Seminario Virus
Seminario Virus
Seminario Virus
Review
5
Box 1. Structure and functions of a mitochondrion National University Cancer Institute,
National University Health System,
Mitochondria are membrane-bound organelles found inside most eukaryotic cells. Referred to as the powerhouse of the
Singapore
cell, the primary function of a mitochondrion is to generate energy in the form of ATP by oxidative phosphorylation
(OXPHOS). Structurally, mitochondria have two distinct membranes, namely outer and inner membranes, that define three
compartments. The inner most compartment or mitochondrial matrix is where numerous enzymatic reactions occur,
including replication and transcription of mitochondrial DNA (mtDNA), protein translation, the citric acid cycle, and fatty acid
catabolism. The inner mitochondrial membrane (IMM) is permeable only to oxygen, carbon dioxide, and water. The IMM
forms cristae that invaginate into the matrix, and contains the electron transport chain (ETC) and the ATP synthase, *Correspondence:
responsible for OXPHOS. During this process, electrons are shuttled from the initial electron donor (NADH or FADH2) micas@nus.edu.sg (S. Alonso).
through a series of electron carriers to the final acceptor, molecular oxygen (O2), eventually resulting in the production of
ATP. The electron carrier proteins are embedded in the IMM; they consist of four protein complexes (complexes I–IV),
forming the ETC. Movement of electrons through the ETC during redox reactions is coupled to the pumping of protons
out of the mitochondrial matrix across the IMM and into the mitochondrial intermembrane space, thereby creating potential
energy in the form of a pH gradient and electrical potential across the IMM. Protons accumulated in the intermembrane
space are then channelled back to the matrix across the IMM through the ATP synthase, which uses this energy to make
ATP from ADP and inorganic phosphate (Pi). At the end of the ETC, electrons combine with molecular oxygen and protons
to make water.
The outer mitochondrial membrane (OMM) is a phospholipid bilayer freely permeable to small molecules and harboring
porins that allow bigger molecules to be transported across – among which are the mitochondrial antiviral-signaling protein
(MAVS), and the pore-forming voltage-dependent anion channel (VDAC), which is a major OMM trafficking protein.
Mitochondria are also involved in the storage and release of calcium ions (Ca2+). In mitochondria, Ca2+ are important for
energy production, signaling events, and opening of the mitochondrial permeability transition pore (mPTP). The interplay
between mitochondria and the endoplasmic reticulum (ER), another important calcium storage site, ensures the calcium
homeostasis of the cell. Ca2+ is transported across the OMM to the mitochondrial intermembrane space through VDAC,
then across the IMM to the matrix through specialized transporters.
membrane (IMM) in a process known as oxidative phosphorylation (OXPHOS) (Figure 1). During
this process, 1–2% of the electrons inevitably leak out of the ETC and reduce surrounding molec-
ular O2 to ROS superoxide (O2•–) [4]. Electron leakage occurs primarily at complexes I and III of the
ETC and is enhanced by saturation with electrons fed from the redox cofactor NADH or when
reverse electron transport occurs due to high proton motive force within the mitochondria during
respiration [5]. Notably, inhibition of electron transfer from complexes I and III to subsequent
electron carriers within the ETC by rotenone and antimycin A, respectively, can lead to accumu-
lation and leakage of electrons from these complexes [5,6]. While ROS can be produced through
various mechanisms, the ETC represents a major source of ROS within the cell. At steady state,
redox homeostasis is maintained through a series of reactions involving the dismutation of O2•–
to hydrogen peroxide (H2O2) by superoxide dismutase enzymes [SOD1 and SOD3 (Cu/ZnSOD;
cytosolic and extracellular) and SOD2 (MnSOD; mitochondrial)], which can be further detoxified
by a number of antioxidant enzymes such as catalases, peroxiredoxins, and glutathione/
glutathione peroxidase systems [7] (Figure 1). Oxidative stress occurs when the rate of
intracellular ROS production outweighs the antioxidant capacity of the cell. This situation
may result from either increased ROS generation and/or disabled antioxidant defenses.
However, challenging the popular dogma that ROS indiscriminately damage cellular macro-
molecules, these reactive molecules have been increasingly recognized and described as
secondary messengers that play a critical role in cell signaling by modulating gene transcription
as well as protein structure, stability, and function [8].
the hepatitis B surface (HBsAg) and core (HBcAg) antigens in the ER led to ER stress and was redox homeostasis by modulating
antioxidant response element
associated with ROS production [22]. (ARE)-dependent transcription and the
expression of antioxidant defense
How ROS production benefits viruses enzymes.
Despite the fact that virus-induced increased ROS levels trigger innate antiviral immunity, studies Reactive oxygen species (ROS):
oxygen species formed by the partial
have increasingly reported that ROS production benefits viruses. These claims are mostly based reduction of oxygen, including radicals
on the observation that antioxidant treatments resulted in lower viral titers in vitro in the culture su- such as superoxide anion (•O−2),
pernatant, and improved disease outcome in animal models. While the latter often results from peroxide radicals (•O–22 ), hydroxyl
preventing detrimental organ and tissue damage linked to excessive production of ROS, the for- radicals (HO•), hydroxyl ion (OH–), as
well as non-radicals such as hydrogen
mer suggests a more direct role of ROS during the virus intracellular life cycle. However, fewer peroxide (H2O2), singlet oxygen (1O2),
studies have pinpointed the molecular mechanisms involved, and this may be due to the pleiotro- and ozone (O3). ROS alter biomolecules'
pic effects of ROS within the cell. The mechanisms by which viruses benefit from increased ROS redox status, including proteins, lipids,
and nucleic acids, thereby affecting their
levels during infection can be broadly categorized into two classes, namely direct and indirect
functions. Generation of ROS has been
mechanisms. The former involves oxidative changes of viral protein activities while the latter con- recognized as an important mechanism
sists of ROS-mediated modulation of host signaling pathways. that regulates signaling events inside the
cell (also known as redox signaling),
while excessive levels of ROS can lead to
Oxidative modification of viral proteins
cell and tissue damage.
Oxidative stress has been known to induce covalent modifications of biological molecules, includ- Redox homeostasis: refers to the
ing DNA, protein, lipids, and glycans, resulting in loss or gain of function [23]. Naturally, viral mol- endogenous capacity of cells to maintain
ecules are also expected to undergo redox-mediated changes, which modulate their function a physiological redox status, essential for
the proper functioning of cellular
and activity. Consistently, the guanylyltransferase activity of alphavirus nsP1 and flavivirus NS5 processes and signaling events.
proteins was found to be enhanced under oxidative conditions, resulting in increased RNA cap-
ping activity and increased RNA genome replication [24]. The redox status was also predicted to
significantly impact binding affinity of the severe acute respiratory syndrome coronavirus 2
(SARS-Cov2) Spike protein to its receptor angiotensin-converting enzyme 2 (ACE2) through
influencing disulfide bond formation in both proteins [25]. Furthermore, SUMOylation is a post-
translation modification that is modulated by oxidative stress [26], and viruses have exploited
this activity. During infection with infectious salmon anemia (ISA) virus, nucleocapsid (NP)-induced
ROS results in SUMOylation of this protein that correlated with improved virus output [27]. SUMO
modification of the RNA-dependent RNA polymerase 3D and protease 3C from EV-A71 was also
reported to modulate their activity [28,29].
Other evidence for redox regulation of viral protein activity is illustrated by the observation that
glutathionylation of cysteine residues Cys67 and Cys95 modulates HIV-1 protease activity [30].
Redox regulation of HIV-1 protease activity is believed to allow timely virus maturation and pre-
vents premature host cell death, given the ability of this viral protein to cleave host prosurvival
Bcl-2 protein [31]. Interestingly, the presence of these cysteine residues at the dimerization inter-
face of other retroviral proteases from feline immunodeficiency virus (FIV) and Simian immunode-
ficiency virus (SIV) suggests that this redox-mediated modulation of viral protease activity may be
conserved among retroviruses [32].
Trends in Microbiology
Figure 1. Interactions between viral proteins and components of the electron transport chain (ETC). (A) Viral proteins such as HBsAg and HBcAg from
hepatitis B virus (HBV) induce stress in the endoplasmic reticulum (ER) and calcium efflux; the calcium can be taken up by mitochondria, leading to the production of
reactive oxygen species (ROS). (B) HBx protein interacts with voltage-dependent anion channel 3 (VDAC3) and promotes influx of mitochondrial calcium. (C) Hepatitis
C virus (HCV) E1, E2, and NS3 proteins inhibit complex I of the ETC, resulting in the production of ROS. (D) Enterovirus A71 (EV-A71) replication complexes interact
with mitochondrial prohibitin 1 (PHB1), which may alter mitochondria integrity and function. (E) Human immunodeficiency virus Vpr protein increases mitochondrial
membrane permeabilization by interacting with adenine nucleotide translocator (ANT) of the mitochondrial permeability transition pore (mPTP), leading to efflux of
apoptogenic factors. (F) HIV Tat protein dysregulates anionic channels, leading to loss of mitochondrial transmembrane potential (MMP).
signaling, and induction of the ER stress/autophagy signaling pathway [39], all of which could be
abrogated by ROS scavengers, with a resultant decrease in viral replication. Modulation of
these pathways during infection has been associated with timely induction of apoptosis for
optimal viral replication within the host cell.
HCV is another example of a virus that benefits from increased ROS levels through modulation
of signaling pathways. HCV NS5A protein has been reported to induce expression of Death
Receptor 6 (DR6) through ROS-mediated activation of NF-κB signaling, which could be
abrogated by antioxidant N-acetylcysteine (NAC) and NF-κB inhibitor, SN50 [40]. Interaction
between NS5A and DR6 was then found to be critical for production of infectious virus, through
Box 2. ER stress
The ER is the site where membrane and secreted cellular proteins are synthesized and folded. This implies a close
monitoring of the protein flux through the ER, to avoid build-up of misfolded proteins, thereby causing ER stress that
may impair cellular functions and homeostasis [99]. An elaborate array of signaling pathways from the ER to the cytosol
and nucleus allow the cell to detect and respond to the presence of misfolded proteins within the ER. Such response is
known as the unfolded protein response (UPR). The UPR starts with ER-resident proteins that are able to sense ER stress
through their domains protruding within the ER lumen and activate a cascade of signaling events via their cytosolic effector
domains. The initial goal of the UPR is to restore cellular function and it does so by downregulating protein translation,
degrading misfolded proteins, and increasing the production of chaperones involved in protein folding. Should this outcome
not be achieved by a certain timeline, the UPR triggers apoptosis. In addition, ER stress is associated with Ca2+ release from
the lumen into the cytosol, and Ca2+ eventually accumulates in the mitochondrial matrix (see Box 1 in the main text), which,
beyond a critical threshold, triggers proapoptotic mitochondrial outer membrane permeabilization (MOMP).
Chronic ER stress and sustained over-activation of the UPR have been increasingly recognized as key players in diseases
such as cancer, metabolic (diabetes) and neurodegenerative diseases, and inhibiting the UPR represents a potential
therapeutic strategy for these diseases. ER stress also occurs during viral infection, where large amounts of viral proteins
are being produced and saturate the ER.
dampening the ability of DR6 to suppress AKT-mediated survival signaling, thereby allowing HCV
to replicate without triggering cell death [40]. Besides NF-κB, HCV-induced ROS production also
activates signal transducer and activator of transcription (STAT)-3 signaling [41], leading to increased
expression of the long noncoding RNAs, lnc-7SK and lnc-IGF2-AF, which promote HCV replication,
and this was associated with upregulation of phosphatidylinositol 4-phosphate (PI4P) [42]. While the
proviral role of PI4P during HCV infection remains poorly defined at the molecular level, other studies
have reported its involvement in HCV replication and secretion [43,44].
Cap-independent translation initiation was first described in picornaviruses, whose viral RNA genome 5′ end is uncapped,
and for which translation is initiated at highly structured untranslated regions (UTRs) called internal ribosome entry sites
(IRESs) where ribosomes, translation initiation factors (IFs), and other factors known as IRES-transacting factors (ITAFs)
are directly recruited. Since then, cap-independent initiation of translation has been described for many RNA viruses
[109]. Viral IRESs can be categorized into four groups depending on their requirements in IFs and ITAFs, with the idea that
the more compact IRES needs less of these auxiliary proteins. Cap-independent initiation of translation also occurs in
eukaryotes, mostly for genes involved in response to cellular stresses such as viral infection, during which cap-dependent
translation is shut down by various viral proteolytic activities. Cellular IRESs are less structured than their viral counterparts
with little sequence and structure conservation among them.
Furthermore, other non-IRES cap-independent mechanisms of translation initiation have also been reported, and these
mechanisms involve cap-independent translation elements (CITEs) that can be found in the 3′ UTR of many plant RNA
viruses with various structures and sequences.
dependent manner was found to be responsible for enhanced HIV replication in CD4+ helper T
cells and secretion of extracellular vesicles [46]. Further studies have shown that ROS-
dependent stabilization of HIF-1α during HIV infection resulted in enhanced expression of
CXCR4 and glucose transporter (Glut), thereby resulting in increased HIV entry and glucose
uptake, respectively [47–50]. These studies thus indicated that the ROS/HIF-1α/Glut/CXCR4
axis is exploited by HIV to support its replication and facilitate infection of neighboring cells.
A number of DNA viruses from the families of poxviruses, baculoviruses, mimiviruses, and
phycodnaviruses have also been documented to encode a copper–zinc superoxide dismutase
(Cu–Zn SOD)-like protein, although the enzymatic activity and/or the role of these proteins during
the virus life cycle remain to be experimentally demonstrated [54]. Functionality of the Cu–Zn SOD
produced by Paramecium bursaria chlorella virus 1 (PBCV-1) of the family Phycodnaviridae, genus
Chlorovirus, has however been reported [54]. The authors proposed that the enzyme is packaged
within the virion and may contribute to reducing ROS levels during the early phase of virus infection.
Furthermore, some viruses have evolved strategies to counteract the potential negative outcome
of oxidative-stress-induced lipid peroxidation on their own components. Lipid peroxidation
results in formation of lipid peroxides and reactive aldehydes that are cytotoxic, leading to
membrane dysfunction and altering DNA and protein structure and functions [55]. NS5A from
HCV was found to induce the production of glutathione peroxidase 4 (GPx4), which reduces
lipid peroxides, and therefore helps to limit incorporation of lipid peroxides into newly formed
virions that may otherwise alter their fusiogenic activity [56]. Furthermore, the discovery of residues
in HCV NS4A and NS5B that are resistant to lipid peroxidation suggests another mechanism by
which viruses are able to cope with oxidative-stress-induced lipid peroxidation [57].
to induce ROS production during the early phase of infection [58], but it increases GSH synthesis
afterwards to antagonize oxidative stress, thereby maintaining mTORC1 activity, which is critical
for cell growth and proliferation [59]. Similarly, in an HCV infection model using the human hepa-
tocyte Huh 7.5 cell line, increased oxidative stress was associated with high HCV replication
activity and cellular apoptotic activity, followed by increased expression of metabolic enzymes
involved in de novo synthesis and regeneration of GSH, resulting in lower HCV replication activity,
thereby favoring viral persistence [60]. This finding thus suggested that HCV controls its
replication activity, and hence the infection status (acute or chronic), through modulation of the
redox environment. The increased levels of GSH may result from increased expression of Nrf2,
a transcription factor that drives GSH de novo synthesis. The same study also showed that
dormant HCV infection in Huh 7.5 cells could be reactivated by addition of the pro-oxidant
drug, auranofin, leading to higher viral replication activity and apoptosis [60].
A biphasic activity was also observed during influenza A virus (IAV) infection where MAPK
activation peaked during virus entry and at late stages of IAV infection when nuclear export of
IAV nucleoprotein (NP) is facilitated via ERK phosphorylation of its Ser392 residue [61,62]. Given
the known role of ROS in regulating MAPKs activity [63], it is plausible that IAV actively manipulates
the host redox status throughout its replication cycle to ensure a successful outcome.
The release of proapoptotic factors as a result of MOMP is regulated by Bcl-2 family proteins [72],
which can be categorized into antiapoptotic members [including Bcl-2, B cell lymphoma-extra-
Trends in Microbiology
Figure 2. Viral modulation of the intrinsic apoptotic pathway. Viruses can inhibit cell apoptosis by directly regulating caspase activity (A), or by antagonizing (B) or
inhibiting (C) proapoptotic Bcl-2 proteins. Alternatively, a virus can induce cell apoptosis by promoting activation of BAK and BAX (D), or by promoting mitochondria
permeabilization and release of cytochrome c (E). A virus may also increase apoptotic sensitivity by relieving Apaf-1 inhibition (F), or by increasing host translation of
proapoptotic factors such as Apaf-1 (G). Abbreviations: ADV, adenovirus; CMV, cytomegalovirus; EV-A71, enterovirus A71; hnRNPA1, heterogeneous nuclear
ribonucleoprotein A1; IAP, inhibitor of apoptosis proteins; IAV, influenza A virus; IRES, internal ribosome entry site; MOMP, mitochondrial outer membrane
permeabilization; SMAC, second mitochondria-activator of caspases; VACV, vaccinia virus.
large (Bcl-xL), induced myeloid leukemia cell differentiation protein (Mcl-1)], and proapoptotic
members [such as Bcl-2-associated X protein (Bax), Bcl-2 homologous antagonist killer (Bak),
Bid, and Bad] (Figure 2).
Since virus-induced ROS promotes cell death, viruses have evolved strategies to interact with
mitochondria in order to take control of the intrinsic apoptotic pathway and counteract undesirable
or untimely cell death. Indeed, while preventing cell host apoptosis allows viruses to complete their
intracellular life cycle, inducing apoptosis represents a means for some of them to exit the host cell
and release their viral progeny.
E1B 19K from adenovirus represents one of the earliest viral Bcl-2 antagonists discovered [73]. In
spite of limited sequence similarity to mammalian Bcl-2, E1B 19K functionally complements
human Bcl-2 and inhibits p53-dependent apoptosis through interaction with proapoptotic
protein Bak, preventing Bax–Bak assembly on the mitochondrial membrane [74]. E1B 19K also
binds to Bcl-2 interacting killer (Bik/Nbk), which disrupts Bak/Mcl-1and Bak/Bcl-xL interactions,
further dampening mitochondria-dependent intrinsic apoptosis. The antagonistic function of E1B
19K towards proapoptotic proteins prolongs the viability of host cells, thereby allowing sufficient
time for virus maturation to complete.
Vaccinia virus of the family Poxviridae also encodes a Bcl-2-antagonizing protein, named F1L
[75]. F1L also lacks discernible sequence homology with mammalian Bcl-2, localizes to the
mitochondrial membrane, and inhibits apoptosis through interaction with proapoptotic Bim and
Bak proteins [76]. F1L also inhibits apoptosis by preventing mitochondrial translocation and
oligomerization of Bax on the mitochondrial membrane in Bak-deficient cells; however, this is
likely to occur via an indirect mechanism, as interaction between F1L and Bax has not been
reported [76], possibly involving the inhibition of Bim upstream [77].
Other viral strategies to prevent premature apoptosis of the host cell have been reported
(Figure 2). Epstein–Barr nuclear antigen (EBNA) 3A and 3C proteins were found to downregulate
the expression of proapoptotic protein Bim, thus promoting survival of EBV-infected B cells [78].
The UL36 gene product from CMV suppresses caspase 8 proteolytic activation by binding to its
pro-domain [79]. Of note, while caspase 8 is an initiator of extrinsic apoptosis, its inhibition also
prevents the activation of proapoptotic Bid, thereby inhibiting further amplification of apoptotic
signal through the mitochondrial pathway [80]. CMV can also directly act on mediators of intrinsic
apoptosis Bax and Bak through viral mitochondrial localized inhibitor of apoptosis (vMIA), a gene
product of UL37 [81–83]. In addition, CMV produces antiapoptotic noncoding miRNAβ2.7,
which binds and stabilizes complex I of the ETC, preventing release of the GRIM19 subunit,
thereby maintaining MMP and respiration in CMV-infected cells and improving cell viability
when challenged with metabolic stressors [84].
Virus-induced apoptosis
A number of viruses rely on cell lysis to release viral progeny in the external environment and infect
neighboring cells. Consistently, viral induction of the intrinsic apoptotic pathway has been
described. Two such prominent examples include EV-A71 [85] and IAV [86] for which a variety
of mechanisms have been reported (Figure 2).
Nonstructural protein 2B (NS2B) of EV-A71 signals Bax translocation to the mitochondrial outer
membrane and induces its conformational activation, resulting in formation of the apoptotic pore
[85]. One might dare to conjecture that an increase in intracellular ROS may represent a probable
signal for the mitochondrial recruitment and oligomerization of Bax, which further fuels mtROS
generation and apoptosis. EV-A71 3C protease has also been implicated in apoptosis induction
through cleavage of host protein heterogenous ribonucleoprotein A1 (hnRNPA1), which relieves
its binding to the IRES sequence in Apaf-1 mRNA and allows IRES-dependent synthesis of
Apaf-1, thereby resulting in downstream caspase activation and apoptosis [87]. Increased
IRES-dependent translation of Apaf-1 is further contributed to by EV-A71 NS2A-mediated cleav-
age of eukaryotic initiation factor 4G1 (eIF4G1), a key factor for cap-dependent mRNA translation
[88]. Downstream cytochrome c release, both M1 and NP proteins from IAV have also been im-
plicated in the upregulation of Apaf-1. M1 binds to Hsp70 and reduces its interaction with Apaf-1,
allowing apoptosome assembly [89]. Conversely, NP downregulates the expression of host ap-
optosis inhibitor 5 (API5), a protein that antagonizes E2F transcription factor 1 (E2F1), which
leads to upregulation of its gene target Apaf-1 [90]. Furthermore, PB1-F2 protein from IAV local-
izes to the mitochondria and increases permeability of the mitochondrial membrane by interacting
with ANT3 and VDAC1 of the mPTP, resulting in MMP loss and increased efflux of cytochrome c
[91]. In addition, IAV-induced apoptosis likely involves activation of proapoptotic Bax/Bak pro-
teins and downregulation of antiapoptotic proteins Bcl-xL and Mcl-1 [92]. Activation of MAPKs
p38 and ERK1/2 during IAV infection could also promote host cell apoptosis through their
However, while clinical benefits of antioxidant treatments in the context of viral infections have been reported, this approach is
unlikely to be sufficient to effectively protect against explosive, acute viral infections. Combining antioxidants with antiviral
drugs instead may represent a promising therapeutic approach [108]. In addition, a deeper understanding of the contribution
of oxidative stress during viral infections may still be required to harness fully the therapeutic potential of antioxidants.
associated signaling targets, namely Bcl-2 and p53 [93,94]. Activation of these MAPKs was ob-
served during the late phase of the IAV replication cycle, which is believed to assist in the release
of viral particles via apoptosis [61,94].
Concluding remarks
Viruses have devised sophisticated and well-orchestrated strategies to interact with multiple
mitochondrial components in order to support their intracellular life cycle. This includes timely
manipulation of the redox metabolism, and mitochondrial integrity and function. A dichotomy of
functional outcomes vis-à-vis redox modifications from the standpoint of cell fate determination
illustrates the remarkable complexity of intracellular perturbations triggered during the acute
and chronic phase of a viral infection. While much knowledge has been generated over the past
decades, and has shed light on the molecular interactions between viruses and mitochondria,
the pleiotropic effects of these interactions has made it very challenging to apprehend fully the
implications of these interactions. More specifically, our incomplete understanding of the molecular
mechanisms and dynamics of redox modulation during viral infection has limited our ability to
harness fully the therapeutic potential of antioxidants (Box 4).
involving mitochondria and lipid peroxidation. Whether and how viruses interfere with, and
manipulate, this pathway remains to be investigated [98].
Declaration of interests
No interests are declared.
References
1. Spinelli, J.B. and Haigis, M.C. (2018) The multifaceted 23. Grimsrud, P.A. et al. (2008) Oxidative stress and covalent mod-
contributions of mitochondria to cellular metabolism. Nat. Cell ification of protein with bioactive aldehydes. J. Biol. Chem. 283,
Biol. 20, 745–754 21837–21841
2. Nunnari, J. and Suomalainen, A. (2012) Mitochondria: in 24. Gullberg, R.C. et al. (2015) Oxidative stress influences positive
sickness and in health. Cell 148, 1145–1159 strand RNA virus genome synthesis and capping. Virology
3. Ren, Z. et al. (2020) Regulation of MAVS expression and 475, 219–229
signaling function in the antiviral innate immune response. 25. Hati, S. and Bhattacharyya, S. (2020) Impact of thiol-disulfide
Front. Immunol. 11, 1030 balance on the binding of Covid-19 Spike protein with
4. Hernansanz-Agustín, P. and Enríquez, J.A. (2021) Generation of angiotensin-converting enzyme 2 receptor. ACS Omega 5,
reactive oxygen species by mitochondria. Antioxidants 10, 415 16292–16298
5. Mailloux, R.J. (2015) Teaching the fundamentals of electron 26. Stankovic-Valentin, N. et al. (2016) Redox regulation of SUMO
transfer reactions in mitochondria and the production and enzymes is required for ATM activity and survival in oxidative
detection of reactive oxygen species. Redox Biol. 4, 381–398 stress. EMBO J. 35, 1312–1329
6. Li, N. et al. (2003) Mitochondrial complex I inhibitor rotenone 27. Fredericksen, F. et al. (2019) Sumoylation of nucleoprotein (NP)
induces apoptosis through enhancing mitochondrial reactive mediated by activation of NADPH oxidase complex is a conse-
oxygen species production. J. Biol. Chem. 278, 8516–8525 quence of oxidative cellular stress during infection by Infectious
7. He, L. et al. (2017) Antioxidants maintain cellular redox homeo- salmon anemia (ISA) virus necessary to viral progeny. Virology
stasis by elimination of reactive oxygen species. Cell. Physiol. 531, 269–279
Biochem. 44, 532–553 28. Liu, Y. et al. (2016) SUMO modification stabilizes enterovirus
8. Knaus, U.G. (2021) Oxidants in physiological processes. 71 polymerase 3D to facilitate viral replication. J. Virol. 90,
Handb. Exp. Pharmacol. 264, 27–47 10472–10485
9. Yang, B. and Bouchard, M.J. (2012) The hepatitis B virus X 29. Chen, S.C. et al. (2011) Sumoylation-promoted enterovirus 71
protein elevates cytosolic calcium signals by modulating mito- 3C degradation correlates with a reduction in viral replication
chondrial calcium uptake. J. Virol. 86, 313–327 and cell apoptosis. J. Biol. Chem. 286, 31373–31384
10. Rahmani, Z. et al. (2000) Hepatitis B virus X protein colocalizes 30. Davis, D.A. et al. (1997) Thioltransferase (glutaredoxin) is
to mitochondria with a human voltage-dependent anion channel, detected within HIV-1 and can regulate the activity of
HVDAC3, and alters its transmembrane potential. J. Virol. 74, glutathionylated HIV-1 protease in vitro. J. Biol.Chem. 272,
2840–2846 25935–25940
11. McClain, S.L. et al. (2007) Hepatitis B virus replication is asso- 31. Strack, P.R. et al. (1996) Apoptosis mediated by HIV protease
ciated with an HBx-dependent mitochondrion-regulated is preceded by cleavage of Bcl-2. Proc. Natl. Acad. Sci. U. S. A.
increase in cytosolic calcium levels. J. Virol. 81, 12061–12065 93, 9571–9576
12. Piccoli, C. et al. (2007) Hepatitis C virus protein expression 32. Davis, D.A. et al. (2003) Reversible oxidative modification as a
causes calcium-mediated mitochondrial bioenergetic dysfunc- mechanism for regulating retroviral protease dimerization and
tion and nitro-oxidative stress. Hepatology (Baltimore) 46, 58–65 activation. J. Virol. 77, 3319–3325
13. Macho, A. et al. (1999) Susceptibility of HIV-1-TAT transfected 33. Cheng, M.L. et al. (2014) Enterovirus 71 induces mitochondrial
cells to undergo apoptosis. Biochemical mechanisms. Oncogene reactive oxygen species generation that is required for efficient
18, 7543–7551 replication. PLoS One 9, e113234
14. Lecoeur, H. et al. (2012) HIV-1 Tat protein directly induces 34. Tung, W.H. et al. (2011) Enterovirus 71 induces integrin β1/
mitochondrial membrane permeabilization and inactivates EGFR-Rac1-dependent oxidative stress in SK-N-SH cells:
cytochrome c oxidase. Cell Death Dis. 3, e282 role of HO-1/CO in viral replication. J. Cell. Physiol. 226,
15. Jacotot, E. et al. (2000) The HIV-1 viral protein R induces 3316–3329
apoptosis via a direct effect on the mitochondrial permeability 35. Li, H. et al. (2020) EV71 infection induces cell apoptosis
transition pore. J. Exp. Med. 191, 33–46 through ROS generation and SIRT1 activation. J. Cell. Biochem.
16. Too, I.H.K. et al. (2018) Prohibitin plays a critical role in Enterovirus 121, 4321–4331
71 neuropathogenesis. PLoS Pathog. 14, e1006778 36. Bai, Z. et al. (2020) EV71 virus reduces Nrf2 activation to
17. Chang, L.Y. et al. (2019) Enterovirus A71 neurologic complica- promote production of reactive oxygen species in infected
tions and long-term sequelae. J. Biomed. Sci. 26, 57 cells. Gut Pathog. 12, 22
18. Nijtmans, L.G. et al. (2000) Prohibitins act as a membrane- 37. Chen, D. et al. (2018) Harmine, a small molecule derived from
bound chaperone for the stabilization of mitochondrial natural sources, inhibits enterovirus 71 replication by targeting
proteins. EMBO J. 19, 2444–2451 NF-κB pathway. Int. Immunopharmacol. 60, 111–120
19. Huang, C.Y. et al. (2012) HIV-1 Vpr triggers mitochondrial 38. Lv, X. et al. (2014) Apigenin inhibits enterovirus 71 replication
destruction by impairing Mfn2-mediated ER-mitochondria through suppressing viral IRES activity and modulating cellular
interaction. PLoS One 7, e33657 JNK pathway. Antivir. Res. 109, 30–41
20. Kang, S.M. et al. (2009) Interaction of hepatitis C virus core 39. Du, N. et al. (2019) Resveratrol-loaded nanoparticles inhibit
protein with Hsp60 triggers the production of reactive oxygen enterovirus 71 replication through the oxidative stress-mediated
species and enhances TNF-alpha-mediated apoptosis. Cancer ERS/autophagy pathway. Int. J. Mol. Med. 44, 737–749
Lett. 279, 230–237 40. Luong, T.T.D. et al. (2017) Hepatitis C virus exploits death
21. Rizzuto, R. et al. (2012) Mitochondria as sensors and regula- receptor 6-mediated signaling pathway to facilitate viral
tors of calcium signalling. Nat. Rev. Mol. Cell Biol. 13, 566–578 propagation. Sci. Rep. 7, 6445
22. Choi, Y.M. et al. (2019) Naturally occurring hepatitis B virus 41. Machida, K. et al. (2006) Hepatitis C virus triggers mitochon-
mutations leading to endoplasmic reticulum stress and their drial permeability transition with production of reactive oxygen
contribution to the progression of hepatocellular carcinoma. species, leading to DNA damage and STAT3 activation.
Int. J. Mol. Sci. 20, 597 J. Virol. 80, 7199N7207
42. Xiong, Y. et al. (2015) STAT3-regulated long non-coding RNAs 67. Wu, Y.H. et al. (2021) Aerobic glycolysis supports hepatitis B
lnc-7SK and lnc-IGF2-AS promote hepatitis C virus replication. virus protein synthesis through interaction between viral sur-
Mol. Med. Rep. 12, 6738–6744 face antigen and pyruvate kinase isoform M2. PLoS Pathog.
43. Berger, K.L. et al. (2011) Hepatitis C virus stimulates the phos- 17, e1008866
phatidylinositol 4-kinase III alpha-dependent phosphatidylinositol 68. Jung, G.S. et al. (2016) Pyruvate dehydrogenase kinase regu-
4-phosphate production that is essential for its replication. lates hepatitis C virus replication. Sci. Rep. 6, 30846
J. Virol. 85, 8870–8883 69. Le, A. et al. (2010) Inhibition of lactate dehydrogenase A
44. Bishé, B. et al. (2012) Role of phosphatidylinositol 4-phosphate induces oxidative stress and inhibits tumor progression. Proc.
(PI4P) and its binding protein GOLPH3 in hepatitis C virus Natl. Acad. Sci. U. S. A. 107, 2037–2042
secretion. J. Biol. Chem. 287, 27637–27647 70. Wu, S.B. and Wei, Y.H. (2012) AMPK-mediated increase of
45. Codo, A.C. et al. (2020) Elevated glucose levels favor SARS- glycolysis as an adaptive response to oxidative stress in
CoV-2 infection and monocyte response through a HIF-1α/ human cells: implication of the cell survival in mitochondrial
glycolysis-dependent axis. Cell Metab. 32, 437–446.e5 diseases. Biochim. Biophys. Acta 1822, 233–247
46. Duette, G. et al. (2018) Induction of HIF-1α by HIV-1 infection in 71. Carneiro, B.A. and El-Deiry, W.S. (2020) Targeting apoptosis in
CD4(+) T cells promotes viral replication and drives extracellular cancer therapy. Nat. Rev. Clin. Oncol. 17, 395–417
vesicle-mediated inflammation. mBio 9, e00757-18 72. Kale, J. et al. (2018) BCL-2 family proteins: changing partners
47. Chetram, M.A. and Hinton, C.V. (2013) ROS-mediated in the dance towards death. Cell Death Differ. 25, 65–80
regulation of CXCR4 in cancer. Front. Biol. 8. https://doi.org/ 73. Subramanian, T. et al. (1995) Functional similarity between
10.1007/s11515-012-1204-4 adenovirus E1B 19-kDa protein and proteins encoded by
48. Lan, X. et al. (2013) High glucose enhances HIV entry into T cells Bcl-2 proto-oncogene and Epstein–Barr virus BHRF1 gene.
through upregulation of CXCR4. J. Leukoc. Biol. 94, 769–777 Curr. Top. Microbiol. Immunol. 199, 153–161
49. Loisel-Meyer, S. et al. (2012) Glut1-mediated glucose transport 74. Radke, J.R. et al. (2014) Adenovirus E1B 19-kilodalton protein
regulates HIV infection. Proc. Natl. Acad. Sci. U. S. A. 109, modulates innate immunity through apoptotic mimicry. J. Virol.
2549–2554 88, 2658–2669
50. Kavanagh Williamson, M. et al. (2018) Upregulation of glucose 75. Wasilenko, S.T. et al. (2003) Vaccinia virus encodes a previously
uptake and hexokinase activity of primary human CD4+ T cells uncharacterized mitochondrial-associated inhibitor of apoptosis.
in response to infection with HIV-1. Viruses 10, 114 Proc. Natl. Acad. Sci. U. S. A. 100, 14345–14350
51. Shim, J.H. et al. (2005) E7-expressing HaCaT keratinocyte 76. Taylor, J.M. et al. (2006) The vaccinia virus protein F1L
cells are resistant to oxidative stress-induced cell death via interacts with Bim and inhibits activation of the pro-apoptotic
the induction of catalase. Proteomics 5, 2112–2122 protein Bax. J. Biol. Chem. 281, 39728–39739
52. Qadri, I. et al. (2004) Induced oxidative stress and activated 77. Chi, X. et al. (2020) The carboxyl-terminal sequence of Bim
expression of manganese superoxide dismutase during hepa- enables Bax activation and killing of unprimed cells. eLife 9,
titis C virus replication: role of JNK, p38 MAPK and AP-1. e44525
Biochem. J. 378, 919–928 78. Anderton, E. et al. (2008) Two Epstein–Barr virus (EBV)
53. Sheyn, U. et al. (2016) Modulation of host ROS metabolism is oncoproteins cooperate to repress expression of the proapoptotic
essential for viral infection of a bloom-forming coccolithophore tumour-suppressor Bim: clues to the pathogenesis of Burkitt's
in the ocean. ISME J. 10, 1742–1754 lymphoma. Oncogene 27, 421–433
54. Kang, M. et al. (2014) Chlorovirus PBCV-1 encodes an active 79. Brune, W. and Andoniou, C.E. (2017) Die another day: inhibi-
copper-zinc superoxide dismutase. J. Virol. 88, 12541–12550 tion of cell death pathways by Cytomegalovirus. Viruses 9, 249
55. Gaschler, M.M. and Stockwell, B.R. (2017) Lipid peroxidation 80. Wilson, W.H. et al. (2010) Navitoclax, a targeted high-affinity
in cell death. Biochem. Biophys. Res. Commun. 482, 419–425 inhibitor of BCL-2, in lymphoid malignancies: a phase 1 dose-
56. Brault, C. et al. (2016) Glutathione peroxidase 4 is reversibly escalation study of safety, pharmacokinetics, pharmacodynamics,
induced by HCV to control lipid peroxidation and to increase and antitumour activity. Lancet Oncol. 11, 1149–1159
virion infectivity. Gut 65, 144–154 81. Goldmacher, V.S. et al. (1999) A cytomegalovirus-encoded
57. Yamane, D. et al. (2014) Regulation of the hepatitis C virus mitochondria-localized inhibitor of apoptosis structurally unre-
RNA replicase by endogenous lipid peroxidation. Nat. Med. lated to Bcl-2. Proc. Natl. Acad. Sci. U. S. A. 96, 12536–12541
20, 927–935 82. Poncet, D. et al. (2004) An anti-apoptotic viral protein that
58. Speir, E. et al. (1996) Role of reactive oxygen intermediates in recruits Bax to mitochondria. J. Biol. Chem. 279, 22605–22614
cytomegalovirus gene expression and in the response of human 83. Norris, K.L. and Youle, R.J. (2008) Cytomegalovirus proteins
smooth muscle cells to viral infection. Circ. Res. 79, 1143–1152 vMIA and m38.5 link mitochondrial morphogenesis to Bcl-2
59. Tilton, C. et al. (2011) Human cytomegalovirus induces multi- family proteins. J. Virol. 82, 6232–6243
ple means to combat reactive oxygen species. J. Virol. 85, 84. Reeves, M.B. et al. (2007) Complex I binding by a virally
12585–12593 encoded RNA regulates mitochondria-induced cell death.
60. Anticoli, S. et al. (2019) Counteraction of HCV-induced Science (New York) 316, 1345–1348
oxidative stress concurs to establish chronic infection in liver 85. Cong, H. et al. (2016) Enterovirus 71 2B induces cell apoptosis
cell cultures. Oxidative Med. Cell. Longev. 2019, 6452390 by directly inducing the conformational activation of the
61. Schreiber, A. et al. (2020) Dissecting the mechanism of proapoptotic protein Bax. J. Virol. 90, 9862–9877
signaling-triggered nuclear export of newly synthesized influ- 86. Ampomah, P.B. and Lim, L.H.K. (2020) Influenza A virus-
enza virus ribonucleoprotein complexes. Proc. Natl. Acad. induced apoptosis and virus propagation. Apoptosis 25, 1–11
Sci. U. S. A. 117, 16557–16566 87. Li, M.L. et al. (2019) EV71 3C protease induces apoptosis by
62. Droebner, K. et al. (2011) Antiviral activity of the MEK-inhibitor cleavage of hnRNP A1 to promote apaf-1 translation. PLoS
U0126 against pandemic H1N1v and highly pathogenic avian One 14, e0221048
influenza virus in vitro and in vivo. Antivir. Res. 92, 195–203 88. Kuo, R.L. et al. (2002) Infection with enterovirus 71 or expres-
63. Son, Y. et al. (2011) Mitogen-activated protein kinases and sion of its 2A protease induces apoptotic cell death. J. Gen.
reactive oxygen species: how can ROS activate MAPK pathways? Virol. 83, 1367–1376
J. Signal Transduct. 2011, 792639 89. Halder, U.C. et al. (2011) Cell death regulation during influenza
64. Thaker, S.K. et al. (2019) Viral hijacking of cellular metabolism. A virus infection by matrix (M1) protein: a model of viral control
BMC Biol. 17, 59 over the cellular survival pathway. Cell Death Dis. 2, e197
65. Kishimoto, N. et al. (2021) Glucose-dependent aerobic glycolysis 90. Mayank, A.K. et al. (2015) Nucleoprotein of influenza A virus
contributes to recruiting viral components into HIV-1 particles to negatively impacts antiapoptotic protein API5 to enhance
maintain infectivity. Biochem. Biophys. Res. Commun. 549, E2F1-dependent apoptosis and virus replication. Cell Death
187–193 Dis. 6, e2018
66. Icard, P. et al. (2021) The key role of Warburg effect in SARS- 91. Zamarin, D. et al. (2005) Influenza virus PB1-F2 protein induces
CoV-2 replication and associated inflammatory response. cell death through mitochondrial ANT3 and VDAC1. PLoS
Biochimie 180, 169–177 Pathog. 1, e4
92. McLean, J.E. et al. (2009) Lack of Bax prevents influenza A tolerance against pathogens by melatonin may impact
virus-induced apoptosis and causes diminished viral replication. outcome of deadly virus infection pertinent to COVID-19.
J. Virol. 83, 8233–8246 Molecules (Basel) 25, 4410
93. De Chiara, G. et al. (2006) Bcl-2 Phosphorylation by p38 MAPK: 102. de Las Heras, N. et al. (2020) Implications of oxidative stress
identification of target sites and biologic consequences. J. Biol. and potential role of mitochondrial dysfunction in COVID-19:
Chem. 281, 21353–21361 therapeutic effects of vitamin D. Antioxidants 9, 897
94. Wang, Y. et al. (2018) Pseudo-Ginsenoside Rh2 induces A549 103. Salehi, B. et al. (2018) Resveratrol: a double-edged sword in
cells apoptosis via the Ras/Raf/ERK/p53 pathway. Exp. health benefits. Biomedicines 6, 91
Therapeut. Med. 15, 4916–4924 104. Xiao, Z. et al. (2021) Network pharmacology reveals that
95. Girardi, E. et al. (2018) On the importance of host MicroRNAs resveratrol can alleviate COVID-19-related hyperinflammation.
during viral infection. Front. Genet. 9, 439 Dis. Mark. 2021, 4129993
96. Akbari, A. et al. (2020) Cross-talk between oxidative stress 105. Shi, Z. and Puyo, C.A. (2020) N-Acetylcysteine to combat
signaling and microRNA regulatory systems in carcinogenesis: COVID-19: an evidence review. Ther. Clin. Risk Manage. 16,
Focused on gastrointestinal cancers. Biomed. Pharmacother. 1047–1055
131, 110729 106. Safe, I.P. et al. (2021) Adjunct N-acetylcysteine treatment in
97. Vezza, T. et al. (2021) MicroRNAs and oxidative stress: an hospitalized patients with HIV-associated tuberculosis
intriguing crosstalk to be exploited in the management of dampens the oxidative stress in peripheral blood: results
type 2 diabetes. Antioxidants (Basel) 10, 802 from the RIPENACTB study trial. Front. Immunol. 11,
98. Wang, M.P. et al. (2021) Ferroptosis in viral infection: the unexplored 602589
possibility. Acta Pharmacol. Sin. Published online December 6, 107. Han, S.N. et al. (2000) Effect of long-term dietary antioxidant
2021. https://doi.org/10.1038/s41401-021-00814-1 supplementation on influenza virus infection. J. Gerontol. A
99. Lin, J.H. et al. (2008) Endoplasmic reticulum stress in disease Biol. Sci. Med. Sci. 55, B496–B503
pathogenesis. Annu. Rev. Pathol. 3, 399–425 108. Uchide, N. and Toyoda, H. (2011) Antioxidant therapy as a
100. Fraternale, A. et al. (2021) Intracellular redox-modulated path- potential approach to severe influenza-associated complications.
ways as targets for effective approaches in the treatment of Molecules 16, 2032–2052
viral infection. Int. J. Mol. Sci. 22, 3603 109. Yang, Y. and Wang, Z. (2019) IRES-mediated cap-independent
101. Tan, D.X. and Hardeland, R. (2020) Targeting host defense translation, a path leading to hidden proteome. J. Mol. Cell Biol.
system and rescuing compromised mitochondria to increase 11, 911–919