute respiratory syndrome coronavirus 2 infection. However, its mechanism of action against equine herpesvirus type 8 (EHV-8) is currently unknown. Here, we demonstrated that hyperoside significantly inhibits EHV-8 adsorption and internalization in susceptible cells. This process induces HO-1 expression via c-Jun N-terminal kinase/nuclear factor erythroid-2-related factor 2/Kelch-like ECH-associated protein 1 axis activation, alleviating oxidative stress and triggering an antiviral interferon response. These findings indicate that hyperoside could be very effective as a drug against EHV-8.
Article
Full-text available
Although the COVID-19 pandemic appears to be diminishing, the emergence of SARS-CoV-2 variants represents a threat to humans due to their inherent transmissibility, immunological evasion, virulence, and invulnerability to existing therapies. The COVID-19 pandemic affected more than 500 million people and caused over 6 million deaths. Vaccines are essential, but in circumstances in which vaccination is not accessible or in individuals with compromised immune systems, drugs can provide additional protection. Targeting host signaling pathways is recommended due to their genomic stability and resistance barriers. Moreover, targeting host factors allows us to develop compounds that are effective against different viral variants as well as against newly emerging virus strains. In recent years, the globe has experienced climate change, which may contribute to the emergence and spread of infectious diseases through a variety of factors. Warmer temperatures and changing precipitation patterns can increase the geographic range of disease-carrying vectors, increasing the risk of diseases spreading to new areas. Climate change may also affect vector behavior, leading to a longer breeding season and more breeding sites for disease vectors. Climate change may also disrupt ecosystems, bringing humans closer to wildlife that transmits zoonotic diseases. All the above factors may accelerate the emergence of new viral epidemics. Plant-derived products, which have been used in traditional medicine for treating pathological conditions, offer structurally novel therapeutic compounds, including those with anti-viral activity. In addition, plant-derived bioactive substances might serve as the ideal basis for developing sustainable/efficient/cost-effective anti-viral alternatives. Interest in herbal antiviral products has increased. More than 50% of approved drugs originate from herbal sources. Plant-derived compounds offer diverse structures and bioactive molecules that are candidates for new drug development. Combining these therapies with conventional drugs could improve patient outcomes. Epigenetics modifications in the genome can affect gene expression without altering DNA sequences. Host cells can use epigenetic gene regulation as a mechanism to silence incoming viral DNA molecules, while viruses recruit cellular epitranscriptomic (covalent modifications of RNAs) modifiers to increase the translational efficiency and transcript stability of viral transcripts to enhance viral gene expression and replication. Moreover, viruses manipulate host cells’ epigenetic machinery to ensure productive viral infections. Environmental factors, such as natural products, may influence epigenetic modifications. In this review, we explore the potential of plant-derived substances as epigenetic modifiers for broad-spectrum anti-viral activity, reviewing their modulation processes and anti-viral effects on DNA and RNA viruses, as well as addressing future research objectives in this rapidly emerging field.
Article
Full-text available
The Epstein-Barr virus (EBV) efficiently transforms primary B cells. Here, we show that this process starts immediately after cellular exposure to infectious viral particles. Virus binding to B cells led to the activation of intracytoplasmic tyrosine kinases and STAT3. Tegument proteins within the virion in turn activated the p38-MK2 pathway upon cell entry, independently of the viral DNA. Engagement of STAT3 and p38/MK2, two pro-inflammatory pathways, was essential for expression of the key EBV transforming gene EBNA2 but also facilitated IL-6 and TNFα release. However, these pathways simultaneously activated ZFP36L1, a stress response protein that targets transcripts with an AU-rich 3′UTR, to reduce IL-6 and TNFα transcription in infected cells. Expression of viral latent proteins after infection amplified the viral effects on p38 and MK2, but also on ZFP36L1, altogether resulting in a transitory and limited increase in IL-6 and TNFα transcription and release. Thus, EBV virions are not merely vehicles that allow injection of the viral DNA into the nucleus but manipulate cellular pathways to initiate transformation while limiting cytokine release. IMPORTANCE The Epstein-Barr virus efficiently infects and transforms B lymphocytes. During this process, infectious viral particles transport the viral genome to the nucleus of target cells. We show here that these complex viral structures serve additional crucial roles by activating transcription of the transforming genes encoded by the virus. We show that components of the infectious particle sequentially activate proinflammatory B lymphocyte signaling pathways that, in turn, activate viral gene expression but also cause cytokine release. However, virus infection activates expression of ZFP36L1, an RNA-binding stress protein that limits the length and the intensity of the cytokine response. Thus, the infectious particles can activate viral gene expression and initiate cellular transformation at the price of a limited immune response.
Article
Hepatitis B virus (HBV) infection is a serious global health problem that threatens the health of human. Tannic acid (TA), a natural polyphenol in foods, fruits, and plants, exhibits a variety of bioactive functions. In our research, we decide to explore the pharmacological mechanism of TA against HBV replication. Our results showed that TA effectively reduced the content of HBV DNA and viral antigens (HBsAg and HBeAg) in HepG2.2.15 cells. Meanwhile, TA significantly decreased the mRNA expression of HBV RNA, which include total HBV RNA, HBV pregenomic RNA, and HBV precore mRNA. Besides, TA evidently downregulated the activity of HBV promoters in HepG2.2.15 cells. Furthermore, we found that TA upregulated the expression of IL-8, TNF-α, IFN-α, and IFN-α-mediated antiviral effectors in HepG2.2.15 cells. On the contrary, TA downregulated the expression of IL-10 and hepatic nuclear factor 4 (HNF4α). In addition, TA activated the NF-κB and MAPK pathways that contributed to the inhibition of HBV replication. Finally, TA treatment led to the occurrence of autophagy, which accelerated the elimination of HBV components in HepG2.2.15 cells. Taken together, our results elucidated the suppressive effect of TA on HBV replication and provided inspiration for its clinical application in HBV treatment.
Article
Introduction: The global Mpox (MPX) disease outbreak caused by the Mpox virus (MPXV) in 2022 alarmed the World Health Organization (WHO) and health regulation agencies of individual countries leading to the declaration of MPX as a Public Health Emergency. Owing to the genetic similarities between smallpox-causing poxvirus and MPXV, vaccine JYNNEOS, and anti-smallpox drugs brincidofovir and tecovirimat were granted emergency use authorization by the United States Food and Drug Administration. The WHO also included cidofovir, NIOCH-14, and other vaccines as treatment options. Areas covered: This article covers the historical development of EUA-granted antivirals, resistance to these antivirals, and the projected impact of signature mutations on the potency of antivirals against currently circulating MPXV. Since a high prevalence of MPXV infections in individuals coinfected with HIV and MPXV, the treatment results among these individuals have been included. Expert opinion: All EUA-granted drugs have been approved for smallpox treatment. These antivirals show good potency against Mpox. However, conserved resistance mutation positions in MPXV and related poxviruses, and the signature mutations in the 2022 MPXV can potentially compromise the efficacy of the EUA-granted treatments. Therefore, MPXV-specific medications are required not only for the current but also for possible future outbreaks.
Article
Full-text available
We describe herein that Apigenin, which is a dietary flavonoid, exerts a strong in vitro and in ovo antiviral efficacy against buffalopox virus (BPXV). Apigenin treatment was shown to inhibit synthesis of viral DNA, mRNA and proteins, without affecting other steps of viral life cycle such as attachment, entry and budding. Although the major mode of antiviral action of Apigenin was shown to be mediated via targeting certain cellular factors, a modest inhibitory effect of Apigenin was also observed directly on viral polymerase. We also evaluated the selection of drug-resistant virus variants under long-term selection pressure of Apigenin. Wherein Apigenin-resistant mutants were not observed up to ~ P20 (passage 20), a significant resistance was observed to the antiviral action of Apigenin at ~ P30. However, a high degree resistance could not be observed even up to P60. To the best of our knowledge, this is the first report describing in vitro and in ovo antiviral efficacy of Apigenin against poxvirus infection. The study also provides mechanistic insights on the antiviral activity of Apigenin and selection of potential Apigenin-resistant mutants upon long-term culture.
Article
Full-text available
Merkel cell carcinoma (MCC) is an aggressive skin cancer with high rates of recurrence and metastasis. Merkel cell polyomavirus (MCPyV) is associated with the majority of MCC cases. MCPyV-induced tumourigenesis is largely dependent on the expression of the small tumour antigen (ST). Recent findings implicate MCPyV ST expression in the highly metastatic nature of MCC by promoting cell motility and migration, through differential expression of cellular proteins that lead to microtubule destabilisation, filopodium formation and breakdown of cell-cell junctions. However, the molecular mechanisms which dysregulate these cellular processes are yet to be fully elucidated. Here we demonstrate that MCPyV ST expression activates p38 MAPK signalling to drive cell migration and motility. Notably, MCPyV ST-mediated p38 MAPK signalling occurs through MKK4, as opposed to the canonical MKK3/6 signalling pathway. In addition, our results indicate that an interaction between MCPyV ST and the cellular phospatase subunit PP4C is essential for its effect on p38 MAPK signalling. These results provide novel opportunities for the treatment of metastatic MCC given the intense interest in p38 MAPK inhibitors as therapeutic agents.
Article
Full-text available
Newcastle disease virus (NDV), a member of the Paramyxoviridae family, can activate PKR/eIF2α signaling cascade to shutoff host and facilitate viral mRNA translation during infection, however, the mechanism remains unclear. In this study, we revealed that NDV infection up-regulated host cap-dependent translation machinery by activating PI3K/Akt/mTOR and p38 MAPK/Mnk1 pathways. In addition, NDV infection induced p38 MAPK/Mnk1 signaling participated 4E-BP1 hyperphosphorylation for efficient viral protein synthesis when mTOR signaling is inhibited. Furthermore, NDV NP protein was found to be important for selective cap-dependent translation of viral mRNAs through binding to eIF4E during NDV infection. Taken together, NDV infection activated multiple signaling pathways for selective viral protein synthesis in infected cells, via interaction between viral NP protein and host translation machinery. Our results may help to design novel targets for therapeutic intervention against NDV infection and to understand the NDV anti-oncolytic mechanism.
Article
Full-text available
Antiviral drugs have traditionally been developed by directly targeting essential viral components. However, this strategy often fails due to the rapid generation of drug-resistant viruses. Recent genome-wide approaches, such as those employing small interfering RNA (siRNA) or clustered regularly interspaced short palindromic repeats (CRISPR) or those using small molecule chemical inhibitors targeting the cellular “kinome,” have been used successfully to identify cellular factors that can support virus replication. Since some of these cellular factors are critical for virus replication, but are dispensable for the host, they can serve as novel targets for antiviral drug development. In addition, potentiation of immune responses, regulation of cytokine storms, and modulation of epigenetic changes upon virus infections are also feasible approaches to control infections. Because it is less likely that viruses will mutate to replace missing cellular functions, the chance of generating drug-resistant mutants with host-targeted inhibitor approaches is minimized. However, drug resistance against some host-directed agents can, in fact, occur under certain circumstances, such as long-term selection pressure of a host-directed antiviral agent that can allow the virus the opportunity to adapt to use an alternate host factor or to alter its affinity toward the target that confers resistance. This review describes novel approaches for antiviral drug development with a focus on host-directed therapies and the potential mechanisms that may account for the acquisition of antiviral drug resistance against host-directed agents.
Article
Full-text available
Evasion of the host immune responses is critical for both persistent human papillomavirus (HPV) infection and associated cancer progression. We have previously shown that expression of the homeostatic chemokine CXCL14 is significantly downregulated by the HPV oncoprotein E7 during cancer progression. Restoration of CXCL14 expression in HPV-positive head and neck cancer (HNC) cells dramatically suppresses tumor growth and increases survival through an immune-dependent mechanism in mice. Although CXCL14 recruits natural killer (NK) and T cells to the tumor microenvironment, the mechanism by which CXCL14 mediates tumor suppression through NK and/or T cells remained undefined. Here we report that CD8+ T cells are required for CXCL14-mediated tumor suppression. Using a CD8+ T-cell receptor transgenic model, we show that the CXCL14-mediated antitumor CD8+ T-cell responses require antigen specificity. Interestingly, CXCL14 expression restores major histocompatibility complex class I (MHC-I) expression on HPV-positive HNC cells downregulated by HPV, and knockdown of MHC-I expression in HNC cells results in loss of tumor suppression even with CXCL14 expression. These results suggest that CXCL14 enacts antitumor immunity through restoration of MHC-I expression on tumor cells and promoting antigen-specific CD8+ T-cell responses to suppress HPV-positive HNC.
Article
We describe herein that Apigenin, which is a dietary flavonoid, exerts a strong in vitro and in ovo antiviral efficacy against buffalopox virus (BPXV). Apigenin treatment was shown to inhibit synthesis of viral DNA, mRNA and proteins, without affecting other steps of viral life cycle such as attachment, entry and budding. Although the major mode of antiviral action of Apigenin was shown to be mediated via targeting certain cellular factors, a modest inhibitory effect of Apigenin was also observed directly on viral polymerase. We also evaluated the selection of drug-resistant virus variants under long-term selection pressure of Apigenin. Wherein Apigenin-resistant mutants were not observed up to ∼P20 (passage 20), a significant resistance was observed to the antiviral action of Apigenin at ∼P30. However, a high degree resistance could not be observed even up to P60. To the best of our knowledge, this is the first report describing in vitro and in ovo antiviral efficacy of Apigenin against poxvirus infection. The study also provides mechanistic insights on the antiviral activity of Apigenin and selection of potential Apigenin-resistant mutants upon long-term culture.
Article
Cytokine storm is a form of uncontrolled systemic inflammatory reaction activated by a variety of factors and leading to a harmful homeostatic process, even to patient’s death. Triggers that start the reaction are infection, systemic diseases and rarely anaphylaxis. Cytokine storm is frequently mentioned in connection to medical interventions such as transplantation or administration of drugs. Presented mini-review would like to show current possibilities how to fight or even stop such a life-threatening, immune-mediated process in order to save lives, not only in COVID-19 patients. Early identification of rising state and multilevel course of treatment is imperative. The most widely used molecule for systemic treatment remains tocilizumab. Except for anti IL-6 treatment, contemporary research opens the possibilities for combination of pharmaceutical, non-pharmaceutical and adjunctive treatment in a successful fight with consequences of cytokine storm. Further work is needed to discover the exact signaling pathways that lead to cytokine storm and to determine how these effector molecules and/or combination of processes can help to resolve this frequently fatal episode of inflammation. It is a huge need for all scientists and clinicians to establish a physiological rational for new therapeutic targets that might lead to more personalized medicine approaches.
Article
The causative agent of the coronavirus disease 2019 (COVID-19) pandemic, severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has infected millions and killed hundreds of thousands of people worldwide, highlighting an urgent need to develop antiviral therapies. Here, we present a quantitative mass spectrometry-based phosphoproteomics survey of SARS-CoV-2 infection in Vero E6 cells, revealing dramatic rewiring of phosphorylation on host and viral proteins. SARS-CoV-2 infection promoted casein kinase II (CK2) and p38 MAP kinase activation, production of diverse cytokines, and shutdown of mitotic kinases resulting in cell cycle arrest. Infection also stimulated a marked induction of CK2-containing filopodia protrusions possessing budding viral particles. Eighty-seven drugs and compounds were identified by mapping global phosphorylation profiles to dysregulated kinases and pathways. We found pharmacologic inhibition of p38, CK2, CDKs, AXL and PIKFYVE kinases to possess antiviral efficacy, representing potential COVID-19 therapies.
Article
p38 MAP kinase (p38) and JNK have been described as playing a critical role in the response to a variety of environmental stresses and proinflammatory cytokines. It was recently reported that hematopoietic cytokines activate not only classical MAP kinases (ERK), but also p38 and JNK. However, the physiological function of these kinases in hematopoiesis remains obscure. We found that all MAP kinases examined, ERK1, ERK2, p38, JNK1, and JNK2, were rapidly and transiently activated by erythropoietin (Epo) stimulation in SKT6 cells, which can be induced to differentiate into hemoglobinized cells in response to Epo. Furthermore, p38-specific inhibitor SB203580 but not MEK-specific inhibitor PD98059 significantly suppressed Epo-induced differentiation and antisense oligonucleotides of p38, JNK1, and JNK2, but neither ERK1 nor ERK2 clearly inhibited Epo-induced hemoglobinization. However, in Epo-dependent FD-EPO cells, inhibition of either ERKs, p38, or JNKs suppressed cell growth. Furthermore, forced expression of a gain-of-function MKK6 mutant, which specifically activated p38, induced hemoglobinization of SKT6 cells without Epo. These results indicate that activation of p38 and JNKs but not of ERKs is required for Epo-induced erythroid differentiation of SKT6 cells, whereas all of these kinases are involved in Epo-induced mitogenesis of FD-EPO cells. © 1998 by The American Society of Hematology.
Article
Influenza A virus infection activates various intracellular signaling pathways, which is mediated by the transcription factors. Here, a quantitative phosphoproteomic analysis of A549 cells after infection with influenza A virus (H5N1) was performed and we found that the transcription factor STAT1 was highly activated. Unexpectedly, upon inhibition of p-STAT1, titers of progeny virus and viral protein synthesis were both reduced. The STAT1 inhibitor Fludarabine (FLUD) inhibited an early progeny step in viral infection and reduced the levels of influenza virus genomic RNA (vRNA). Concomitantly, there was reduced expression of inflammatory cytokines in p-STAT1 inhibited cells. In vivo, suppression of p-STAT1 improved the survival of H5N1 virus-infected mice, reduced the pulmonary inflammatory response and viral burden. Thus, our data demonstrated a critical role for p-STAT1 in influenza virus replication and inflammatory responses. We speculate that STAT1 is an example of a putative antiviral signaling component to support effective replication.