Skip to main content
Erschienen in: Inflammation 5/2022

Open Access 30.05.2022 | COVID-19 | Review

Disengaging the COVID-19 Clutch as a Discerning Eye Over the Inflammatory Circuit During SARS-CoV-2 Infection

verfasst von: Mohammed Moustapha Anwar, Ranjit Sah, Sunil Shrestha, Akihiko Ozaki, Namrata Roy, Zareena Fathah, Alfonso J. Rodriguez-Morales

Erschienen in: Inflammation | Ausgabe 5/2022

Abstract

Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) causes the cytokine release syndrome (CRS) and leads to multiorgan dysfunction. Mitochondrial dynamics are fundamental to protect against environmental insults, but they are highly susceptible to viral infections. Defective mitochondria are potential sources of reactive oxygen species (ROS). Infection with SARS-CoV-2 damages mitochondria, alters autophagy, reduces nitric oxide (NO), and increases both nicotinamide adenine dinucleotide phosphate oxidases (NOX) and ROS. Patients with coronavirus disease 2019 (COVID-19) exhibited activated toll-like receptors (TLRs) and the Nucleotide-binding and oligomerization domain (NOD-), leucine-rich repeat (LRR-), pyrin domain-containing protein 3 (NLRP3) inflammasome. The activation of TLRs and NLRP3 by SARS‐CoV‐2 induces interleukin 6 (IL-6), IL-1β, IL-18, and lactate dehydrogenase (LDH). Herein, we outline the inflammatory circuit of COVID-19 and what occurs behind the scene, the interplay of NOX/ROS and their role in hypoxia and thrombosis, and the important role of ROS scavengers to reduce COVID-19-related inflammation.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
ACE2
Angiotensin-converting enzyme 2
ALI
Acute lung injury
ARDS
Acute respiratory distress syndrome
ATII
Angiotensin II
BBB
Blood-brain barrier
CGD
Chronic granulomatous disease
COVID-19
Coronavirus disease-19
DIC
Disseminated intravascular coagulation
DPI
Diphenyleneiodonium chloride
HIV
Human immunodeficiency virus
HUVECs
Human umbilical vein endothelial cells
IFN-I
Type I interferon
IFN-α
Interferon-alpha
IFN-γ
Interferon-gamma
IL-6
Interleukin-6
KO
Knockout
LTB4
Leukotriene B4
MAPK
Mitogen-activated protein kinase
MAVS
Mitochondrial antiviral signalling molecule
MCAO
Middle cerebral artery occlusion
MI
Myocardial infarction
MyD88
Myeloid differentiation primary response 88
NF-κB
Nuclear factor kappa-light-chain-enhancer of activated B cells
NO
Nitric oxide
NOX
Nicotinamide adenine dinucleotide phosphate oxidases
OGD
Oxygen/glucose deprivation
ORF-9b
Open reading frame 9b
PAI-1
Plasminogen activator inhibitor-1
PAR1
Proteinase-activated receptor 1
PAR2
Proteinase-activated receptor 2
PKC
Protein kinase C
ROS
Reactive oxygen species
SARS-CoV-2
Severe acute respiratory syndrome coronavirus 2
TF
Tissue factor
TGF-β1
Transforming growth factor-beta 1
TLRs
Toll-like receptors
TNF-α
Tumour necrosis factor-alpha
tPA
Tissue-plasminogen activator
TRIF
TIR-Domain-containing adapter-inducing interferon

INTRODUCTION

Coronavirus disease-19 (COVID-19) poses a menace to public health with almost half a billion cases and approximately six million deaths worldwide [1, 2]. Invading the human lungs, severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) interacts with the mucous membranes across different organs, such as the eyes, nose, and mouth. Older people with comorbidities such as the metabolic syndrome and diabetes experience severe COVID-19 symptoms. Moreover, increased mortality due COVID-19 was attributed to other risk factors such as older age, diabetes, hypertension, and renal disease. For instance, more than 65% of COVID-19 patients had diabetes and cardiovascular diseases, of which 63% were above 60 years [3]. In addition, SARS-CoV-2 damages mitochondria, alters autophagy, reduces nitric oxide (NO), increasing nicotinamide adenine dinucleotide phosphate oxidases (NOX) as well as reactive oxygen species (ROS). In COVID-19, SARS-CoV-2 also activates both toll-like receptors (TLRs) and the NOD, LRR, and pyrin domain-containing protein 3 (NLRP3) inflammasome [412]. The SARS-CoV open reading frame 9b (ORF-9b) manipulates the human mitochondrial antiviral signalling molecule (MAVS) to evade the innate host immunity, limit the antiinflammatory response, and overproduce ROS [10, 13]. The NOX protein family produces ROS that enhance viral pathogenicity in inflammatory cells [10, 11]. Mammalian NOX enzymes and subunits include NOX1-5, p22phox, p67phox, NOXO1 that are elevated in response to angiotensin II (ATII) in the kidneys, heart, and endothelial cells. Such enzymes and subunits are also involved in COVID-19 [14, 15]. Infection with SARS-CoV-2 mediates inflammatory cytokines and chemokines, where ATII-induced interleukin-6 (IL-6) synthesis usually requires NOX-derived ROS [7]. Patients and mice who are NOX2-deficient had enhanced immune response with a tendency to develop autoantibodies with low ROS levels [8, 9]. The activation of TLRs and NLRP3 by SARS‐CoV‐2 induces IL‐6, IL-1β, IL-18, and lactate dehydrogenase (LDH) [4, 5, 1623]. Currently, research has discussed a higher number of involved systems in COVID-19, but from an individual perspective. Herein, the present review article combines the simultaneous detrimental effects of mitochondrial dysfunction, autophagy, NOX, NO, ROS, NLRP3, and TLRs during COVID-19 (Fig. 1). Moreover, we referred to the potential role of ROS scavengers in COVID-19.

BEHIND-THE-SCENE IN COVID-19

1.
The NOX-Mediated ROS Pathway of Inflammation
 
The dysregulation of NOX signalling is evident in COVID-19 patients with comorbidities, including obesity, diabetes, coronary artery disease, and heart failure [24]. In COVID-19 patients with acute respiratory distress syndrome (ARDS), ATII increases NOX and causes vasoconstriction and thrombosis via ROS, IL-6, tumour necrosis factor-Alpha (TNF-α), and other cytokines (Fig. 2) [25, 26]. The generation of NOX-dependent ROS elevates TNF-α, transforming growth factor-beta 1 (TGF-β1), ATII, and plasminogen activator inhibitor-1 (PAI-1), all of which are increased in COVID-19 patients [24, 2730]. Numerous endogenous and exogenous processes produce ROS, such as NOX, the electron transport chain, xanthine oxidase, smoking, heavy metals, drugs, processed meat, and radiation (Fig. 2) [31]. Interferon-Gamma (IFN-γ) and ATII in vascular smooth muscle trigger NOX1 expression, while hypoxia/ischaemia and TNF-α stimulate NOX4 [32, 33]. Endosomal NOX2 produces the proinflammatory leukotriene B4 (LTB4) and increases the levels of IL-6 and ROS in virus-mediated pathogenicity [10, 3438]. For example, influenzae A virus causes significantly less lung injury in the absence of NOX2, highlighting that NOX2-mediated ROS stimulates viral infection [35, 39]. In COVID-19, SARS-CoV-2 upregulates both ACE and ATII and therefore activates the phagocytes, metabolises haemoglobin, and causes hyperferritinaemia to produce hydroxyl radical (OH), increasing the likelihood of inflammation and thrombosis (Fig. 3) [4051].
The formation of OH correlates with oxidative stress products such as 4-hydroxynonenal and malondialdehyde guanine adducts of DNA, which also are the products of the radical oxidation of phospholipids, related to COVID-19 dyslipidaemia [5255]. Reactive oxygen species interact with lipids, carbohydrates, proteins, and nucleic acids, causing permanent destruction or alterations in their functions [56]. Hydroxyl radical is the most reactive and most toxic ROS that causes severe cellular damage by strongly interacting with DNA, carbohydrates, proteins, and lipids [5760]. Haemochromatosis in different diseases (e.g., ageing and Parkinson’s disease) has gained attention because iron catalyses the formation of OH [6164]. Hydroxyl radical directly reacts with all DNA components, such as purine and pyrimidine bases, deoxyribose sugar backbone and causes single and double stranded breaks in DNA strand breaks and chemical modifications of nucleobases or nucleotides [60, 65, 66]. The uncontrolled production of ROS significantly contributes to infectious, inflammatory, and numerous chronic disorders. This evidence underpins the current hypothesis that NOX is an essential regulator in COVID-19 pathogenesis, and that blocking the expression of NOX might hinder the production of ATII-induced ROS and IL-6, minimising inflammation and tissue injury (Fig. 2).
2.
The Inflammatory Role of NLRP3
 
The tissues of postmortem COVID-19 patients show the active NLRP3 inflammasome and its products, including IL-1β, IL-18, and LDH [1623]. Acute and chronic respiratory diseases, traumatic brain injury, acute kidney injury (AKI), and chronic kidney disease (CKD) also reported the involvement of the NLRP3 inflammasome [67]. Viral infections, metabolic abnormalities, tissue damage, and dysfunctional mitochondria generate ROS (e.g., OH) that activate the NLRP3 inflammasome, triggering the production of proinflammatory cytokines [6873]. Fortunately, mitochondria-targeted antioxidants such as molecular hydrogen (H2) can suppress the production of mitochondrial OH, and therefore inhibit the expression of NLRP3 inflammasome, caspase-1, and IL-1β [74]. Molecular hydrogen is a potent scavenger that selectively scavenges OH without adverse effects on the human body [75]. A recent multicentre trial revealed that the inhalation of hydrogen–oxygen gas mixture reduced COVID-19-related acute and chronic inflammation [76]. The intraperitoneal H2-rich saline suppressed the activation of the NLRP3 inflammasome, the activity of nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), and the production of TNF-α and IL-1β in a mouse model with acute pancreatitis. Moreover, H2-rich saline improved the survival rate and ameliorated intestinal damage and inflammatory response, oedema, and apoptosis ameliorated intestinal ischaemia/reperfusion-mediated coagulopathy in rats. Molecular hydrogen-rich saline inhibited the activation of NF-κB and NLRP3 inflammasomes in peripheral blood mononuclear cells (PBMCs) [77]. Given this, H2 may reduce the SARS-CoV-2-induced inflammation by inhibiting the NLRP3 cascade and the release of proinflammatory cytokines.
3.
The Nitric Oxide (NO)/ROS Imbalance
 
Persistent inflammation due to COVID-19 disturbs the nitric oxide (NO)/ROS balance and causes multiorgan failure [78]. Patients with COVID-19 and common comorbidities (e.g., hypertension and diabetes) displayed significantly reduced endothelial NO, suggesting a strong relationship with acute lung injury (ALI) and NO/ROS imbalance [7985]. Severe acute respiratory syndrome coronavirus 2 downregulates the expression of angiotensin-converting enzyme 2 (ACE2), producing proinflammatory cytokines and ROS that cause excessive inflammatory responses and lower the levels of NO by causing endothelial cell apoptosis (Fig. 4) [8689]. Viral SARS-CoV-2 particles easily bind their protein spikes and enter into the cells due to the higher expression of ACE-2 receptors. Hence, people with impaired metabolic health are more prone to COVID-19 and comorbidities [3]. Severely ill COVID-19 patients exhibit excessive mitochondrial ROS that lead to mitochondrial dysfunction, reducing the production and bioavailability of NO by the activation of nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB), AP-1 as well as the overexpression of cytokines and adhesion molecules (Fig. 2) [9092]. The NO donor S-nitroso-N-acetylpenicillamine (SNAP) significantly inhibited cysteine proteases encoded by SARS-CoV-1 ORF1a and the membrane fusion of offspring virus S protein, decreasing viral replication by > 80% in VeroE6 cells [9397]. Both SARS-CoV-2 and SARS-CoV exhibit a high degree of similarity in the receptor-binding domains of the spike proteins [98, 99]. Consequently, inhaled NO may prevent SARS-CoV-2 infection or treat mild, moderate, or severe COVID-19 patients, and could be used as an adjuvant therapy in mechanically ventilated patients (Fig. 4) [83, 100, 101].
4.
Mitochondrial Dysfunction and Autophagy
 
Hypoxia and other inflammatory mediators impair the function of mitochondria during COVID-19 [102, 103]. Mitochondrial dysfunction is a potential source of ROS that affect healthy mitochondria and promote cell death [104]. Mitochondria have emerged as critical dynamic organelles to maintain cellular homeostasis, metabolism, innate immune response, and determine the severity of viral infections [105]. Mitochondrial dynamics such as fusion, fission, and mitophagy protect against environmental insults; although, they are susceptible to viral infections, due to viral proteins or physiological alterations (e.g., disruption of Ca2+ homeostasis, endoplasmic reticulum stress, oxidative stress, and hypoxia) [106108]. By interfering with mitochondria, viruses distort mitochondrial functions to create a favorable stressful environment for viral proliferation (i.e., low and higher amounts of mitochondrial ATP and ROS, respectively) and impeding mitochondria-associated antiviral signaling [109]. Defective mitochondria are a potential source of ROS that can also lead to damage of healthy mitochondria. Therefore, disturbances of the rapid clearance of dysfunctional mitochondria create higher levels of ROS, promoting cell death [102, 104, 110, 111]. Afterwards, viruses (e.g., SARS-CoV-2) start to proliferate and propagate via changing potential targets, including NLRP3 inflammasome and autophagy [112].
In COVID-19-related sepsis, the SARS-CoV-2-host interaction releases the cytokine storm that ultimately leads to multiorgan failure [113]. The proinflammatory cytokine TNF-α increased mitochondrial ROS mediated by mitochondrial damage in human umbilical vein endothelial cells (HUVECs) [114]. Similarly, COVID-19 significantly upregulates TNF-α alongside other cytokines and chemokines (Figs. 1 and 3). Accordingly, SARS-CoV-2 presumably counteracts the antiviral response by upregulating TNF-α and causing mitochondrial ultrastructural abnormalities to produce higher amounts of ROS [115]. Viruses modulate mitochondria-mediated antiviral immune responses by altering autophagy, mitophagy, and cellular metabolism to facilitate their proliferation [112].
Autophagy is an essential target in SARS-COV-2-mediated COVID-19 [112]. The possible inhibition of autophagy by SARS-CoV might elaborate more the pathophysiological role of mitochondrial dysfunction during COVID-19. Cells adopt autophagy (i.e., a self-destruction mechanism) to remove dysfunctional and superfluous cellular components via the initiation and elongation of isolation membrane, autophagosomes formation, and fusion and degradation of autophagosome-lysosome [112]. Mitochondria regulate autophagy to remove harmful components by producing ROS, whereas autophagy controls mitochondrial homeostasis using mitophagy [116, 117]. The lack of normal autophagy due to viral infections leads to mitochondrial dysfunction and ROS generation (Fig. 2) [118]. Cardiovascular, neurodegenerative, chronic liver, and kidney diseases also confirmed the interaction between autophagy deterioration, mitochondrial dysfunction, and ROS generation [119122]. These data support the fact that loss of normal autophagy might be one of the primary contributors to SARS-CoV-2 infection in disturbing the mitochondrial homeostasis. However, numerous studies reported that SARS-CoV, SARS-CoV-2, Middle East respiratory syndrome coronavirus (MERS-CoV), and mouse hepatitis virus (MHV) induce and inhibit autophagy. Further research on modulating autophagy (i.e., induction or inhibition of autophagy) would elaborate the consequences on SARS-CoV-2 treatment [123132].
5.
Loss of Autophagy and ROS
 
Elderly COVID-19 patients exhibit a vulnerable antioxidant defence and an exaggerated oxidative damage. The onset of ARDS in COVID-19 patients requires the activation of the “ROS machinery” combined with innate immunity to facilitate NF-κB, exacerbating the proinflammatory host response (Fig. 2) [133]. The overproduction of ROS significantly disturbed the antioxidant system during the SARS-CoV pathogenesis, severity, and progression of the respiratory disease in vitro and in vivo [134, 135]. Humans share age-related loss of autophagy or shocking exposure to ROS. Autophagy may contribute to the ageing phenotype, denoting that ageing alters the adaptive immune response and the proinflammatory state of the host [136]. For example, older mice severely experienced SARS-CoV-induced lung lesions than younger mice [137]. Older macaques upregulate virus-host response with inflammation due to differential gene expression with NF-kB as a central player [137]. Elderly patients also had significantly higher incidence of multilobe lesions than young and middle-aged COVID-19 patients [138]. The concurrent decline in mitochondrial dysfunction due to the inhibition of autophagy and the predisposing comorbidities in elderly patients, might explain why old COVID-19 patients show severe clinical manifestations that eventually lead to multiorgan failure compared to younger patients (Fig. 2). The World Health Organization declared that currently approved medications (e.g., clozapine, glyburide, carbetapentane) could be used for the treatment of COVID-19, by targeting the NLRP3 inflammasome and autophagy to inhibit the propagation of SARS-CoV-2 [139143].
6.
The Possible Crosstalk Between TLRs, NOX, and ROS
 
Evidence supports the association between NOX, ROS, inflammatory mediators, and SARS-CoV-2 pathogenesis as well as the relationship between ROS signalling with TLR4 activation during TLR4/NOX interaction (Fig. 2) [144, 145]. The administration of diphenyleneiodonium chloride (DPI) suppressed the upregulation of TLR2, 4, and 9 in alcohol-induced fatty liver injury [146]. Human cells highlighted the potential role of NOX2 inhibitors in viral infections. In respiratory syncytial virus, rhinovirus, and human immunodeficiency virus (HIV), TLR7 activates NOX2 to produce ROS and modifies the single cysteine residue of TLR7, inhibiting the key antiviral and humoral signalling [147]. The syncytial viral cytoplasmic components recognise TLR7 and other sensor molecules; the mitochondria produce large amounts of ·OH that oxidise mitochondrial DNA, driving the cascade from NLRP3 to the release of proinflammatory cytokines (Fig. 2) [72, 148].
Severe acute respiratory syndrome coronavirus 2 binds to TLRs to activate and regulate pro-IL-1, NLRP3, IL-1β, IL-6, IL-10, and TNF-α. Such cascade causes lung inflammation and fibrosis, suggesting that the TLR pathways are protective mechanisms in SARS-CoV infections [149151]. Toll-like receptors (e.g., TLR3, 4, 7, 8, and 9) identify many viral conserved patterns where myeloid differentiation primary response 88 (MyD88)—an essential component of the TLR pathway—assembles NOX to generate ROS in neutrophils and macrophages (Fig. 2) [152, 153]. Myeloid differentiation primary response 88 activates the TIR-domain-containing adapter-inducing interferon (TRIF)‐dependent signalling to activate the IFN-1, NF‐kB, and mitogen-activated protein kinase (MAPK) pathway [154]. The activation of the TLR-MyD88 downstream signalling and NF-kB is a hallmark of SARS-CoV infections, where the inhibition of NF-kB significantly reduced respiratory coronavirus infection and increased survival in mice [151, 155].
Convalescent SARS-CoV-infected patients experienced mitochondrial- and ROS-responding gene upregulation [144]. For example, ROS/NF‐kB/TLR (mainly TL4) signalling pathways lead to ALI upon triggering by SARS-CoV. The TLR4-TRIF-TRAF6 pathogenic pathway mediates the severity of ALI. The loss of TLR4 or TRIF expression protected mice from H5N1-induced ALI, indicating that the severity of ALI depends on ROS and innate immunity.

THE POTENTIAL ROLE OF THE NOX/ROS INTERPLAY IN MEDIATING HYPOXIA, ISCHAEMIC INJURY, THROMBOSIS, AND FIBROSIS IN COVID-19

Severe hypoxia occurring during the COVID-19 cytokine storm is the leading cause of myocardial and liver damage, toxic encephalopathy, extremity ischaemia, and abnormal coagulation [156159]. Although the activation of NOX in pulmonary endothelium mediates an increase in ischaemia-mediated ROS, data remain scarce to support the role of the NOX family in hypoxia/ischaemia in COVID-19 patients [160, 161]. A murine model of coronary artery ligation showed that NOX2 led to adverse cardiac injury [162]. Rhinovirus, SARS-CoV, and the anoxia of human platelets generate NOX2-dependent ROS in vitro [163]. The genetic deletion of NOX2 quenched the cognitive deficits promoted by intermittent hypoxia and oxidative stress in mice [164, 165]. Mice transplanted with p47phox-deficient bone marrow had decreased levels of lung ischaemia and proinflammatory cytokines [166]. Apocynin—NOX2 inhibitor—reduced vascular permeability in sheep, and aborted ischaemic lung and hepatic injury, cell necrosis and tissue injury, cytokine release, and ROS production in different murine models [167173]. These data highlight that the inhibition of NOX, ROS, or p47phox could hold promise for designing effective molecules to limit the ischaemic injury in COVID-19 patients [7, 103, 174].
1.
Brain Ischaemia
 
Patients with COVID-19 present with ischaemic strokes. Brain ischaemic stroke comprises more than 80% of all strokes and occurs due to an immediate halting of blood flow by middle cerebral artery blockade [175, 176]. The excessive production of ROS aggravates oxidative stress and contributes to brain damage during ischaemia, suggesting that decreasing ROS might be helpful in the management of cerebral stroke (Fig. 1) [177180]. Studies demonstrated that NOX1, NOX2, NOX4, and NOX5 are associated with cerebral disorders and ROS release [181186]. The genetic deletion of NOX2 had protective effects against cerebral stroke in middle cerebral artery occlusion (MCAO) model. Functional NOX2-deficient and NOX2 knockout (KO) mice had significant reduction of oedema, lesion volume, and blood–brain barrier (BBB) leakage, postischaemic inflammatory gene expression and oxidative stress markers, and better neurological function during cerebral ischaemia [187190]. Mouse model of retinal ischaemia with NOX2-deficient hippocampal neurons experiences low ROS levels upon exposure to oxygen/glucose deprivation (OGD) with attenuated neuronal cell death [191]. Consequently, the treatment of stroke should adopt an effective NOX inhibitory strategy, especially NOX2. However, extensive research that simulates the human biological system is crucial to validate the data emerging from in vivo models given the small organs and the relatively large penumbra in the lesioned tissues.
2.
Thrombosis and Fibrosis
 
Microthrombosis, pulmonary embolism, endothelial failure, and disseminated intravascular coagulation (DIC) are reported in COVID-19 patients [7, 192196]. Viruses activate the coagulation pathway to overproduce proinflammatory cytokines via proteinase-activated receptors (PAR1 and PAR2) mediated by mitochondrial ROS [196201]. Both PAR1/PAR2—expressed on platelets, endothelial and epithelial cells, and vascular and nonvascular smooth muscles—are involved in inflammation [202205]. The upregulation of the NOX subunit p22phox in endothelial cells generates ROS that promote PAR1- and PAR2-mediated tissue factor (TF) induction, causing acute and chronic inflammation (Fig. 2) [206209]. During inflammation, iron (III) generate OH that convert soluble plasma fibrinogen into abnormal fibrin clots in the form of dense matted deposits resistant to enzymatic degradation (i.e., blood coagulation) (Fig. 4) [210212]. Tissue-plasminogen activator (tPA) downregulates both IL-1α and IL-1β in endothelial cells during inflammation [213, 214]. Three mechanically-ventilated COVID-19 patients demonstrated that tPA has a therapeutic role in ARDS, showing a transient improvement in the ratio of arterial oxygen partial pressure/fractional inspired oxygen [215]. However, this improvement is lost after the end of treatment due to the fact that NOX-dependent ROS inhibits tPA activity, leading to thrombosis [216, 217].
The biopsies of liver and lung injury in deceased COVID-19 patients showed severe inflammatory responses with higher levels of IL-2, IL-6, IL-8. IL-10, and IFN-γ [218]. Profibrotic responses are triggered upon the activation of PAR1 and PAR2, inducing the release of NF-kB and IL-6, IL-8, and MCP-1 that contribute to leucocyte recruitment during SARS-CoV-2 infection as well (Fig. 1) [219]. The direct upregulation of PAR (i.e., PAR2) in chronic liver disease and pulmonary fibrosis increases the production of ROS, enhancing fibrogenesis by inducing hepatocyte apoptosis, airway obstruction, and lung oedema [209, 220223]. This is consistent with the fact that PAR-2-deficient mice showed reduced inflammation and improved survival [224, 225]. Therefore, it is expected that PAR-2-dependent ROS could contribute to lung and liver injuries in COVID-19 patients, especially with predisposing diseases such as liver disease, leading to immunosuppression and disease aggression [25, 226, 227].

COULD ROS SCAVENGERS BE EFFECTIVE AGAINST COVID-19?

Natural compounds such as lycopene, polyphenols, quercetin, phloretin, berberine, and sulforaphane show a preventive potential against SARS-CoV-2 infection [228231]. The lecithinised superoxide dismutase (PC-SOD) enzyme possesses excellent bioavailability, safety (confirmed in phase I and II studies), and modulatory effect to reduce the harms of oxidative stress in COVID-19 [232236]. It is a synthetic product with long-life and high bioavailability compared to non-lecithinsed forms of the enzyme [237, 238]. For example, the intravenous administration of PC-SOD was safe and suppressed pulmonary emphysema and fibrosis, lung inflammation or ARDS, and activation of proteases, and the expression in vitro and in animal models [234, 239243]. The lecithinised superoxide dismutase reduced serum LDH and surfactant protein A in patients with stage III-IV idiopathic pulmonary fibrosis without significant side effects. It would exert a more pulmonary protective effect if administered earlier during the course of the disease [233].

CONCLUSIONS AND FUTURE DIRECTIONS

This review has shed light on the close relationship between mitochondrial dysfunction, NOX, ROS, NLRP3 inflammasome, TLRs, and NO as the “inflammatory circuit” of COVID-19. The lack of normal autophagy leads to central problems such as mitochondrial dysfunction and the production of ROS. Subsequently, there could be an interplay between autophagy and SARS-CoV-2, but the exact nature of such an interaction remains unclear.
The proposed crosstalk between ROS and NOX during SARS-CoV-2 infection unequivocally constitutes an emerging molecular analysis and drug design route for COVID-19. Other probable interfering signalling pathways (i.e., PAR, TLR-MyD88, ROS/NF‐kB/TLR, and TLR4/TRIF/TRAF6) take place during SARS-CoV-2 pathogenesis. The coronavirus proteases, especially 3C-like protease (Mpro or 3CLpro), are attractive antiviral drug targets because they are essential for coronaviral replication. Such antiviral drugs would inhibit viral replication and the dysregulation of signalling cascades in infected cells that may lead to the death of healthy cells [6].
Future investigations may unveil the mitochondrial innate antiviral signalling during COVID-19, SARS-CoV-2–host interactions, and how SARS-CoV-2 exploits alterations to the mitochondrial morphophysiology to its benefit [244, 245]. The inhibitors of NOS and ROX might be promising compounds to reduce the SARS-CoV-2-related hyperinflammatory states during the cytokine response, vascular hyperpermeability, microthrombosis, tissue injury/ischaemia and fibrosis, and multiorgan failure. Nevertheless, the essential functions of NOX and ROS in normal physiology should be considered. The use of antioxidants may face potential challenges, such as physiological interferences, biological functions of NOX/ROS, lack of target access, and the inability to attain adequate ROS concentrations.

ACKNOWLEDGEMENTS

None

DECLARATIONS

Not applicable.
Not applicable.

Conflict of Interest

The authors declare no competing interests.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Jetzt e.Med zum Sonderpreis bestellen!

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

Jetzt bestellen und 100 € sparen!

e.Dent – Das Online-Abo der Zahnmedizin

Online-Abonnement

Mit e.Dent erhalten Sie Zugang zu allen zahnmedizinischen Fortbildungen und unseren zahnmedizinischen und ausgesuchten medizinischen Zeitschriften.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Jetzt bestellen und 100 € sparen!

Literatur
1.
Zurück zum Zitat Rabi, F.A., M.S. Al Zoubi, A.D. Al-Nasser, G.A. Kasasbeh, and D.M. Salameh. 2020. Sars-cov-2 and coronavirus disease 2019: What we know so far. Pathogens 9 (3): 1–14.CrossRef Rabi, F.A., M.S. Al Zoubi, A.D. Al-Nasser, G.A. Kasasbeh, and D.M. Salameh. 2020. Sars-cov-2 and coronavirus disease 2019: What we know so far. Pathogens 9 (3): 1–14.CrossRef
3.
Zurück zum Zitat Singh, P.S., A. Bhatnagar, K.S. Singh, K.S. Patra, N. Kanwar, A. Kanwal, et al. 2022. SARS-CoV-2 Infections, Impaired Tissue, and Metabolic Health: Pathophysiology and Potential Therapeutics. Mini-Reviews in Medicinal Chemistry 22:1. Available from: http://www.eurekaselect.com/article/120610. Singh, P.S., A. Bhatnagar, K.S. Singh, K.S. Patra, N. Kanwar, A. Kanwal, et al. 2022. SARS-CoV-2 Infections, Impaired Tissue, and Metabolic Health: Pathophysiology and Potential Therapeutics. Mini-Reviews in Medicinal Chemistry 22:1. Available from: http://​www.​eurekaselect.​com/​article/​120610.
6.
Zurück zum Zitat Pratap, S.S. 2020. Clinical Application of the Main Viral Proteinase (Mpro or 3clpro) Inhibitors for Coronavirus Therapy. Biomedical Journal of Scientific and Technical Research 30 (3): 23352–23354. Pratap, S.S. 2020. Clinical Application of the Main Viral Proteinase (Mpro or 3clpro) Inhibitors for Coronavirus Therapy. Biomedical Journal of Scientific and Technical Research 30 (3): 23352–23354.
7.
Zurück zum Zitat Zhang, W., Y. Zhao, F. Zhang, Q. Wang, T. Li, Z. Liu, et al. 2020. The use of anti-inflammatory drugs in the treatment of people with severe coronavirus disease 2019 (COVID-19): The Perspectives of clinical immunologists from China. Clinical Immunology 214:108393. Available from: https://pubmed.ncbi.nlm.nih.gov/32222466. Zhang, W., Y. Zhao, F. Zhang, Q. Wang, T. Li, Z. Liu, et al. 2020. The use of anti-inflammatory drugs in the treatment of people with severe coronavirus disease 2019 (COVID-19): The Perspectives of clinical immunologists from China. Clinical Immunology 214:108393. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32222466.
9.
Zurück zum Zitat Kelkka, T., D. Kienhöfer, M. Hoffmann, M. Linja, K. Wing, O. Sareila, et al. 2014. Reactive oxygen species deficiency induces autoimmunity with type 1 interferon signature. Antioxidants and Redox Signaling 21(16):2231–45. Available from: https://pubmed.ncbi.nlm.nih.gov/24787605. Kelkka, T., D. Kienhöfer, M. Hoffmann, M. Linja, K. Wing, O. Sareila, et al. 2014. Reactive oxygen species deficiency induces autoimmunity with type 1 interferon signature. Antioxidants and Redox Signaling 21(16):2231–45. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​24787605.
12.
Zurück zum Zitat Li, T. 2020. Diagnosis and clinical management of severe acute respiratory syndrome Coronavirus 2 (SARS-CoV-2) infection: an operational recommendation of Peking Union Medical College Hospital (V2.0): Working Group of 2019 Novel Coronavirus, Peking Union Medical Colle. Emerging Microbes and Infections 9(1):582–5. Li, T. 2020. Diagnosis and clinical management of severe acute respiratory syndrome Coronavirus 2 (SARS-CoV-2) infection: an operational recommendation of Peking Union Medical College Hospital (V2.0): Working Group of 2019 Novel Coronavirus, Peking Union Medical Colle. Emerging Microbes and Infections 9(1):582–5.
14.
16.
17.
18.
Zurück zum Zitat Han, Y., H. Zhang, S. Mu, W. Wei, C. Jin, C. Tong, et al. 2020. Lactate dehydrogenase, an independent risk factor of severe COVID-19 patients: a retrospective and observational study. Aging (Albany NY). 12(12):11245–58. Available from: https://pubmed.ncbi.nlm.nih.gov/32633729. Han, Y., H. Zhang, S. Mu, W. Wei, C. Jin, C. Tong, et al. 2020. Lactate dehydrogenase, an independent risk factor of severe COVID-19 patients: a retrospective and observational study. Aging (Albany NY). 12(12):11245–58. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32633729.
22.
Zurück zum Zitat Ratajczak, M.Z., and M. Kucia. 2020. SARS-CoV-2 infection and overactivation of Nlrp3 inflammasome as a trigger of cytokine “storm” and risk factor for damage of hematopoietic stem cells. Leukemia 34(7):1726–9. Available from: https://pubmed.ncbi.nlm.nih.gov/32483300. Ratajczak, M.Z., and M. Kucia. 2020. SARS-CoV-2 infection and overactivation of Nlrp3 inflammasome as a trigger of cytokine “storm” and risk factor for damage of hematopoietic stem cells. Leukemia 34(7):1726–9. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32483300.
23.
Zurück zum Zitat Rodrigues, T.S., K.S.G. de Sá, A.Y. Ishimoto, A. Becerra, S. Oliveira, L. Almeida, et al. 2021. Inflammasomes are activated in response to SARS-CoV-2 infection and are associated with COVID-19 severity in patients. J Exp Med. 218(3):e20201707. Available from: https://pubmed.ncbi.nlm.nih.gov/33231615. Rodrigues, T.S., K.S.G. de Sá, A.Y. Ishimoto, A. Becerra, S. Oliveira, L. Almeida, et al. 2021. Inflammasomes are activated in response to SARS-CoV-2 infection and are associated with COVID-19 severity in patients. J Exp Med. 218(3):e20201707. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​33231615.
25.
28.
37.
42.
49.
Zurück zum Zitat Ivanov, A.V., V.T. Valuev-Elliston, O.N. Ivanova, S.N. Kochetkov, E.S. Starodubova, B. Bartosch, et al. 2016. Oxidative Stress during HIV Infection: Mechanisms and Consequences. Oxidative Medical and Cellular Longevity 2016:8910396. Available from: https://pubmed.ncbi.nlm.nih.gov/27829986. Ivanov, A.V., V.T. Valuev-Elliston, O.N. Ivanova, S.N. Kochetkov, E.S. Starodubova, B. Bartosch, et al. 2016. Oxidative Stress during HIV Infection: Mechanisms and Consequences. Oxidative Medical and Cellular Longevity 2016:8910396. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​27829986.
52.
Zurück zum Zitat Sorokin, A.V., S.K. Karathanasis, Z.-H. Yang, L. Freeman, K. Kotani, A.T. Remaley. 2020. COVID-19-Associated dyslipidemia: Implications for mechanism of impaired resolution and novel therapeutic approaches. FASEB Journal 34(8):9843–53. Available from: https://pubmed.ncbi.nlm.nih.gov/32588493. Sorokin, A.V., S.K. Karathanasis, Z.-H. Yang, L. Freeman, K. Kotani, A.T. Remaley. 2020. COVID-19-Associated dyslipidemia: Implications for mechanism of impaired resolution and novel therapeutic approaches. FASEB Journal 34(8):9843–53. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32588493.
54.
Zurück zum Zitat Kell, D.B., and E. Pretorius. 2014. Serum ferritin is an important inflammatory disease marker, as it is mainly a leakage product from damaged cells. Metallomics 6(4):748–73. Available from: https://doi.org/10.1039/c3mt00347g. Kell, D.B., and E. Pretorius. 2014.  Serum ferritin is an important inflammatory disease marker, as it is mainly a leakage product from damaged cells. Metallomics 6(4):748–73. Available from: https://​doi.​org/​10.​1039/​c3mt00347g.
55.
Zurück zum Zitat Nakabeppu, Y. 2014. Cellular levels of 8-oxoguanine in either DNA or the nucleotide pool play pivotal roles in carcinogenesis and survival of cancer cells. International Journal of Molecular Sciences 15(7):12543–57. Available from: https://pubmed.ncbi.nlm.nih.gov/25029543. Nakabeppu, Y. 2014. Cellular levels of 8-oxoguanine in either DNA or the nucleotide pool play pivotal roles in carcinogenesis and survival of cancer cells. International Journal of Molecular Sciences 15(7):12543–57. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​25029543.
58.
Zurück zum Zitat Petlicki, J., and T.G.M. van de Ven. 1998. The equilibrium between the oxidation of hydrogen peroxide by oxygen and the dismutation of peroxyl or superoxide radicals in aqueous solutions in contact with oxygen. Journal of Chemical Society, Faraday Transactions 94(18):2763–7. Available from: https://doi.org/10.1039/A804551H. Petlicki, J., and T.G.M. van de Ven. 1998. The equilibrium between the oxidation of hydrogen peroxide by oxygen and the dismutation of peroxyl or superoxide radicals in aqueous solutions in contact with oxygen. Journal of Chemical Society, Faraday Transactions 94(18):2763–7. Available from: https://​doi.​org/​10.​1039/​A804551H.
64.
Zurück zum Zitat Acton, R.T., J.C. Barton, L.V. Passmore, P.C. Adams, G.D. McLaren, C. Leiendecker-Foster, et al. Accuracy of family history of hemochromatosis or iron overload: the hemochromatosis and iron overload screening study. Clinical Gastroenterology and Hepatology 6(8):934–8. Available from: https://pubmed.ncbi.nlm.nih.gov/18585964. Acton, R.T., J.C. Barton, L.V. Passmore, P.C. Adams, G.D. McLaren, C. Leiendecker-Foster, et al. Accuracy of family history of hemochromatosis or iron overload: the hemochromatosis and iron overload screening study. Clinical Gastroenterology and Hepatology 6(8):934–8. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​18585964.
65.
Zurück zum Zitat Buxton, G.V., C.L. Greenstock, W.P. Helman, and A.B. Ross. 1988. Critical Review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (⋅OH/⋅O− in Aqueous Solution. Journal of Physical and Chemical Reference Data 17(2):513–886. Available from: https://doi.org/10.1063/1.555805. Buxton, G.V., C.L. Greenstock, W.P. Helman, and A.B. Ross. 1988. Critical Review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (⋅OH/⋅O− in Aqueous Solution. Journal of Physical and Chemical Reference Data 17(2):513–886. Available from: https://​doi.​org/​10.​1063/​1.​555805.
66.
69.
72.
Zurück zum Zitat Hirano, S.-I., Y. Ichikawa, B. Sato, H. Yamamoto, Y. Takefuji, and F. Satoh. 2021. Potential Therapeutic Applications of Hydrogen in Chronic Inflammatory Diseases: Possible Inhibiting Role on Mitochondrial Stress. International Journal of Molecular Sciences 22(5):2549. Available from: https://pubmed.ncbi.nlm.nih.gov/33806292. Hirano, S.-I., Y. Ichikawa, B. Sato, H. Yamamoto, Y. Takefuji, and F. Satoh. 2021. Potential Therapeutic Applications of Hydrogen in Chronic Inflammatory Diseases: Possible Inhibiting Role on Mitochondrial Stress. International Journal of Molecular Sciences 22(5):2549. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​33806292.
75.
Zurück zum Zitat Ohsawa, I., M. Ishikawa, K. Takahashi, M. Watanabe, K. Nishimaki, K. Yamagata, et al. 2007. Hydrogen acts as a therapeutic antioxidant by selectively reducing cytotoxic oxygen radicals. Nature Medicine 13(6):688–94. Available from: https://doi.org/10.1038/nm1577. Ohsawa, I., M. Ishikawa, K. Takahashi, M. Watanabe, K. Nishimaki, K. Yamagata, et al. 2007. Hydrogen acts as a therapeutic antioxidant by selectively reducing cytotoxic oxygen radicals. Nature Medicine 13(6):688–94. Available from: https://​doi.​org/​10.​1038/​nm1577.
76.
Zurück zum Zitat Guan, W.-J., C.-H. Wei, A.-L. Chen, X.-C. Sun, G.-Y. Guo, X. Zou, et al. 2020. Hydrogen/oxygen mixed gas inhalation improves disease severity and dyspnea in patients with Coronavirus disease 2019 in a recent multicenter, open-label clinical trial. Journal of Thoracic Disease 12(6):3448–52. Available from: https://pubmed.ncbi.nlm.nih.gov/32642277. Guan, W.-J., C.-H. Wei, A.-L. Chen, X.-C. Sun, G.-Y. Guo, X. Zou, et al. 2020. Hydrogen/oxygen mixed gas inhalation improves disease severity and dyspnea in patients with Coronavirus disease 2019 in a recent multicenter, open-label clinical trial. Journal of Thoracic Disease 12(6):3448–52. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32642277.
77.
Zurück zum Zitat Yang, L., Y. Guo, X. Fan, Y. Chen, B. Yang, K.-X. Liu, et al. 2020. Amelioration of Coagulation Disorders and Inflammation by Hydrogen-Rich Solution Reduces Intestinal Ischemia/Reperfusion Injury in Rats through NF-κB/NLRP3 Pathway. Mediators of Inflammation 2020:4359305. Available from: https://pubmed.ncbi.nlm.nih.gov/32587471. Yang, L., Y. Guo, X. Fan, Y. Chen, B. Yang, K.-X. Liu, et al. 2020. Amelioration of Coagulation Disorders and Inflammation by Hydrogen-Rich Solution Reduces Intestinal Ischemia/Reperfusion Injury in Rats through NF-κB/NLRP3 Pathway. Mediators of Inflammation 2020:4359305. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32587471.
79.
Zurück zum Zitat Fraser, D.D., E.K. Patterson, M. Slessarev, S.E. Gill, C. Martin, M. Daley, et al. 2020. Endothelial Injury and Glycocalyx Degradation in Critically Ill Coronavirus Disease 2019 Patients: Implications for Microvascular Platelet Aggregation. Critical Care Explorations 2(9):e0194–e0194. Available from: https://pubmed.ncbi.nlm.nih.gov/32904031. Fraser, D.D., E.K. Patterson, M. Slessarev, S.E. Gill, C. Martin, M. Daley, et al. 2020. Endothelial Injury and Glycocalyx Degradation in Critically Ill Coronavirus Disease 2019 Patients: Implications for Microvascular Platelet Aggregation. Critical Care Explorations 2(9):e0194–e0194. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32904031.
81.
86.
Zurück zum Zitat Banu, N., S.S. Panikar, L.R. Leal, and A.R. Leal. 2020. Protective role of ACE2 and its downregulation in SARS-CoV-2 infection leading to Macrophage Activation Syndrome: Therapeutic implications. Life Sciences 256:117905. Available from: https://pubmed.ncbi.nlm.nih.gov/32504757. Banu, N., S.S. Panikar, L.R. Leal, and A.R. Leal. 2020. Protective role of ACE2 and its downregulation in SARS-CoV-2 infection leading to Macrophage Activation Syndrome: Therapeutic implications. Life Sciences 256:117905. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32504757.
93.
Zurück zum Zitat Akerström, S., M. Mousavi-Jazi, J. Klingström, M. Leijon, A. Lundkvist, and A. Mirazimi. 2005. Nitric oxide inhibits the replication cycle of severe acute respiratory syndrome coronavirus. Journal of Virology 79(3):1966–9. Available from: https://pubmed.ncbi.nlm.nih.gov/15650225. Akerström, S., M. Mousavi-Jazi, J. Klingström, M. Leijon, A. Lundkvist, and A. Mirazimi. 2005. Nitric oxide inhibits the replication cycle of severe acute respiratory syndrome coronavirus. Journal of Virology 79(3):1966–9. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​15650225.
94.
Zurück zum Zitat Akerström, S., V. Gunalan, C.T. Keng, and Y.-T. Tan, A. 2009. Mirazimi. Dual effect of nitric oxide on SARS-CoV replication: viral RNA production and palmitoylation of the S protein are affected. Virology 395(1):1–9. Available from: https://pubmed.ncbi.nlm.nih.gov/19800091. Akerström, S., V. Gunalan, C.T. Keng, and Y.-T. Tan, A. 2009. Mirazimi. Dual effect of nitric oxide on SARS-CoV replication: viral RNA production and palmitoylation of the S protein are affected. Virology 395(1):1–9. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​19800091.
103.
Zurück zum Zitat Ma, J., P. Xia, Y. Zhou, Z. Liu, X. Zhou, J. Wang, et al. 2020. Potential effect of blood purification therapy in reducing cytokine storm as a late complication of critically ill COVID-19. Clinical Immunology 214:108408. Available from: https://pubmed.ncbi.nlm.nih.gov/32247038. Ma, J., P. Xia, Y. Zhou, Z. Liu, X. Zhou, J. Wang, et al. 2020. Potential effect of blood purification therapy in reducing cytokine storm as a late complication of critically ill COVID-19. Clinical Immunology 214:108408. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32247038.
107.
Zurück zum Zitat Kim, S.-J., M. Khan, J. Quan, A. Till, S. Subramani, and A. Siddiqui. 2013. Hepatitis B virus disrupts mitochondrial dynamics: induces fission and mitophagy to attenuate apoptosis. PLoS Pathogens 9(12):e1003722–e1003722. Available from: https://pubmed.ncbi.nlm.nih.gov/24339771. Kim, S.-J., M. Khan, J. Quan, A. Till, S. Subramani, and A. Siddiqui. 2013. Hepatitis B virus disrupts mitochondrial dynamics: induces fission and mitophagy to attenuate apoptosis. PLoS Pathogens 9(12):e1003722–e1003722. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​24339771.
112.
Zurück zum Zitat Singh, S.P., S. Amar, P. Gehlot, S.K. Patra, N. Kanwar, and A. Kanwal. 2021. Mitochondrial Modulations, Autophagy Pathways Shifts in Viral Infections: Consequences of COVID-19. International Journal of Molecular Sciences 22(15):8180. Available from: https://pubmed.ncbi.nlm.nih.gov/34360945. Singh, S.P., S. Amar, P. Gehlot, S.K. Patra, N. Kanwar, and A. Kanwal. 2021. Mitochondrial Modulations, Autophagy Pathways Shifts in Viral Infections: Consequences of COVID-19. International Journal of Molecular Sciences 22(15):8180. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​34360945.
114.
118.
Zurück zum Zitat Venco, P., M. Bonora, C. Giorgi, E. Papaleo, A. Iuso, H. Prokisch, et al. 2015. Mutations of C19orf12, coding for a transmembrane glycine zipper containing mitochondrial protein, cause mis-localization of the protein, inability to respond to oxidative stress and increased mitochondrial Ca2+. Frontiers in Genetics 6:185. Available from: https://www.frontiersin.org/article/10.3389/fgene.2015.00185. Venco, P., M. Bonora, C. Giorgi, E. Papaleo, A. Iuso, H. Prokisch, et al. 2015. Mutations of C19orf12, coding for a transmembrane glycine zipper containing mitochondrial protein, cause mis-localization of the protein, inability to respond to oxidative stress and increased mitochondrial Ca2+. Frontiers in Genetics 6:185. Available from: https://​www.​frontiersin.​org/​article/​10.​3389/​fgene.​2015.​00185.
123.
Zurück zum Zitat Kindrachuk, J., B. Ork, B.J. Hart, S. Mazur, M.R. Holbrook, M.B. Frieman, et al. 2015. Antiviral potential of ERK/MAPK and PI3K/AKT/mTOR signaling modulation for Middle East respiratory syndrome coronavirus infection as identified by temporal kinome analysis. Antimicrobial Agents and Chemotheraphy 59(2):1088–99. Available from: https://pubmed.ncbi.nlm.nih.gov/25487801. Kindrachuk, J., B. Ork, B.J. Hart, S. Mazur, M.R. Holbrook, M.B. Frieman, et al. 2015. Antiviral potential of ERK/MAPK and PI3K/AKT/mTOR signaling modulation for Middle East respiratory syndrome coronavirus infection as identified by temporal kinome analysis. Antimicrobial Agents and Chemotheraphy 59(2):1088–99. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​25487801.
124.
Zurück zum Zitat Reggiori, F., I. Monastyrska, M.H. Verheije, T. Calì, M. Ulasli, S. Bianchi, et al. Coronaviruses Hijack the LC3-I-positive EDEMosomes, ER-derived vesicles exporting short-lived ERAD regulators, for replication. Cell Host and Microbe 7(6):500–8. Available from: https://pubmed.ncbi.nlm.nih.gov/20542253. Reggiori, F., I. Monastyrska, M.H. Verheije, T. Calì, M. Ulasli, S. Bianchi, et al. Coronaviruses Hijack the LC3-I-positive EDEMosomes, ER-derived vesicles exporting short-lived ERAD regulators, for replication. Cell Host and Microbe 7(6):500–8. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​20542253.
125.
126.
127.
Zurück zum Zitat Gassen, N.C., Niemeyer D, Muth D, Corman VM, Martinelli S, Gassen A, et al. 2019. SKP2 attenuates autophagy through Beclin1-ubiquitination and its inhibition reduces MERS-Coronavirus infection. Nature Communications 10(1):5770. Available from: https://pubmed.ncbi.nlm.nih.gov/31852899. Gassen, N.C.,  Niemeyer D, Muth D, Corman VM, Martinelli S, Gassen A, et al. 2019. SKP2 attenuates autophagy through Beclin1-ubiquitination and its inhibition reduces MERS-Coronavirus infection. Nature Communications 10(1):5770. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​31852899.
128.
Zurück zum Zitat Keyaerts, E., L. Vijgen, P. Maes, J. Neyts, and M. Van Ranst. 2004. In vitro inhibition of severe acute respiratory syndrome coronavirus by chloroquine. Biochemical and Biophysics Research Communications 323(1):264–8. Available from: https://pubmed.ncbi.nlm.nih.gov/15351731. Keyaerts, E., L. Vijgen, P. Maes, J. Neyts, and M. Van Ranst. 2004. In vitro inhibition of severe acute respiratory syndrome coronavirus by chloroquine. Biochemical and Biophysics Research Communications 323(1):264–8. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​15351731.
129.
Zurück zum Zitat Zhu, J., W. Yu, B. Liu, Y. Wang, J. Shao, J. Wang, et al. 2017. Escin induces caspase-dependent apoptosis and autophagy through the ROS/p38 MAPK signalling pathway in human osteosarcoma cells in vitro and in vivo. Cell Death and Disease 8(10):e3113–e3113. Available from: https://pubmed.ncbi.nlm.nih.gov/29022891. Zhu, J., W. Yu, B. Liu, Y. Wang, J. Shao, J. Wang, et al. 2017. Escin induces caspase-dependent apoptosis and autophagy through the ROS/p38 MAPK signalling pathway in human osteosarcoma cells in vitro and in vivo. Cell Death and Disease 8(10):e3113–e3113. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​29022891.
130.
Zurück zum Zitat Yao, X., F. Ye, M. Zhang, C. Cui, B. Huang, P. Niu, et al. In Vitro Antiviral Activity and Projection of Optimized Dosing Design of Hydroxychloroquine for the Treatment of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). Clinical Infectious Disease 71(15):732–9. Available from: https://pubmed.ncbi.nlm.nih.gov/32150618. Yao, X., F. Ye, M. Zhang, C. Cui, B. Huang, P. Niu, et al. In Vitro Antiviral Activity and Projection of Optimized Dosing Design of Hydroxychloroquine for the Treatment of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). Clinical Infectious Disease 71(15):732–9. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32150618.
131.
Zurück zum Zitat Wu, C.-Y., J.-T. Jan, S.-H. Ma, C.-J. Kuo, H.-F. Juan, Y-S.E. Cheng, et al. 2004. Small molecules targeting severe acute respiratory syndrome human coronavirus. Proceedings of the National Academy of Sciences U S A 101(27):10012–7. Available from: https://pubmed.ncbi.nlm.nih.gov/15226499. Wu, C.-Y., J.-T. Jan, S.-H. Ma, C.-J. Kuo, H.-F. Juan, Y-S.E. Cheng, et al. 2004. Small molecules targeting severe acute respiratory syndrome human coronavirus. Proceedings of the National Academy of Sciences U S A 101(27):10012–7. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​15226499.
132.
Zurück zum Zitat Klein, B., K. Wörndl, U. Lütz-Meindl, and H.H. Kerschbaum. 2011. Perturbation of intracellular K+ homeostasis with valinomycin promotes cell death by mitochondrial swelling and autophagic processes. Apoptosis 16(11):1101. Available from: https://doi.org/10.1007/s10495-011-0642-9. Klein, B., K. Wörndl, U. Lütz-Meindl, and H.H. Kerschbaum. 2011. Perturbation of intracellular K+ homeostasis with valinomycin promotes cell death by mitochondrial swelling and autophagic processes. Apoptosis 16(11):1101. Available from: https://​doi.​org/​10.​1007/​s10495-011-0642-9.
133.
134.
135.
Zurück zum Zitat van den Brand, J.M.A., B.L. Haagmans, D. van Riel, A.D.M.E. Osterhaus, and T. Kuiken. 2014. The pathology and pathogenesis of experimental severe acute respiratory syndrome and influenza in animal models. Journal of Comparative Pathology 151(1):83–112. Available from: https://pubmed.ncbi.nlm.nih.gov/24581932. van den Brand, J.M.A., B.L. Haagmans, D. van Riel, A.D.M.E. Osterhaus, and T. Kuiken. 2014. The pathology and pathogenesis of experimental severe acute respiratory syndrome and influenza in animal models. Journal of Comparative Pathology 151(1):83–112. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​24581932.
137.
Zurück zum Zitat Smits, S.L., A. de Lang, J.M.A. van den Brand, L.M. Leijten, W.F. van IJcken, M.J.C. Eijkemans, et al. Exacerbated innate host response to SARS-CoV in aged non-human primates. PLoS Pathogens 6(2):e1000756–e1000756. Available from: https://pubmed.ncbi.nlm.nih.gov/20140198. Smits, S.L., A. de Lang, J.M.A. van den Brand, L.M. Leijten, W.F. van IJcken, M.J.C. Eijkemans, et al. Exacerbated innate host response to SARS-CoV in aged non-human primates. PLoS Pathogens 6(2):e1000756–e1000756. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​20140198.
139.
Zurück zum Zitat Albayrak, Y., and K. Hashimoto. 2017. Sigma-1 Receptor Agonists and Their Clinical Implications in Neuropsychiatric Disorders BT - Sigma Receptors: Their Role in Disease and as Therapeutic Targets. In: Smith SB, Su T-P, editors. Cham: Springer International Publishing; p. 153–61. Available from: https://doi.org/10.1007/978-3-319-50174-1_11. Albayrak, Y., and K. Hashimoto. 2017. Sigma-1 Receptor Agonists and Their Clinical Implications in Neuropsychiatric Disorders BT - Sigma Receptors: Their Role in Disease and as Therapeutic Targets. In: Smith SB, Su T-P, editors. Cham: Springer International Publishing; p. 153–61. Available from: https://​doi.​org/​10.​1007/​978-3-319-50174-1_​11.
140.
146.
Zurück zum Zitat Gustot, T., A. Lemmers, C. Moreno, N. Nagy, E. Quertinmont, C. Nicaise, et al. 2006. Differential liver sensitization to Toll-like receptor pathways in mice with alcoholic fatty liver. Hepatology 43(5):989–1000. Available from: https://doi.org/10.1002/hep.21138. Gustot, T., A. Lemmers, C. Moreno, N. Nagy, E. Quertinmont, C. Nicaise, et al. 2006. Differential liver sensitization to Toll-like receptor pathways in mice with alcoholic fatty liver. Hepatology 43(5):989–1000. Available from: https://​doi.​org/​10.​1002/​hep.​21138.
147.
149.
Zurück zum Zitat Lu, H. 2020. Drug treatment options for the 2019-new coronavirus (2019-nCoV). Bioscience Trends 14 (1): 69–71.CrossRef Lu, H. 2020. Drug treatment options for the 2019-new coronavirus (2019-nCoV). Bioscience Trends 14 (1): 69–71.CrossRef
150.
Zurück zum Zitat Jin, Y.-H., L. Cai, Z.-S. Cheng, H. Cheng, T. Deng, Y.-P. Fan, et al. 2020. A rapid advice guideline for the diagnosis and treatment of 2019 novel coronavirus (2019-nCoV) infected pneumonia (standard version). Mil Med Res [Internet]. 2020 Feb 6;7(1):4. Available from: https://pubmed.ncbi.nlm.nih.gov/32029004. Jin, Y.-H., L. Cai, Z.-S. Cheng, H. Cheng, T. Deng, Y.-P. Fan, et al. 2020. A rapid advice guideline for the diagnosis and treatment of 2019 novel coronavirus (2019-nCoV) infected pneumonia (standard version). Mil Med Res [Internet]. 2020 Feb 6;7(1):4. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32029004.
151.
Zurück zum Zitat Totura, A.L., A. Whitmore, S. Agnihothram, A. Schäfer, M.G. Katze, M.T. Heise, et al. 2015. Toll-Like Receptor 3 Signaling via TRIF Contributes to a Protective Innate Immune Response to Severe Acute Respiratory Syndrome Coronavirus Infection. MBio 6(3):e00638. Available from: https://pubmed.ncbi.nlm.nih.gov/26015500. Totura, A.L., A. Whitmore, S. Agnihothram, A. Schäfer, M.G. Katze, M.T. Heise, et al. 2015. Toll-Like Receptor 3 Signaling via TRIF Contributes to a Protective Innate Immune Response to Severe Acute Respiratory Syndrome Coronavirus Infection. MBio 6(3):e00638. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​26015500.
154.
Zurück zum Zitat DeDiego, M.L., J.L. Nieto-Torres, J.A. Regla-Nava, J.M. Jimenez-Guardeño, R. Fernandez-Delgado, C. Fett, et al. 2013. Inhibition of NF-κB-mediated inflammation in severe acute respiratory syndrome coronavirus-infected mice increases survival. Journal of Virology 88(2):913–24. Available from: https://pubmed.ncbi.nlm.nih.gov/24198408. DeDiego, M.L., J.L. Nieto-Torres, J.A. Regla-Nava, J.M. Jimenez-Guardeño, R. Fernandez-Delgado, C. Fett, et al. 2013. Inhibition of NF-κB-mediated inflammation in severe acute respiratory syndrome coronavirus-infected mice increases survival. Journal of Virology 88(2):913–24. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​24198408.
159.
Zurück zum Zitat Zhang, Y., L. Zheng, L. Liu, M. Zhao, J. Xiao, and Q. Zhao. 2020. Liver impairment in COVID-19 patients: A retrospective analysis of 115 cases from a single centre in Wuhan city, China. Liver International 40(9):2095–103. Available from: https://doi.org/10.1111/liv.14455. Zhang, Y., L. Zheng, L. Liu, M. Zhao, J. Xiao, and Q. Zhao. 2020. Liver impairment in COVID-19 patients: A retrospective analysis of 115 cases from a single centre in Wuhan city, China. Liver International 40(9):2095–103. Available from: https://​doi.​org/​10.​1111/​liv.​14455.
160.
161.
Zurück zum Zitat Weissmann, N., A. Sydykov, H. Kalwa, U. Storch, B. Fuchs, M. Mederos y Schnitzler, et al. 2012. Activation of TRPC6 channels is essential for lung ischaemia–reperfusion induced oedema in mice. Nature Communications 3(1):649. Available from: https://doi.org/10.1038/ncomms1660. Weissmann, N., A. Sydykov, H. Kalwa, U. Storch, B. Fuchs, M. Mederos y Schnitzler, et al. 2012. Activation of TRPC6 channels is essential for lung ischaemia–reperfusion induced oedema in mice. Nature Communications 3(1):649. Available from: https://​doi.​org/​10.​1038/​ncomms1660.
163.
Zurück zum Zitat Basili, S., P. Pignatelli, G. Tanzilli, E. Mangieri, R. Carnevale, C. Nocella, et al. 2011. Anoxia-Reoxygenation Enhances Platelet Thromboxane A2 Production via Reactive Oxygen Species–Generated NOX2. Arteriosclerosis, Thrombosis, and Vascular Biology 31(8):1766–71. Available from: https://doi.org/10.1161/ATVBAHA.111.227959. Basili, S., P. Pignatelli, G. Tanzilli, E. Mangieri, R. Carnevale, C. Nocella, et al. 2011. Anoxia-Reoxygenation Enhances Platelet Thromboxane A2 Production via Reactive Oxygen Species–Generated NOX2. Arteriosclerosis, Thrombosis, and Vascular Biology 31(8):1766–71. Available from: https://​doi.​org/​10.​1161/​ATVBAHA.​111.​227959.
165.
Zurück zum Zitat Nair. D., E.A. Dayyat, S.X. Zhang, Y. Wang, and D. Gozal. 2011. Intermittent hypoxia-induced cognitive deficits are mediated by NADPH oxidase activity in a murine model of sleep apnea. PLoS One 6(5):e19847–e19847. Available from: https://pubmed.ncbi.nlm.nih.gov/21625437. Nair. D., E.A. Dayyat, S.X. Zhang, Y. Wang, and D. Gozal. 2011. Intermittent hypoxia-induced cognitive deficits are mediated by NADPH oxidase activity in a murine model of sleep apnea. PLoS One 6(5):e19847–e19847. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​21625437.
166.
Zurück zum Zitat Yang, Z., A.K. Sharma, M. Marshall, I.L. Kron, and V.E. Laubach. 2009. NADPH oxidase in bone marrow-derived cells mediates pulmonary ischemia-reperfusion injury. American Journal of Respiratory Cell and Molecular Biology 40(3):375–81. Available from: https://pubmed.ncbi.nlm.nih.gov/18787174. Yang, Z., A.K. Sharma, M. Marshall, I.L. Kron, and V.E. Laubach. 2009. NADPH oxidase in bone marrow-derived cells mediates pulmonary ischemia-reperfusion injury. American Journal of Respiratory Cell and Molecular Biology 40(3):375–81. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​18787174.
169.
170.
Zurück zum Zitat Shiotani, S., M. Shimada, A. Taketomi, Y. Soejima, T. Yoshizumi, K. Hashimoto, et al. 2007. Rho-kinase as a novel gene therapeutic target in treatment of cold ischemia/reperfusion-induced acute lethal liver injury: effect on hepatocellular NADPH oxidase system. Gene Therapy 14(19):1425–33. Available from: https://doi.org/10.1038/sj.gt.3303000. Shiotani, S., M. Shimada, A. Taketomi, Y. Soejima, T. Yoshizumi, K. Hashimoto, et al. 2007. Rho-kinase as a novel gene therapeutic target in treatment of cold ischemia/reperfusion-induced acute lethal liver injury: effect on hepatocellular NADPH oxidase system. Gene Therapy 14(19):1425–33. Available from: https://​doi.​org/​10.​1038/​sj.​gt.​3303000.
171.
Zurück zum Zitat Csányi, G., E. Cifuentes-Pagano, I. Al Ghouleh, D.J. Ranayhossaini, L. Egaña, L.R. Lopes, et al. 2011. Nox2 B-loop peptide, Nox2ds, specifically inhibits the NADPH oxidase Nox2. Free Radical Biology and Medicine 51(6):1116–25. Available from: https://pubmed.ncbi.nlm.nih.gov/21586323. Csányi, G., E. Cifuentes-Pagano, I. Al Ghouleh, D.J. Ranayhossaini, L. Egaña, L.R. Lopes, et al. 2011. Nox2 B-loop peptide, Nox2ds, specifically inhibits the NADPH oxidase Nox2. Free Radical Biology and Medicine 51(6):1116–25. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​21586323.
172.
Zurück zum Zitat Dorman, R.B., C. Wunder, H. Saba, J.L. Shoemaker, L.A. MacMillan-Crow, and R.W. Brock. 2006. NAD(P)H oxidase contributes to the progression of remote hepatic parenchymal injury and endothelial dysfunction, but not microvascular perfusion deficits. American Journal of Physiology Liver Physiology 290(5):G1025–32. Available from: https://doi.org/10.1152/ajpgi.00246.2005. Dorman, R.B., C. Wunder, H. Saba, J.L. Shoemaker, L.A. MacMillan-Crow, and R.W. Brock. 2006. NAD(P)H oxidase contributes to the progression of remote hepatic parenchymal injury and endothelial dysfunction, but not microvascular perfusion deficits. American Journal of Physiology Liver Physiology 290(5):G1025–32. Available from: https://​doi.​org/​10.​1152/​ajpgi.​00246.​2005.
173.
Zurück zum Zitat Paterniti, I., M. Galuppo, E. Mazzon, D. Impellizzeri, E. Esposito, P. Bramanti, et al. 2010. Protective effects of apocynin, an inhibitor of NADPH oxidase activity, in splanchnic artery occlusion and reperfusion. Journal of Leukocyte Biology 88(5):993–1003. Available from: https://doi.org/10.1189/jlb.0610322. Paterniti, I., M. Galuppo, E. Mazzon, D. Impellizzeri, E. Esposito, P. Bramanti, et al. 2010. Protective effects of apocynin, an inhibitor of NADPH oxidase activity, in splanchnic artery occlusion and reperfusion. Journal of Leukocyte Biology 88(5):993–1003. Available from: https://​doi.​org/​10.​1189/​jlb.​0610322.
176.
Zurück zum Zitat Hacke, W., M. Kaste, E. Bluhmki, M. Brozman, A. Dávalos, D. Guidetti, et al. 2008. Thrombolysis with Alteplase 3 to 4.5 Hours after Acute Ischemic Stroke. New England Journal of Medicine 359(13):1317–29. Available from: https://doi.org/10.1056/NEJMoa0804656. Hacke, W., M. Kaste, E. Bluhmki, M. Brozman, A. Dávalos, D. Guidetti, et al. 2008. Thrombolysis with Alteplase 3 to 4.5 Hours after Acute Ischemic Stroke. New England Journal of Medicine 359(13):1317–29. Available from: https://​doi.​org/​10.​1056/​NEJMoa0804656.
177.
178.
180.
183.
Zurück zum Zitat Montezano, A.C., D. Burger, T.M. Paravicini, A.Z. Chignalia, H. Yusuf, M. Almasri, et al. 2010. Nicotinamide adenine dinucleotide phosphate reduced oxidase 5 (Nox5) regulation by angiotensin II and endothelin-1 is mediated via calcium/calmodulin-dependent, rac-1-independent pathways in human endothelial cells. Circulation Research 106(8):1363–73. Available from: https://pubmed.ncbi.nlm.nih.gov/20339118. Montezano, A.C., D. Burger, T.M. Paravicini, A.Z. Chignalia, H. Yusuf, M. Almasri, et al. 2010. Nicotinamide adenine dinucleotide phosphate reduced oxidase 5 (Nox5) regulation by angiotensin II and endothelin-1 is mediated via calcium/calmodulin-dependent, rac-1-independent pathways in human endothelial cells. Circulation Research 106(8):1363–73. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​20339118.
190.
Zurück zum Zitat Yoshioka, H., K. Niizuma, M. Katsu, N. Okami, H. Sakata, G.S. Kim, et al. 2010. Nadph Oxidase Mediates Striatal Neuronal Injury after Transient Global Cerebral Ischemia. Journal of Cerebral Blood Flow and Metabolism 31(3):868–80. Available from: https://doi.org/10.1038/jcbfm.2010.166. Yoshioka, H., K. Niizuma, M. Katsu, N. Okami, H. Sakata, G.S. Kim, et al. 2010. Nadph Oxidase Mediates Striatal Neuronal Injury after Transient Global Cerebral Ischemia. Journal of Cerebral Blood Flow and Metabolism 31(3):868–80. Available from: https://​doi.​org/​10.​1038/​jcbfm.​2010.​166.
191.
Zurück zum Zitat Yokota, H., S.P. Narayanan, W. Zhang, H. Liu, M. Rojas, Z. Xu, et al. 2011. Neuroprotection from retinal ischemia/reperfusion injury by NOX2 NADPH oxidase deletion. Investigative Ophthalmology and Visual Science 52(11):8123–31. Available from: https://pubmed.ncbi.nlm.nih.gov/21917939. Yokota, H., S.P. Narayanan, W. Zhang, H. Liu, M. Rojas, Z. Xu, et al. 2011. Neuroprotection from retinal ischemia/reperfusion injury by NOX2 NADPH oxidase deletion. Investigative Ophthalmology and Visual Science 52(11):8123–31. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​21917939.
193.
Zurück zum Zitat Fox, S.E., A. Akmatbekov, J.L. Harbert, G. Li, J. Quincy Brown, and R.S. Vander Heide. 2020. Pulmonary and cardiac pathology in African American patients with COVID-19: an autopsy series from New Orleans. Lancet Respiratory Medicine 8(7):681–6. Available from: https://pubmed.ncbi.nlm.nih.gov/32473124. Fox, S.E., A. Akmatbekov, J.L. Harbert, G. Li, J. Quincy Brown, and R.S. Vander Heide. 2020. Pulmonary and cardiac pathology in African American patients with COVID-19: an autopsy series from New Orleans. Lancet Respiratory Medicine 8(7):681–6. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32473124.
195.
Zurück zum Zitat Zhou, F., T. Yu, R. Du, G. Fan, Y. Liu, Z. Liu, et al. 2020. Clinical course and risk factors for mortality of adult inpatients with COVID-19 in Wuhan, China: a retrospective cohort study. Lancet (London, England) 395(10229):1054–62. Available from: https://pubmed.ncbi.nlm.nih.gov/32171076. Zhou, F., T. Yu, R. Du, G. Fan, Y. Liu, Z. Liu, et al. 2020. Clinical course and risk factors for mortality of adult inpatients with COVID-19 in Wuhan, China: a retrospective cohort study. Lancet (London, England) 395(10229):1054–62. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32171076.
196.
197.
198.
Zurück zum Zitat Yang, X., Y. Yu, J. Xu, H. Shu, J. Xia, H. Liu, et al. 2020. Clinical course and outcomes of critically ill patients with SARS-CoV-2 pneumonia in Wuhan, China: a single-centered, retrospective, observational study. Lancet Respiratory Medicine 8(5):475–81. Available from: https://pubmed.ncbi.nlm.nih.gov/32105632. Yang, X., Y. Yu, J. Xu, H. Shu, J. Xia, H. Liu, et al. 2020. Clinical course and outcomes of critically ill patients with SARS-CoV-2 pneumonia in Wuhan, China: a single-centered, retrospective, observational study. Lancet Respiratory Medicine 8(5):475–81. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32105632.
199.
Zurück zum Zitat Perrone, L.A., J.A. Belser, D.A. Wadford, J.M. Katz, and T.M. Tumpey. 2013. Inducible Nitric Oxide Contributes to Viral Pathogenesis Following Highly Pathogenic Influenza Virus Infection in Mice. Journal of Infectious Disease 207(10):1576–84. Available from: https://doi.org/10.1093/infdis/jit062. Perrone, L.A., J.A. Belser, D.A. Wadford, J.M. Katz, and T.M. Tumpey. 2013. Inducible Nitric Oxide Contributes to Viral Pathogenesis Following Highly Pathogenic Influenza Virus Infection in Mice. Journal of Infectious Disease 207(10):1576–84. Available from: https://​doi.​org/​10.​1093/​infdis/​jit062.
202.
Zurück zum Zitat R. D’Andrea, M., C.K. Derian, D. Leturcq, S.M. Baker SM, Brunmark A, Ling P, et al. 1998. Characterization of Protease-activated Receptor-2 Immunoreactivity in Normal Human Tissues. Journal of Histochemistry and Cytochemistry 46(2):157–64. Available from: https://doi.org/10.1177/002215549804600204. R. D’Andrea, M., C.K. Derian, D. Leturcq, S.M. Baker SM, Brunmark A, Ling P, et al. 1998. Characterization of Protease-activated Receptor-2 Immunoreactivity in Normal Human Tissues. Journal of Histochemistry and Cytochemistry 46(2):157–64. Available from: https://​doi.​org/​10.​1177/​0022155498046002​04.
208.
Zurück zum Zitat Ott, I. 2003. Tissue Factor in Acute Coronary Syndromes. Semin Vasc Med. 03 (02): 185–192.CrossRef Ott, I. 2003. Tissue Factor in Acute Coronary Syndromes. Semin Vasc Med. 03 (02): 185–192.CrossRef
209.
Zurück zum Zitat Banfi, C., M. Brioschi, S.S. Barbieri, S. Eligini, S. Barcella, E. Tremoli, et al. 2009. Mitochondrial reactive oxygen species: a common pathway for PAR1- and PAR2-mediated tissue factor induction in human endothelial cells. Journal of Thrombosis and Haemostasis 7(1):206–16. Available from: https://doi.org/10.1111/j.1538-7836.2008.03204.x. Banfi, C., M. Brioschi, S.S. Barbieri, S. Eligini, S. Barcella, E. Tremoli, et al. 2009. Mitochondrial reactive oxygen species: a common pathway for PAR1- and PAR2-mediated tissue factor induction in human endothelial cells. Journal of Thrombosis and Haemostasis 7(1):206–16. Available from: https://​doi.​org/​10.​1111/​j.​1538-7836.​2008.​03204.​x.
211.
Zurück zum Zitat Schaer, D.J., P.W. Buehler, A.I. Alayash, J.D. Belcher, and G.M. Vercellotti. 2013 Hemolysis and free hemoglobin revisited: exploring hemoglobin and hemin scavengers as a novel class of therapeutic proteins. Blood 121(8):1276–84. Available from: https://pubmed.ncbi.nlm.nih.gov/23264591. Schaer, D.J., P.W. Buehler, A.I. Alayash, J.D. Belcher, and G.M. Vercellotti. 2013 Hemolysis and free hemoglobin revisited: exploring hemoglobin and hemin scavengers as a novel class of therapeutic proteins. Blood 121(8):1276–84. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​23264591.
212.
213.
Zurück zum Zitat Bevilacqua, M.P., R.R. Schleef, M.A. Gimbrone Jr, and D.J. Loskutoff. 1986. Regulation of the fibrinolytic system of cultured human vascular endothelium by interleukin 1. Journal of Clinical Investigations 78(2):587–91. Available from: https://pubmed.ncbi.nlm.nih.gov/3090105. Bevilacqua, M.P., R.R. Schleef, M.A. Gimbrone Jr, and D.J. Loskutoff. 1986. Regulation of the fibrinolytic system of cultured human vascular endothelium by interleukin 1. Journal of Clinical Investigations 78(2):587–91. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​3090105.
214.
215.
Zurück zum Zitat Wang, J., N. Hajizadeh, E.E. Moore, R.C. McIntyre, P.K. Moore, L.A. Veress, et al. 2020. Tissue plasminogen activator (tPA) treatment for COVID-19 associated acute respiratory distress syndrome (ARDS): A case series. Journal of Thrombosis and Haemostasis 18(7):1752–5. Available from: https://pubmed.ncbi.nlm.nih.gov/32267998. Wang, J., N. Hajizadeh, E.E. Moore, R.C. McIntyre, P.K. Moore, L.A. Veress, et al. 2020. Tissue plasminogen activator (tPA) treatment for COVID-19 associated acute respiratory distress syndrome (ARDS): A case series. Journal of Thrombosis and Haemostasis 18(7):1752–5. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​32267998.
216.
217.
Zurück zum Zitat Van Guilder, G.P., G.L. Hoetzer, J.J. Greiner, B.L. Stauffer, and C.A. DeSouza. 2008. Acute and chronic effects of vitamin C on endothelial fibrinolytic function in overweight and obese adult humans. Journal of Physiology 586(14):3525–35. Available from: https://pubmed.ncbi.nlm.nih.gov/18499730. Van Guilder, G.P., G.L. Hoetzer, J.J. Greiner, B.L. Stauffer, and C.A. DeSouza. 2008. Acute and chronic effects of vitamin C on endothelial fibrinolytic function in overweight and obese adult humans. Journal of Physiology 586(14):3525–35. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​18499730.
218.
Zurück zum Zitat Rabaan, A.A., S.H. Al-Ahmed, J. Muhammad, A. Khan, A.A. Sule, R. Tirupathi, et al. 2021. Role of Inflammatory Cytokines in COVID-19 Patients: A Review on Molecular Mechanisms, Immune Functions, Immunopathology and Immunomodulatory Drugs to Counter Cytokine Storm. Vaccines 9(5):436. Available from: https://pubmed.ncbi.nlm.nih.gov/33946736. Rabaan, A.A., S.H. Al-Ahmed, J. Muhammad, A. Khan, A.A. Sule, R. Tirupathi, et al. 2021. Role of Inflammatory Cytokines in COVID-19 Patients: A Review on Molecular Mechanisms, Immune Functions, Immunopathology and Immunomodulatory Drugs to Counter Cytokine Storm. Vaccines 9(5):436. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​33946736.
220.
Zurück zum Zitat Borensztajn, K., J.H. von der Thüsen, M.P. Peppelenbosch, and C.A. Spek. 2010. The coagulation factor Xa/protease activated receptor-2 axis in the progression of liver fibrosis: a multifaceted paradigm. Journal of Cell and Molecular Medicine 14(1–2):143–53. Available from: https://pubmed.ncbi.nlm.nih.gov/19968736. Borensztajn, K., J.H. von der Thüsen, M.P. Peppelenbosch, and C.A. Spek. 2010. The coagulation factor Xa/protease activated receptor-2 axis in the progression of liver fibrosis: a multifaceted paradigm. Journal of  Cell and Molecular Medicine 14(1–2):143–53. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​19968736.
221.
225.
Zurück zum Zitat Takizawa, T., M. Tamiya, T. Hara, J. Matsumoto, N. Saito, T. Kanke, et al. 2005. Abrogation of bronchial eosinophilic inflammation and attenuated eotaxin content in protease-activated receptor 2-deficient mice. Journal of Pharmacological Sciences 98(1):99–102. Available from: http://europepmc.org/abstract/MED/15879675. Takizawa, T., M. Tamiya, T. Hara, J. Matsumoto, N. Saito, T. Kanke, et al. 2005. Abrogation of bronchial eosinophilic inflammation and attenuated eotaxin content in protease-activated receptor 2-deficient mice. Journal of Pharmacological Sciences 98(1):99–102. Available from: http://​europepmc.​org/​abstract/​MED/​15879675.
231.
232.
233.
Zurück zum Zitat Kamio, K., A. Azuma, K. Ohta, Y. Sugiyama, T. Nukiwa, S. Kudoh, et al. 2014. Double-blind controlled trial of lecithinized superoxide dismutase in patients with idiopathic interstitial pneumonia - short term evaluation of safety and tolerability. BMC Pulmonary Medicine 14:86. Available from: https://pubmed.ncbi.nlm.nih.gov/24886036. Kamio, K., A. Azuma, K. Ohta, Y. Sugiyama, T. Nukiwa, S. Kudoh, et al. 2014. Double-blind controlled trial of lecithinized superoxide dismutase in patients with idiopathic interstitial pneumonia - short term evaluation of safety and tolerability. BMC Pulmonary Medicine 14:86. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​24886036.
234.
Zurück zum Zitat Broeyer, F.J.F., B.E. van Aken, J. Suzuki, M.J.B. Kemme, H.C. Schoemaker, A.F. Cohen, et al. 2008. The pharmacokinetics and effects of a long-acting preparation of superoxide dismutase (PC-SOD) in man. Brirtish Journal of Clinical Pharmacology 65(1):22–9. Available from: https://pubmed.ncbi.nlm.nih.gov/17610527. Broeyer, F.J.F., B.E. van Aken, J. Suzuki, M.J.B. Kemme, H.C. Schoemaker, A.F. Cohen, et al. 2008. The pharmacokinetics and effects of a long-acting preparation of superoxide dismutase (PC-SOD) in man. Brirtish Journal of Clinical Pharmacology 65(1):22–9. Available from: https://​pubmed.​ncbi.​nlm.​nih.​gov/​17610527.
235.
Zurück zum Zitat Suzuki, J., F. Broeyer, A. Cohen, M. Takebe, J. Burggraaf, and Y. Mizushima. 2008. Pharmacokinetics of PC-SOD, a Lecithinized Recombinant Superoxide Dismutase, After Single- and Multiple-Dose Administration to Healthy Japanese and Caucasian Volunteers. Journal of Clinical Pharmacology 48(2):184–92. Available from: https://doi.org/10.1177/0091270007309705. Suzuki, J., F. Broeyer, A. Cohen, M. Takebe, J. Burggraaf, and Y. Mizushima. 2008. Pharmacokinetics of PC-SOD, a Lecithinized Recombinant Superoxide Dismutase, After Single- and Multiple-Dose Administration to Healthy Japanese and Caucasian Volunteers. Journal of Clinical Pharmacology 48(2):184–92. Available from: https://​doi.​org/​10.​1177/​0091270007309705​.
240.
Zurück zum Zitat Tanaka, K.-I., T. Ishihara, A. Azuma, S. Kudoh, M. Ebina, T. Nukiwa, et al. 2009. Therapeutic effect of lecithinized superoxide dismutase on bleomycin-induced pulmonary fibrosis. American Journal of Physiology Cellular and Molecular Physiology 298(3):L348–60. Available from: https://doi.org/10.1152/ajplung.00289.2009. Tanaka, K.-I., T. Ishihara, A. Azuma, S. Kudoh, M. Ebina, T. Nukiwa, et al. 2009. Therapeutic effect of lecithinized superoxide dismutase on bleomycin-induced pulmonary fibrosis. American Journal of Physiology Cellular and Molecular Physiology 298(3):L348–60. Available from: https://​doi.​org/​10.​1152/​ajplung.​00289.​2009.
241.
Zurück zum Zitat Tanaka, K.-I., K. Sato, K. Aoshiba, A. Azuma, and T. Mizushima. 2012. Superiority of PC-SOD to other anti-COPD drugs for elastase-induced emphysema and alteration in lung mechanics and respiratory function in mice. American Journal of Physiology Cellular and Molecular Physiology 302(12):L1250–61. Available from: https://doi.org/10.1152/ajplung.00019.2012. Tanaka, K.-I., K. Sato, K. Aoshiba, A. Azuma, and T. Mizushima. 2012. Superiority of PC-SOD to other anti-COPD drugs for elastase-induced emphysema and alteration in lung mechanics and respiratory function in mice. American Journal of Physiology Cellular and Molecular Physiology 302(12):L1250–61. Available from: https://​doi.​org/​10.​1152/​ajplung.​00019.​2012.
242.
Zurück zum Zitat Tanaka, K., F. Tamura, T. Sugizaki, M. Kawahara, K. Kuba, Y. Imai, et al. 2016. Evaluation of Lecithinized Superoxide Dismutase for the Prevention of Acute Respiratory Distress Syndrome in Animal Models. American Journal of Respiratory Cellular and Molecular Biology 56(2):179–90. Available from: https://doi.org/10.1165/rcmb.2016-0158OC. Tanaka, K., F. Tamura, T. Sugizaki, M. Kawahara, K. Kuba, Y. Imai, et al. 2016. Evaluation of Lecithinized Superoxide Dismutase for the Prevention of Acute Respiratory Distress Syndrome in Animal Models. American Journal of Respiratory Cellular and Molecular Biology 56(2):179–90. Available from: https://​doi.​org/​10.​1165/​rcmb.​2016-0158OC.
244.
Zurück zum Zitat Vasquez-Bonilla, W.O., R. Orozco, V. Argueta, M. Sierra, L.I. Zambrano, F. Muñoz-Lara, D.S. López-Molina, K. Arteaga-Livias, Z. Grimes, C. Bryce, A. Paniz-Mondolfi, and A.J. Rodríguez-Morales. 2020. A review of the main histopathological findings in coronavirus disease 2019. Human Pathology 105:74–83. https://doi.org/10.1016/j.humpath.2020.07.023. Epub 2020 Aug 2. PMID: 32750378; PMCID: PMC7395947. Vasquez-Bonilla, W.O., R. Orozco, V. Argueta, M. Sierra, L.I. Zambrano, F. Muñoz-Lara, D.S. López-Molina, K. Arteaga-Livias, Z. Grimes, C. Bryce, A. Paniz-Mondolfi, and A.J. Rodríguez-Morales. 2020. A review of the main histopathological findings in coronavirus disease 2019. Human Pathology 105:74–83. https://​doi.​org/​10.​1016/​j.​humpath.​2020.​07.​023. Epub 2020 Aug 2. PMID: 32750378; PMCID: PMC7395947.
245.
Zurück zum Zitat Iqbal Yatoo, M., Z. Hamid, O.R. Parray, A.H. Wani, A. Ul Haq, A. Saxena, S.K. Patel, M. Pathak, R. Tiwari, Y.S. Malik, R. Sah, A.A. Rabaan, A.J. Rodriguez Morales, and K. Dhama. 2020. COVID-19 - Recent advancements in identifying novel vaccine candidates and current status of upcoming SARS-CoV-2 vaccines. Human Vaccines and Immunotherapeutics 16(12):2891–2904. https://doi.org/10.1080/21645515.2020.1788310. Epub 2020 Jul 23. PMID: 32703064; PMCID: PMC8641591. Iqbal Yatoo, M., Z. Hamid, O.R. Parray, A.H. Wani, A. Ul Haq, A. Saxena, S.K. Patel, M. Pathak, R. Tiwari, Y.S. Malik, R. Sah, A.A. Rabaan, A.J. Rodriguez Morales, and K. Dhama. 2020. COVID-19 - Recent advancements in identifying novel vaccine candidates and current status of upcoming SARS-CoV-2 vaccines. Human Vaccines and Immunotherapeutics 16(12):2891–2904. https://​doi.​org/​10.​1080/​21645515.​2020.​1788310. Epub 2020 Jul 23. PMID: 32703064; PMCID: PMC8641591.
Metadaten
Titel
Disengaging the COVID-19 Clutch as a Discerning Eye Over the Inflammatory Circuit During SARS-CoV-2 Infection
verfasst von
Mohammed Moustapha Anwar
Ranjit Sah
Sunil Shrestha
Akihiko Ozaki
Namrata Roy
Zareena Fathah
Alfonso J. Rodriguez-Morales
Publikationsdatum
30.05.2022
Verlag
Springer US
Schlagwort
COVID-19
Erschienen in
Inflammation / Ausgabe 5/2022
Print ISSN: 0360-3997
Elektronische ISSN: 1573-2576
DOI
https://doi.org/10.1007/s10753-022-01674-5

Weitere Artikel der Ausgabe 5/2022

Inflammation 5/2022 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Erhebliches Risiko für Kehlkopfkrebs bei mäßiger Dysplasie

29.05.2024 Larynxkarzinom Nachrichten

Fast ein Viertel der Personen mit mäßig dysplastischen Stimmlippenläsionen entwickelt einen Kehlkopftumor. Solche Personen benötigen daher eine besonders enge ärztliche Überwachung.

Nach Herzinfarkt mit Typ-1-Diabetes schlechtere Karten als mit Typ 2?

29.05.2024 Herzinfarkt Nachrichten

Bei Menschen mit Typ-2-Diabetes sind die Chancen, einen Myokardinfarkt zu überleben, in den letzten 15 Jahren deutlich gestiegen – nicht jedoch bei Betroffenen mit Typ 1.

15% bedauern gewählte Blasenkrebs-Therapie

29.05.2024 Urothelkarzinom Nachrichten

Ob Patienten und Patientinnen mit neu diagnostiziertem Blasenkrebs ein Jahr später Bedauern über die Therapieentscheidung empfinden, wird einer Studie aus England zufolge von der Radikalität und dem Erfolg des Eingriffs beeinflusst.

Costims – das nächste heiße Ding in der Krebstherapie?

28.05.2024 Onkologische Immuntherapie Nachrichten

„Kalte“ Tumoren werden heiß – CD28-kostimulatorische Antikörper sollen dies ermöglichen. Am besten könnten diese in Kombination mit BiTEs und Checkpointhemmern wirken. Erste klinische Studien laufen bereits.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.