Please cite this paper as:

Boga, J.A. and Coto-Montes, A. 2020. ER stress and autophagy induced by SARS-CoV-2: The targets for melatonin treatment. Melatonin Research. 3, 3 (Jun. 2020), 346-361. DOI:https://doi.org/https://doi.org/10.32794/mr11250067.


Review

ER stress and autophagy induced by SARS-CoV-2: The targets for melatonin treatment

Ana Coto-Montes1,3a, José Antonio Boga2,3a*

1Department of Morphology and Cell Biology, Faculty of Medicine, University of Oviedo, Avda. Julián Clavería s/n, 33006, Oviedo, Principado de Asturias, Spain.

2Microbiology Service, Hospital Universitario Central de Asturias, Avda. Roma, s/n, 33011, Oviedo, Principado de Asturias, Spain.

3Instituto de Investigación Sanitaria del Principado de Asturias (ISPA), Avda. Roma, s/n, 33011, Oviedo, Principado de Asturias, Spain

aMembers of Research Consortium cROS (cellular Response to Oxidative Stress).

*Correspondence: joseantonio.boga@sespa.es, Tel: +34-985108720

Running Title: Melatonin as treatment against COVID19

Received: May 04, 2020; Accepted: June 7, 2020

 

ABSTRACT

      Coronavirus disease 19 (COVID-19) is a viral disease caused by the new coronavirus SARS-CoV-2. Like other coronaviral infections, SARS-CoV-2 causes oxidative and ER stress triggering cellular response pathways, mainly PERK and IRE1 branches of the UPR. This excessive oxidative stress and the increasing of unfolded and misfolded proteins induce autophagy. Once this process is triggered, the blockage of the fusion of autophagosomes and lysosomes induced by virus leads to an incomplete autophagy. Double-membraned vesicles, which create a membranous support for viral RNA replication complexes, are formed. Melatonin is a pleiotropic molecule, which reduces oxidative and ER stress, regulates immune system, and modulates autophagy pathway. Thus, melatonin reinforces UPR and unlocks autophagy blockage, allowing autophagosomes to bind to lysosomes, completing the process of autophagy and decreasing viral replication capacity. Based on these activities of melatonin the recommendation of melatonin for patients with COVID-19 should be seriously considered, especially in elderlies and patients with different comorbidities, which are the highest risk population for serious cases.

 

Key words: Sars-CoV2, ER-stress, autophagy, melatonin, unfolded protein response (UPR), frailty, aging, viral infection.

__________________________________________________________________________

 

1. INTRODUCTION

     A new member of the human coronavirus (HCoV), SARS-CoV-2, which causes coronavirus disease 19 (COVID-19), has become a threat to global health. So, it is most necessary ever that researchers contribute with their knowledge to develop a treatment to block or minimize the damages caused by the virus. This must lead them to be pragmatic and imaginative, without thereby moving away from scientific evidences.

     In this sense, the methoxyindoleamine, melatonin (N-acetyl-5-methoxytryptamine), which is the hormone of the pineal gland of vertebrates and whose first described functions were as a skin lightening agent in amphibians and as a regulator of circadian rhythms in mammals, has been also involved in antioxidant, antiexcitatory actions, vasomotor control, adrenal function, immunomodulation including antiinflammatory properties, and energy metabolism (1-3) Therefore, several clinical trials have shown that melatonin could be efficient in preventing cell damage in metabolic and neurodegenerative diseases, cancer, inflammation, and aging (4- 7). These beneficial effects of melatonin can be explained by its properties as a potent antioxidant, a regulator of programmed cell death (PCD) mechanisms, such as apoptosis and autophagy, and a positive regulator of immune functions.

     Viral infections, which usually come up with oxidative stress, inflammatory injury and triggering of PCD mechanisms, seems to be fully counteracted by the beneficial effects of melatonin. Nevertheless, there are a quite few of articles about the roles of this indolamine on animal viral diseases (8- 11). Surprisingly, a relatively large number of studies about melatonin on virus can also be found in plants (12-14).

     The need to find an effective treatment for COVID-19 patients is urgent. Thus, the aim of this review is to explain the potential mechanisms of melatonin, which could justify its use as a treatment or adjuvant molecule for SARS-CoV-2 infection. Two approaches have been studied for melatonin’s potential on COVID-19. Firstly, the role of melatonin as a regulator of immune system depends on the status of the host, especially frailty elderlies. Secondly, at the cellular level, melatonin modifies the endoplasmic reticulum (ER) stress and autophagy caused by the guest, the SARS-CoV-2.

 

2. THE HOST: IMPORTANCE OF THE RISK GROUPS

     COVID-19 is an infectious disease caused by novel coronavirus SARS-CoV-2.  it originated in the Wuhan, a city of China, in December 2019, then, the disease spread globally resulting in a pandemic, which has caused more than 3 million cases and more than 250 thousand deaths in almost all countries and territories of the world until now.

     Most cases show mild symptoms, such as fever, cough, fatigue, breath shortness, and anosmia. Nevertheless, some cases progress to viral pneumonia, multiorgan failure, and eventually death (15).

     Several studies have reported that the risk for severity and death rate from COVID-19 are higher in older age population.  A recent study in the United States has shown that  no ICU admissions or deaths were reported among persons aged ≤19 years, while 31% of cases, 45% of hospitalizations, 53% of ICU admissions, and 80% of death  associated with COVID-19 were among individuals aged ≥65 years and the highest percentage of severe outcomes was among persons aged ≥85 years (MMWR Morb Mortal Wkly Rep 2020). Another recent report has addressed the role of comorbidities in the outcome of infected patients. The results indicated that 94% of the patients had a chronic health problem, and 88 % of them had two or more health issues. The three most prevalent conditions were hypertension (56.6%), obesity (41.7%), and diabetes (33.8%) (16).

     Therefore, the elderlies, especially those with comorbidities, are the victims of highest risk for serious symptoms and death. Although elderly people are a very heterogeneous group including healthy individuals and those with age-associated pathologies, it is well known that even healthy elderlies show a clear differential response to conditions based on their frailty. The concept of frailty is defined as the condition that a person has a poor capacity to respond to stress. Based on this, a frail elderly has a higher predisposition to suffer both age-related disorders and non-age-dependent diseases (17). Among these latter, it has been reported that frailty is a powerful predictor of the vulnerability to HIV infection (18- 20) and the lack of an adequate response to retroviral therapy (21). Further studies are necessary to use the frailty as a predictor of the outcome of other viral diseases.

     Studies carried out in our laboratory have showed a significant increase in Interleukin 6 (IL-6) levels in frail elderlies when they have faced an adverse event, which is not observed in non-fragile individuals (22-24). Thus, IL-6 increase has been served as a clear marker of dependency in the elderly. Other cohort studies have found that high levels of TNF-α are also markers of frailty (25). 

     The role of these proinflammatory cytokines in the serious cases of COVID-19 has been clearly established. In fact, an overproduction of early response proinflammatory cytokines (tumor necrosis factor (TNF), IL-6, and IL-1β, called as cytokine storm, causes an increased risk of vascular hyperpermeability, multiorgan failure, and eventually death (15). The need to find effective treatments for this SARS-CoV-2 infected frailty patients, especially those with comorbidities, is urgent.  Since the reasons why this group of patients is particularly vulnerable are not clear, to consider all the possibilities is necessary. In this sense, it is worth to note the clearly established role of melatonin as regulator of immune system. Such role has been reported in several animal models where a treatment with this molecule helps to reduce age-associated functional deficits in many organs, including those of the immune system (26-28). Furthermore, several studies have reported that melatonin supplementation acts as anti-inflammatory factor attenuating the function of toll-like receptors (TLRs) and toll-like receptor-associated activator of interferon (TRIF) (29, 30), significantly decreased TNF-α, IL-6 and IL-1β levels in different proinflammatory disorders and diseases (31-34), the overexpressed cytokines in serious cases of COVID-19 (15). Furthermore, melatonin has been highly efficient in critically ill and especially vulnerable people (35, 36), such as is suggested by extensive bibliography about the beneficial effects of melatonin against multiple diseases (37, 38). All these facts support administration of melatonin in patients with COVID-19, specially in elderlies, whose endogenous production of melatonin is drastically reduced to practically zero levels from the age of 70 (28).

 

3. THE GUEST: THE NEED FOR REPLICATION

     Based on sequence comparisons of entire viral genomes, the family Coronaviridae are a group of viruses infecting a wide variety of hosts including avian, swine and humans. They are categorized into four main genera, Alpha-Beta-Gamma- and Deltacoronaviruses (39). HCoVs are identified to be either in the Alpha- or Alpha-Betacoronavirus genera, including Alphacoronaviruses HCoV-229E and HCoV-NL63, and Betacoronaviruses, HCoV-HKU1 SARS-CoV, MERS-CoV and HCoV-OC43.

     SARS-CoV-2, which is a new member of the genus Betacoronavirus, is an enveloped, spherical and about 120 nm in diameter virus. Its genome is a monopartite, linear ssRNA(+) of 27-32kb in size (the largest of all RNA virus genomes), which encodes for two ORFs (ORF1a and the longest ORF1ab containing ORF1a in its 5´ region). ORFs 2-10 encodes four structural and six auxiliary/accessory proteins, which are expressed as subgenomic RNAs. Structural proteins are spike (S), envelope (E) and membrane (M) proteins, which are involved in the formation of the coat, as well as the nucleocapsid (N) protein, which is involved in the packaging of the RNA. Other auxiliary/accessory proteins are responsible for several important functions in genome maintenance and virus replication (40).

     Similar to other CoVs, the life cycle of SARS-CoV-2 begins with the binding of S protein to cellular receptor angiotensin-converting enzyme 2 (ACE2). Following entry into the host cell through an endosomal pathway, the viral RNA is unveiled in the cytoplasm. The translation of ORFs 1a and 1ab produce polypeptides pp1a and pp1ab, which are processed by two virally encoded proteases (3CLpro and PLpro) to yield 16 non-structural proteins that form the RNA replicase-transcriptase complex. This complex drives the production of full-length (−) RNA copies of the genome, which are used as templates for full-length (+) RNA genomes as well as a subset of 10 sub-genomic (sg)RNAs through discontinuous transcription. These sgRNAs encodes both structural and accessory proteins. Viral nucleocapsids are assembled from genomic RNA and N protein in the cytoplasm. S, E and M proteins are translated by ribosomes that are bound to the endoplasmic reticulum (ER), presented on its surface and moved along the secretory pathway into the endoplasmic reticulum–Golgi intermediate compartment (ERGIC). There, viral genomes encapsidated by N protein bud into membranes of the ERGIC containing S, E and M structural proteins, forming mature virions. Virions are then released from the infected cell through exocytosis (41).

 

4. ROUGH ENDOPLASMIC RETICULUM: FACTORY OF VIRAL PROTEINS

     Several stressors, such as aging (42, 43), frailty (24), chemotherapy (44), oxidative injury (45) and bacterial infections (46), can affect ER homeostasis and causes ER-stress. Among these agents, stand up viral infections, because viruses need ER machinery for producing their own proteins which overwhelms ER capacity (47). Such overload condition alters ER-associated proteostasis resulting the accumulated misfolded or unfolded proteins in the ER lumen and inducing a significant increase in oxidative stress (48). This ER-stress activates a signaling cascade, known as unfolded protein response (UPR) (47). This response is initiated by three ER transmembrane sensors, each of which triggers the activation of a different route to achieve a global action of cellular recovery. These sensors are pancreatic ER kinase (PKR)-like ER kinase (PERK), inositol requiring enzyme 1 (IRE1), and activating transcription factor 6 (ATF6) (43). So, PERK starts a route that involves two main substrates: eIF2α and ATF4, which regulate antioxidant response and induces global arrest of protein synthesis among other functions (47). IRE-1 develops several functions as redox homeostasis, antioxidant response and facilitates degradation of unfolded and misfolded proteins by the proteasome via splicing of X-Box binding protein 1 (XBP1) mRNA (49). Finally, ATF6 branch is activated by cleavage of ATF6 obtaining a transcription factor implicated in expansion of the ER to better cope with the proteotoxic stress (50). 

    As mentioned above, translation of some SARS-CoV-2 proteins is necessarily via ER. The overload of this organelle causes oxidative and ER stress triggering UPR. This response can be considered common in viral infections and it is widely documented in several viral families (51-54). The subsequent viral induced UPR must be carefully interpreted , since the literature shows conflicting data depending on the virus, which could indicate that the virus may try to block or facilitate the response to ER stress (47, 55). For example,  in Ebola infection, UPR seems to be considered as an antiviral mechanism (56) while in lytic infections, such as that caused by adenovirus,  the triggering of UPR favors viral replication (57). Although no data about regulation of UPR by SARS-CoV-2 are available up to date, PKR and eIF2α phosphorylation was detected in SARS-CoV-infected cells. Nevertheless, PKR knockdown by using antisense peptide-conjugated phosphorodiamidate morpholino oligomers did not affect SARS-CoV protein replication and virus production. Therefore, a strategy to counteract against the antiviral effects of PKR enabling viral mRNA translation to proceed regardless of whether eIF2α phosphorylation is used. PERK was also found to be activated during SARS-CoV infection, possibly through its S and 3a proteins (58). The IRE-1 branch is also affected by SARS-CoV. Thus, a significant XBP1 splicing and higher rate of apoptosis was observed in infected cells by recombinant SARS-CoV containing deleted E protein (59). In the case of ATF6, no SARS-CoV induction of these branch has been reported.

     Several reports have shown the role of melatonin in ER protection by reduction of both oxidative and ER stress produced by several medical conditions  including obesity, cardiopathy or Alzheimer (60- 63), and the consequent reduction of UPR (64, 65). Furthermore, melatonin has also been implicated in activation of UPR and favoring cellular survival (66), suggesting that melatonin could have a beneficial role in regulation of the UPR, which is triggered by viral induced ER stress. The fact that melatonin reduced both oxidative and ER stress produced in cardiomyocytes of coxsackie virus B3-infected mice (67), as well as activated UPR in hepatocytes of rabbits infected by rabbit hemorrhagic disease virus (68), support the use of melatonin in viral infections such as those caused by SARS-CoV-2.

     It is worth to note that UPR is effective only in conditions of low-to medium ER stress. An excessive high stress overtakes UPR capacity and, then, the survival mechanisms, such as autophagy, has to be triggered. This is the last step before the activation of pro-death signaling (47, 69) and induction of apoptosis.

 

5. ENDOPLASMIC RETICULUM STRESS AND LUNG

     The lungs deserve special attention due to the essential role that they play in the fight against SARS-CoV 2 since they become one of the targets for virus attacking and the hosting organs of this virus during infection. In lungs. Type II alveolar cells, whose transformation into Type I or Type II pneumocytes depends on organ requirements, are essential part of local immune system and they are also a common destination for viral invasion (70, 71). They face a highly increase in oxidative stress after infection and this favors the development of lung fibrosis and a significant reduction of pulmonary capacity (72-75). The pathway from oxidative stress to UPR by increased ER stress has been extensively discussed above and this process has also been addressed in the lungs (76). ER stress is favored not only by such oxidative stress but also by viral infection, which, in turn, increases oxidative stress developing a vicious circle hardly to avoid (77). The reduction of oxidative and ER stress by treatment with antioxidants is a successful therapy for lung damage (78, 79) and viral infections (80). Among these antioxidants, the role of melatonin on pulmonary ER has been well documented (81, 82).

     Angiotensin-converting enzyme 2 (ACE2), which is the main receptor for several human coronaviruses including NL63, SARS-CoV and SARS-CoV2, is expressed in Type II alveolar cells. Thus, these pneumocytes become the target for SARS-CoV-2 infection (83, 84). After pneumocytes have been invaded, virus spread at the pulmonary level followed by those previously mentioned steps as a pattern of viral invasion: oxidative stress (85), ER stress and UPR (77). Therefore, treatment with several antioxidants, including melatonin, to reduce these events on the lungs have been repeatedly proposed (86-88).

 

6. AUTOPHAGY: DUAL ROLE IN VIRAL INFECTIONS

     Unfolded and misfolded proteins from ER are usually degraded by ubiquitin-proteasome system (89, 90). When ER stress is triggered, cytosolic proteasomes are overwhelmed, and proteins are accumulated. The protein aggregates must be degraded via macroautophagy, which is the only survival mechanism that the cell can activate before triggering apoptosis (47, 91).

     Macroautophagy (hereafter referred to as autophagy) is an evolutionarily conserved lysosomal degradation pathway in all eukaryotic cells, which is critical for maintaining cell homeostasis. Autophagy is upregulated in response to extra- or intracellular stress and signals, being present in several physiological and pathological situations from cell development (92) to cancer (93) and pathogen infection (94). An altered autophagy is observed during aging (95), being an impairment hallmark of frailty (24).

     ER stress and autophagy tightly interplay each other usually via UPR system (96), the PERK and IRE1 branches are the main autophagy signaling pathways (97, 98). Furthermore, ER-stress triggers autophagy by downregulation of mammalian target of rapamicine (mTOR), one of main regulators of mammalian metabolism (99).

     The autophagy in viral infection plays dual roles (100). On the one hand, autophagy can reduce virus induced ER stress and thus achieve cell survival by cleaning damaged ER and viruses in autolysosomes, a process called virophagy (101). Example of this mechanism is the autophagic degradation of Sindbis virus by selective degradation of the viral capsid (102). On the other hand, the virus can hijack the process of autophagy and use it to accelerate different stages of the viral life cycle (103-106). Thus, induction of autophagy via UPR by expression of the replicase protein NS4A of flaviviruses, such as dengue virus and Modoc virus, protects cells against death and enhances virus replication (107). Some viruses can increase their replication capacity by inhibition of the fusion of the autophagosome to the lysosome. This leads to an incomplete autophagy creating double-membraned vesicles (DMV), which provide membranous supports for viral RNA replication complexes (108). A similar mechanism has been described for several coronaviruses. Thus, an incomplete autophagy has been induced in HEK293T cells transfected with the membrane-associated papain-like protease PLP2 (PLP2-TM) of MERS-CoV and SARS-CoV. Furthermore, PLP2-TM interacts with the key autophagy regulators, LC3 and Beclin1, and promotes Beclin1 interaction with STING, the key regulator for antiviral IFN signaling (109). A recent study has reported that ubiquitination of BECN1 by S-phase kinase-associated protein 2 (SKP2) resulted in reduction of BECN1 levels and blocks the fusion of autophagosomes and lysosomes in MERS-CoV infected VeroB4 cells (110). The co-localization of viral proteins and LC3, a protein marker for autophagic vacuoles, at DMVs has been reported in SARS-CoV infected Vero cells (111). Nevertheless, replication of a coronavirus, mouse hepatitis virus (MHV), was impaired in autophagy knockout, APG5-/-, embryonic stem cell lines, suggesting that autophagy is required for formation of double membrane-bound MHV replication complexes (112). Although some authors consider that induction of autophagy and following blockage are contradictory (113), others suggest that this can be a strategy of certain viruses (11, 111, 112).

     Autophagy has also been involved in regulation of innate and adaptive immunity systems. Thus, secretion of inflammatory mediators, such as type I interferon (IFN) and cytokines, such as IL-1β and IL-18 to fight microbial infection is promoted by innate immune system, which is stimulated by pattern recognition receptors (PRRs) (114). An aberrant activation can cause excessive inflammation and provoke a severe tissue damage (115). The regulation of the secretion of inflammatory cytokines by autophagy is essential for the correct functioning of the innate immune system (114, 116). On the other hand, processed peptides from pathogen proteins are presented on major histocompatibility complex (MHC) class I and II molecules and recognized by adaptive immune system CD8+ and CD4+ T cells, respectively. The crucial role of autophagy in the presentation of MHC class I and II molecules has been widely reported (116).

     Several studies have reported that melatonin can induce or inhibit autophagy in different biological processes (11). Regarding to viral infections, it has been described that rabbit hemorrhagic disease virus (RHDV) induce autophagy in rabbit hepatocytes, which is inhibited by melatonin administration (117). The dual role of melatonin as autophagy inductor or inhibitor is based on its antioxidant activity. Thus, an incomplete autophagy produces an accumulation of autophagosomes increasing the oxidative stress, whose reduction by melatonin action leads to complete autophagy (118). Thus, the melatonin can be classified as an autophagy regulator.

     Several treatments for COVID-19 are being developed, including those based on the blockage of autophagy flux directly by using rapamycin or mTOR activators (119) or indirectly by using chlorpromazine, which produces an incomplete autophagy and inhibition of clathrin-mediated endocytosis (113). The inhibition of SKP2 protein has also been considered as a target to inhibit viral replication, such as has been suggested by Gassen and coworkers (110). Given that the mechanism of melatonin is the oxidative stress reduction, the use of this molecule to increase the antioxidant capacity as coadjutant to antiviral treatments has ben also suggested (77, 120-122) alone or combined with vitamin D since both  share many control and modulation mechanisms on  immune system (123) or with  other potential drug combinations (124). Melatonin can also act against COVID 19 as mitochondrial metabolic modulator in immune cells (125, 126). Further studies are necessary to find a useful treatment to fight against SARS-CoV-2.

Figure 1-1.png

Fig. 1. Effects of viral infection and melatonin on the process of autophagy.  

     SARS-CoV-2 infection causes ER stress that increases oxidative stress and triggers cellular response pathways, such as PERK and IRE1 branches of the UPR. An excessive oxidative stress and unfolded and misfolded proteins also induce autophagy. An incomplete autophagy by blocking the fusion of autophagosomes and lysosomes is also induced by virus. The incomplete autophagy generates d ouble-membraned vesicles (DMVs) where viral proteins and LC3 are co-localized, are a membranous support for viral RNA replication complexes. Melatonin reduces oxidative stress, reinforces UPR and unlocks the blockage of autophagy caused by virus, allowing autophagosomes to bind to lysosomes and decreases the DMV formation and viral replication capacity.

 

7. CONCLUSIONS

     SARS-CoV-2 infection causes ER stress that increases oxidative stress and triggers cellular response pathways, such as PERK and IRE1 branches of the UPR. An excessive oxidative stress and unfolded and misfolded proteins induce autophagy. The fusion of autophagosomes and lysosomes is blocked by some virus infections leading an incomplete autophagy and forming double-membraned vesicles, which create a membranous support for viral RNA replication complexes. Such as has been discussed in this review, these processes triggered by SARS-CoV-2 may be counteracted by the beneficial effects of melatonin. This indolamine reduces oxidative stress, reinforces UPR and unlocks autophagy blockage induced by virus, allowing autophagosomes to bind to lysosomes and decreasing viral replication capacity (Figure 1). So, the administration of this indolamine should be seriously considered in patients with COVID-19, specially in elderlies and patients with different comorbidities, which are the population with the highest risk for serious symptoms  and death.


ACKNOWLEDGMENTS

      This work was supported by FISS-14-PI13/02741 (Instituto de Salud Carlos III (ISCIII), Spanish Ministry of Science, Innovation and Universities). Both of authors are members of Instituto de Investigación Sanitaria del Principado de Asturias (ISPA) and A C-M is member of Instituto de Neurociencias del Principado de Asturias (INEUROPA). We would also like to acknowledge to K Boga-Coto for improving the English manuscript and N Boga-Coto for his help in figure drawing.

 

AUTHORSHIP

    Both authors have equally contributed to develop of the article.

 

CONFLICT OF INTEREST

     The authors declare that they have no conflict of interest.

 

REFERENCES

 

  1. Pandi-Perumal SR, Srinivasan V, Maestroni GJ, Cardinali DP, Poeggeler B, Hardeland R (2006) Melatonin: Nature's most versatile biological signal? FEBS J. 273 (13): 2813‐2838. DOI:10.1111/j.1742-4658.2006.05322.x.

  2. Hardeland R, Poeggeler B, Srinivasan V, Trakht I, Pandi-Perumal SR, Cardinali DP (2008) Melatonergic drugs in clinical practice. Arzneimittelforschung 58 (1): 1‐10. DOI:10.1055/s-0031-1296459.

  3. Hardeland R, Coto-Montes A, Poeggeler B (2003) Circadian rhythms, oxidative stress, and antioxidative defense mechanisms. Chronobiol. Int. 20 (6): 921‐962. DOI:10.1081/cbi-120025245.

  4. Bonnefont-Rousselot D, Collin F (2010) Melatonin: action as antioxidant and potential applications in human disease and aging. Toxicology 278 (1): 55‐67. DOI:10.1016/j.tox.2010.04.008.

  5. Wu H, Liu J, Yin Y, Zhang D, Xia P, Zhu G. (2019) Therapeutic opportunities in colorectal cancer: focus on melatonin antioncogenic action. Biomed. Res. Int. 9740568: DOI:10.1155/2019/9740568.

  6. Lim Y, Cho H, Kim EK (2016) Brain metabolism as a modulator of autophagy in neurodegeneration. Brain Res. 1649 (Pt B): 158‐165. DOI:10.1016/j.brainres.2016.02.049.

  7. Yang L, Liu X, Song L, Su G, Di A, Bai Ch, Wei Z, Li G (2019) Inhibiting repressive epigenetic modification promotes telomere rejuvenation in somatic cell reprogramming. FASEB J. 33 (12): 13982‐13997. DOI:10.1096/fj.201901486RR.

  8. Junaid A, Tang H, van Reeuwijk A, Abouleila Y, Wuelfroth P, van Duinen V, Stam W, van Zonneveld AJ, Hankemeier T, Mashaghi A (2020) Ebola hemorrhagic shock syndrome-on-a-chip. iScience 23 (1): 100765. DOI:10.1016/j.isci.2019.100765.

  9. Crespo I, Miguel BS, Laliena A, Alvarez M, Culebras JM, Gonzalez-Galleo J, Tuñón MJ (2010) Melatonin prevents the decreased activity of antioxidant enzymes and activates nuclear erythroid 2-related factor 2 signaling in an animal model of fulminant hepatic failure of viral origin. J. Pineal Res. 49 (2): 193–200. DOI:10.1111/j.1600-079X.2010.00787.x.

  10. Crespo I, San-Miguel B, Sánchez DI, Gonzalez-Fernandez B, Alvarez M, Gonzalez-Gallego J, Tuñón MJ (2016) Melatonin inhibits the sphingosine kinase 1/sphingosine-1-phosphate signaling pathway in rabbits with fulminant hepatitis of viral origin. J. Pineal Res. 61 (2): 168–176. DOI:10.1111/jpi.12335.

  11. Boga JA, Caballero B, Potes Y, Perez-Martinez Z, Reiter RJ, Vega-Naredo I, Coto-Montes A (2019) Therapeutic potential of melatonin related to its role as an autophagy regulator: A review. J. Pineal Res. 66 (1): e12534. DOI:10.1111/jpi.12534.

  12. Moustafa-Farag M, Almoneafy A, Mahmoud A, Elkelish A, Arnao MB, Li L, Ai S (2019) Melatonin and its protective role against biotic stress impacts on plants. Biomolecules 10 (1): 54. DOI:10.3390/biom10010054.

  13. Lu R, Liu Z, Shao Y, Sun F, Zhang Y, Cui J, Zhou Y, Shen W, Zhou T (2019) Melatonin is responsible for rice resistance to rice stripe virus infection through a nitric oxide-dependent pathway. Virol. J. 16 (1): 141 DOI:10.1186/s12985-019-1228-3.

  14. Wang M, Zhang S, Ding F (2020) Melatonin mitigates chilling-induced oxidative stress and photosynthesis inhibition in tomato plants. Antioxidants (Basel) 9 (3): 218. DOI:10.3390/antiox9030218.

  15. Jose RJ, Manuel A. COVID-19 (202) Cytokine storm: the interplay between inflammation and coagulation. Lancet Respir. Med. S2213-2600 (20): 30216-2. DOI: 10.1016/S2213-2600(20)30216-2.

  16. Richardson S, Hirsch JS, Narasimhan M, Crawford JM, McGinn T, Davidson KW; and the Northwell COVID-19 Research Consortium, Barnaby DP, Becker LB, Chelico JD, Cohen SL, Cookingham J, Coppa K, Diefenbach MA, Dominello AJ, Duer-Hefele J, Falzon L, Gitlin J, Hajizadeh N, Harvin TG, Hirschwerk DA, Kim EJ, Kozel ZM, Marrast LM, Mogavero JN, Osorio GA, Qiu M, Zanos TP (2020) Presenting characteristics, comorbidities, and outcomes among 5700 patients hospitalized with COVID-19 in the New York City area. JAMA DOI: 10.1001/jama.2020.6775.

  17. Ribeiro AR, Howlett SE, Fernandes A. (2020) Frailty-A promising concept to evaluate disease vulnerability. Mech. Ageing Dev. 187: 111217. DOI:10.1016/j.mad.2020.111217.

  18. Clohisey S, Baillie JK (2019) Host susceptibility to severe influenza A virus infection. Crit. Care 23 (1): 303. DOI:10.1186/s13054-019-2566-7.

  19. Butterfield TR, Landay AL, Anzinger JJ (2020) Dysfunctional immunometabolism in HIV infection: Contributing factors and implications for age-related comorbid diseases. Curr. HIV/AIDS Rep. 17 (2): 125–137. DOI:10.1007/s11904-020-00484-4.

  20. Dessie ZG, Zewotir T, Mwambi H, North D (2020) Modeling viral suppression, viral rebound and state-specific duration of HIV patients with CD4 count adjustment: parametric multistate frailty model approach. Infect. Dis. Ther. DOI:10.1007/s40121-020-00296-4.

  21. Tan H, Wu WZ, Zhao Q, Hong ZC, Feng JY, Zeng H (2020) HIV/AIDS related frailty syndrome in the elderly and related research progress. Zhonghua Liu Xing Bing Xue Za Zhi 41 (1): 127–130. DOI: 10.3760/cma.j.issn.0254-6450.2020.01.023.

  22. de Gonzalo-Calvo D, Neitzert K, Fernández M, Vega-Naredo I, Caballero B, García-Macía M, Suárez FM, Rodríguez-Colunga MJ, Solano JJ, Coto-Montes A (2010) Differential inflammatory responses in aging and disease: TNF-alpha and IL-6 as possible biomarkers. Free Radic. Biol. Med. 49 (5): 733–737. DOI:10.1016/j.freeradbiomed.2010.05.019.

  23. de Gonzalo-Calvo D, de Luxán-Delgado B, Rodríguez-González S, García-Macia M, Suárez FM, Solano JJ, Rodríguez-Colunga MJ, Coto-Montes A (2012) Interleukin 6, soluble tumor necrosis factor receptor I and red blood cell distribution width as biological markers of functional dependence in an elderly population: a translational approach. Cytokine 58 (2): 193–198. DOI:10.1016/j.cyto.2012.01.005.

  24. Potes Y, de Luxán-Delgado B, Rodriguez-González S, Rodrigues Moreira Guimarães M, Solano JJ, Fernandez-Fernandez M, Bermúdez M, Boga JA, Vega-Naredo I, Coto-Montes A (2017) Overweight in elderly people induces impaired autophagy in skeletal muscle. Free Radic. Biol. Med. 110: 31–41. DOI:10.1016/j.freeradbiomed.2017.05.018.

  25. Michaud M, Balardy L, Moulis G, Gaudin C, Peyrot C, Vellas B, Cesari M, Nourhashemi F (2013) Proinflammatory cytokines, aging, and age-related diseases. J. Am. Med. Dir. Assoc. 14 (12): 877-882.

  26. Lardone PJ, Alvarez-García O, Carrillo-Vico A, Vega-Naredo I, Caballero B, Guerrero JM, Coto-Montes A (2006) Inverse correlation between endogenous melatonin levels and oxidative damage in some tissues of SAM P8 mice. J. Pineal Res. 40 (2): 153–157. DOI:10.1111/j.1600-079X.2005.00289.x.

  27. Hardeland R (2018) Melatonin and inflammation-story of a double-edged blade (2018) J. Pineal Res. 65 (4): e12525. DOI:10.1111/jpi.12525.

  28. Hardeland R (2019) Aging, melatonin, and the pro- and anti-inflammatory networks. Int J. Mol. Sci. 20 (5): 1223. DOI:10.3390/ijms20051223.

  29. Chuffa LG, Fioruci-Fontanelli BA, Mendes LO, Seiva FRF, Martínez M, Fávaro WJ, Domeniconi RF, Pinheiro PFF, Dos Santos LD, Martínez FE (2015) Melatonin attenuates the TLR4-mediated inflammatory response through MyD88- and TRIF-dependent signaling pathways in an in vivo model of ovarian cancer. BMC Cancer 15: 34. DOI:10.1186/s12885-015-1032-4.

  30. Chuffa LGA, Reiter RJ, Lupi LA (2017) Melatonin as a promising agent to treat ovarian cancer: molecular mechanisms. Carcinogenesis 38 (10): 945‐952. DOI:10.1093/carcin/bgx054

  31. Zarezadeh M, Khorshidi M, Emami M, Janmohammadi P, Kord-Varkaneh H, Mousavi SM, Mohammed SH, Saedisomeolia A, Alizadeh S (2019) Melatonin supplementation and pro-inflammatory mediators: a systematic review and meta-analysis of clinical trials. Eur. J. Nutr. DOI: 10.1007/s00394-019-02123-0.

  32. Favero G, Franco C, Stacchiotti A, Rodella LF, Rezzani R. Sirtuin1 (2020) Role in the melatonin protective effects against obesity-related heart injury. Front. Physiol. 11: 103. DOI: 10.3389/fphys.2020.00103. eCollection 2020.

  33. Zarezadeh M, Khorshidi M, Emami M, Janmohammadi P, Kord-Varkaneh H, Mouvasi SM, Mohammed SH, Saedisomeolia A, Alizadeh S (2019) Melatonin supplementation and pro-inflammatory mediators: a systematic review and meta-analysis of clinical trials Eur. J. Nutr. 10.1007/s00394-019-02123-0. DOI:10.1007/s00394-019-02123-0.

  34. Zhang Y, He F, Chen Z, Su Q, Yan M, Zhang Q, Tan J, Qian L, Han Y (2019) Melatonin modulates IL-1β-induced extracellular matrix remodeling in human nucleus pulposus cells and attenuates rat intervertebral disc degeneration and inflammation. Aging (Albany NY) 11 (22): 1049910512. DOI:10.18632/aging.102472.

  35. Bourne RS, Mills GH (2006) Melatonin: possible implications for the postoperative and critically ill patient. Intensive Care Med. 32 (3): 371–379.

  36. Rinaldi S, Landucci F, De Gaudio AR (2009) Antioxidant therapy in critically septic patients. Curr. Drug Targets 10 (9): 872‐880. DOI:10.2174/138945009789108774.

  37. Sato K, Meng F, Francis H, Wu N, Chen L, Kennedy L, Zhou T, Franchitto A, Onori P, Gaudio E, Glaser S, Alpini G (2020) Melatonin and circadian rhythms in liver diseases: Functional roles and potential therapies. J. Pineal Res. 68 (3): e12639. DOI:10.1111/jpi.12639

  38. Imenshahidi M, Karimi G, Hosseinzadeh H. (2020) Effects of melatonin on cardiovascular risk factors and metabolic syndrome: a comprehensive review. Naunyn Schmiedebergs Arch. Pharmacol. 393 (4): 521–536. DOI:10.1007/s00210-020-01822-4.

  39. Gorbalenya AE, Snijder EJ, Spaan WJM (2004) Severe acute respiratory syndrome coronavirus phylogeny: toward consensus.  J. Virol. 78: 7863–7866. DOI: 10.1128/JVI.78.15.7863-7866.2004.

  40. Chan JF, Kok KH, Zhu Z, Chu H, To KK, Yuan S, Yuen KY (2020) Genomic characterization of the novel human-pathogenic coronavirus isolated from a patient with atypical pneumonia after visiting Wuhan. Emerg. Microbes Infect. 9 (1): 221-236.

  41. Fehr A.R., Perlman S (2015) Coronaviruses: An overview of their replication and pathogenesis. In: Maier H., Bickerton E., Britton P. (eds) Coronaviruses. Methods in Molecular Biology, vol 1282. Humana Press, New York, NY.

  42. Estébanez B, de Paz JA, Cuevas MJ, González-Gallego (2018) Endoplasmic reticulum unfolded protein response, aging and exercise: an update. Front. Physiol. 9:1744. DOI:10.3389/fphys.2018.01744.

  43. Salminen A, Kaarniranta K, Kauppinen A (2020) ER stress activates immunosuppressive network: implications for aging and Alzheimer's disease J. Mol. Med. (Berl.) 10.1007/s00109-020-01904-z. DOI:10.1007/s00109-020-01904-z.

  44. Zhao A, Zhang Z, Zhou Y, Li X, Li X, Li B, Ma B, Zhang Q (2020) β-Elemonic acid inhibits the growth of human Osteosarcoma through endoplasmic reticulum (ER) stress-mediated PERK/eIF2α/ATF4/CHOP activation and Wnt/β-catenin signal suppression Phytomedicine. 69: 153183. DOI:10.1016/j.phymed.2020.153183.

  45. Uddin MJ, Dorotea D, Pak ES, Ha H (2020) Fyn kinase: A potential therapeutic target in acute kidney injury. Biomol. Ther. (Seoul) 28 (3): 213‐221. DOI:10.4062/biomolther.2019.214.

  46. Kuss-Duerkop SK, Keestra-Gounder AM (2020) NOD1 and NOD2 activation by diverse stimuli: a possible role for sensing pathogen-induced ER stress. Infect. Immun. IAI.00898-19. DOI:10.1128/IAI.00898-19.

  47. Sprooten J, Garg AD (2020) Type I interferons and endoplasmic reticulum stress in health and disease. Int. Rev. Cell. Mol. Biol. 350: 63–118. DOI: 10.1016/bs.ircmb.2019.10.004.

  48. Gorbatyuk MS, Starr CR, Gorbatyuk OS (2020) Endoplasmic reticulum stress: New insights into the pathogenesis and treatment of retinal degenerative diseases Prog. Retin. Eye Res. 100860: DOI:10.1016/j.preteyeres.2020.100860.

  49. Almanza A, Carlesso, A, Chintha C, Creedican S, Doultsinos D, Leuzzi B, Luıs A, McCarthy N, Montibeller L, More S, Papaioannou A, Püschel F, Sassano ML, Skoko J, Agostinis P, de Belleroche J, Eriksson LA, Fulda S, Gorman AM, Healy S, Kozlov A, Muñoz-Pinedo C, Rehm M, Chevet E, Samali A (2019) Endoplasmic reticulum stress signalling—from basic mechanisms to clinical applications. FEBS J. 286: 241–278.

  50. Oka OB van Lith M, Rudolf J, Tungkum W, Pringle MA, Bulleid NJ (2019) ERp18 regulates activation of ATF6α during unfolded protein response. EMBO J. 38: e100990. DOI: 10.15252/embj.2018100990

  51. Kolpikova EP, Tronco AR, Hartigh ABD, Jackson KJ, Iwawaki T, Fink SL (2020) IRE1α promotes Zika virus infection via XBP1. Viruses 12 (3): 278. DOI:10.3390/v12030278

  52. Lee MK, Hyeon S, Ahn JH (2020) Loss of VCP/p97 induced by the human cytomegalovirus transmembrane protein pUL50 and its regulation by a small isoform of pUL50.  J. Virol. JVI.00110-20. DOI:10.1128/JVI.00110-20.

  53. Lewy TG, Offerdahl DK, Grabowski JM, Kellman E, Mlera L, Chiramel A, Bloom ME (2020) PERK-mediated unfolded protein response signaling restricts replication of the tick-borne flavivirus langat virus. Viruses 12 (3): 328. DOI:10.3390/v12030328.

  54. Ríos-Ocampo WA, Navas MC, Buist-Homan M, Faber KN, Daemen T, Moshage H (2020) Hepatitis C virus proteins core and NS5A are highly sensitive to oxidative stress-induced degradation after eIF2α/ATF4 pathway activation. Viruses. 12 (4): E425. DOI:10.3390/v12040425.

  55. Shinjo, S, Mizotani, Y, Tashiro, E, Imoto, M (2013) Comparative analysis of the expression patterns of UPR-target genes caused by UPR-inducing compounds. Biosci. Biotechnol. Biochem. 77: 729–735. DOI: 10.1271/bbb.120812.

  56. Elfiky AA (2020) Ebola virus glycoprotein GP1-host cell-surface HSPA5 binding site prediction. Cell Stress Chaperones. 10.1007/s12192-020-01106-z. DOI:10.1007/s12192-020-01106-z

  57. Prasad V, Suomalainen M, Jasiqi Y, Hemmi S, Hearing P, Burgert H-G, Greber UF (2020) The UPR sensor IRE1α and the adenovirus E3-19K glycoprotein sustain persistent and lytic infections. Nat. Commun. 11 (1): 1997. DOI:10.1038/s41467-020-15844-2.

  58. Lim YX, Ng YL, Tam JP, Liu DX (2016) Human coronaviruses: a review of virus-host interactions. Diseases 4 (3): 26. DOI: 10.3390/diseases4030026.

  59. De Diego ML, Nieto-Torres J.L., Jimenez-Guardeño J.M., Regla-Nava J.A., Álvarez E., Oliveros J.C., Zhao J., Fett C., Perlman S., Enjuanes L (2011) Severe acute respiratory syndrome coronavirus envelope protein regulates cell stress response and apoptosis. PloS Pathog7 (10): e1002315. DOI: 10.1371/journal.ppat.1002315. Epub 2011 Oct 20.

  60. Lin CH, Nicol CJB, Chen Y-Ch, Chen S-J, Yen Ch-H, Huang R-N, Chiang M-Ch (2018) Rosiglitazone rescues human neural stem cells from amyloid-beta induced ER stress via PPARγ dependent signaling. Exp. Cell Res. 370 (2): 312‐321. DOI:10.1016/j.yexcr.2018.06.033.

  61. de Luxán-Delgado B, Potes Y, Rubio-González A, Caballero B, Solano JJ, Fernández-Fernández M, Bermúdez M, Rodrigues Moreira Guimarães M, Vega-Naredo I, Boga JA, Coto-Montes A (2016) Melatonin reduces endoplasmic reticulum stress and autophagy in liver of leptin-deficient mice. J. Pineal Res. 61 (1): 108‐123. DOI:10.1111/jpi.12333.

  62. Zhang J, Wang L, Xie W, Hu S, Zhou H, Zhu P, Zhu H (2020) Melatonin attenuates ER stress and mitochondrial damage in septic cardiomyopathy: A new mechanism involving BAP31 upregulation and MAPK-ERK pathway. J. Cell Physiol. 235 (3): 2847‐2856. DOI:10.1002/jcp.29190.

  63. Shi C, Zeng J, Li Z, Chen Q, Hang W, Xia L, Wu Y, Chen J, Shi A (2018) Melatonin mitigates kainic acid-induced neuronal tau hyperphosphorylation and memory deficits through alleviating ER stress. Front. Mol. Neurosci. 11: 5. DOI:10.3389/fnmol.2018. 00005.

  64. Bermejo-Millo JC, Guimarães MRM, de Luxán-Delgado B, Potes Y, Perez-Martinez Z, Diaz-Luis A, Caballero B, Solano JJ, Vega-Naredo I, Coto-Montes A (2018) High-Fructose Consumption Impairs the Redox System and Protein Quality Control in the Brain of Syrian Hamsters: Therapeutic Effects of Melatonin. Mol. Neurobiol. 55 (10): 7973‐7986. DOI:10.1007/s12035-018-0967-2.

  65. Rubio-González A, Bermejo-Millo JC, de Luxán-Delgado B, Potes Y, Perez-Martinez Z, Boga JA, Vega-Naredo I, Caballero B, Solano JJ, Coto-Montes A (2018) Melatonin prevents the harmful effects of obesity on the brain, including at the behavioral level. Mol. Neurobiol. 55 (7): 5830‐5846. DOI:10.1007/s12035-017-0796-8.

  66. Kleber A, Kubulus D, Rössler D, Wolf B, Volk T, Speer T, Fink T (2014) Melatonin modifies cellular stress in the liver of septic mice by reducing reactive oxygen species and increasing the unfolded protein response. Exp. Mol. Pathol. 97 (3): 565‐571. DOI:10.1016/j.yexmp.2014.10.009.

  67. Ouyang H, Zhong J, Lu J, Zhong Y, Hu Y, Tan Y (2019) Inhibitory effect of melatonin on Mst1 ameliorates myocarditis through attenuating ER stress and mitochondrial dysfunction. J. Mol. Histol. 50 (5): 405–415. DOI:10.1007/s10735-019-09836-w.

  68. Tuñón MJ, San-Miguel B, Crespo I, Laliena A, Vallejo D, Álvarez M, Prieto J, González-Gallego J (2013) Melatonin treatment reduces endoplasmic reticulum stress and modulates the unfolded protein response in rabbits with lethal fulminant hepatitis of viral origin. J. Pineal Res. 55 (3): 221–222DOI:10.1111/jpi.12063.

  69. Estornes, Y, Aguileta, MA, Dubuisson, C, De Keyser, J, Goossens, V, Kersse, K, Samali, A, Vandenabeele, P, Bertrand, MJM (2014) RIPK1 promotes death receptor-independent caspase-8-mediated apoptosis under unresolved ER stress conditions. Cell Death Dis. 6 (6): e1798. DOI: 10.1038/cddis.2015.175.

  70. Fiege JK, Stone IA, Dumm RE, Waring BM, Fife BT, Agudo j, Brown BD, Heaton NS, Langlois RA (2019) Long-term surviving influenza infected cells evade CD8+ T cell mediated clearance. PLoS Pathog. 15 (9): e1008077.DOI:10.1371/journal.ppat.1008077.

  71. Takezaki A, Tsukumo SI, Setoguchi Y, Ledford JG, Goto H, Hosomichi K, Uehara H, Nishioka Y, Yasutomo K (2019) A homozygous SFTPA1 mutation drives necroptosis of type II alveolar epithelial cells in patients with idiopathic pulmonary fibrosis. J. Exp. Med. 216 (12): 2724‐2735. DOI:10.1084/jem.20182351.

  72. Hsu YH, Cho LC, Wang LS, Chen LK, Lee JJ, Yang HH. (2006) Acute respiratory distress syndrome associated with rabies: a case report. Kaohsiung J. Med. Sci. 22 (2): 94‐98. DOI:10.1016/S1607-551X(09)70227-X.

  73. Casola A, Burger N, Liu T, Jamaluddin M, Brasier AR, Garofalo RP. (2001) Oxidant tone regulates RANTES gene expression in airway epithelial cells infected with respiratory syncytial virus: Role in viral-induced interferon regulatory factor activation. J. Biol. Chem. 276 (23): 19715‐19722. DOI:10.1074/jbc.M101526200.

  74. Richter K and Kietzmann T (2016) Reactive oxygen species and fibrosis: further evidence of a significant liaison. Cell Tissue Res. 365: 591–605.

  75. Khomich OA, Kochetkov SN, Bartosch B, and Ivanov AV (2018) Redox biology of respiratory viral infections. Viruses 10 (8): 392.

  76. Kropski JA, Blackwell TS (2018) Endoplasmic reticulum stress in the pathogenesis of fibrotic disease. J. Clin. Invest. 128 (1): 64‐73. DOI:10.1172/JCI93560.

  77. Abouhashem AS, Singh K, Azzazy HME, Sen CK (2020) Is low alveolar type II cell SOD3 in the lungs of elderly linked to the observed severity of COVID-19? Antioxid. Redox Signal. 10.1089/ars.2020.8111. DOI:10.1089/ars.2020.8111

  78. Borok Z, Horie M, Flodby P, Wang H, Liu Y, Ganesh S, Firth AL, Minoo P, Li C, Beers MF, Lee AS, Zhou B (2020) Grp78 loss in epithelial progenitors reveals an age-linked role for endoplasmic reticulum stress in pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 201 (2): 198‐211. DOI:10.1164/rccm.201902-0451OC.

  79. Tanaka K, Ishihara T, Azuma A, Kudoh S, Ebina M, Nukiwa T, Sugiyama Y, Tasaka Y, Namba T, Ishihara T, Sato K, Mizushima Y, and Mizushima T (2010) Therapeutic effect of lecithinized superoxide dismutase on bleomycin-induced pulmonary fibrosis. Am. J. Physiol. Lung Cell Mol. Physiol. 298: L348–L360.

  80. Mileva M, Dimitrova A, Krastev D, Alexandrova A, Tsvetanova E, Georgieva A, Galabov A (2020) Oseltamivir and S-adenosyl-l-methionine combination as effective therapeutic strategy for suppression of oxidative damage in lung caused by influenza virus infection in mice. Drug Res. (Stuttg). 10.1055/a-1147-8824. DOI:10.1055/a-1147-8824.

  81. He B, Zhang W, Qiao J, Peng Z, Chai X (2019) Melatonin protects against COPD by attenuating apoptosis and endoplasmic reticulum stress via upregulating SIRT1 expression in rats. Can. J. Physiol. Pharmacol. 97 (5): 386‐391. DOI:10.1139/cjpp-2018-0529.

  82. He B, Chen Q, Zhou D, Wang L, Liu Z (2019) Role of reciprocal interaction between autophagy and endoplasmic reticulum stress in apoptosis of human bronchial epithelial cells induced by cigarette smoke extract. IUBMB Life 71 (1): 66‐80. DOI:10.1002/iub.1937.

  83. Magrone T, Magrone M, Jirillo E (2020) Focus on receptors for coronaviruses with special reference to angiotensin-converting enzyme 2 as a potential drug target - a perspective. Endocr. Metab. Immune Disord. Drug Targets. 10.2174/1871530320666200427112902. DOI:10.2174/1871530320666200427112902.

  84. Zou X, Chen K, Zou J, Han P, Hao J, Han Z (2020) Single-cell RNA-seq data analysis on the receptor ACE2 expression reveals the potential risk of different human organs vulnerable to 2019-nCoV infection. Front Med. 14 (2): 185‐192. DOI:10.1007/s11684-020-0754-0.

  85. Farnoosh G, Ghanei M, Khorramdelazad H, Alishiri G, Farahani AJ, Shahriary A, Zijoud SRH (2020) Are Iranian Sulfur mustard-exposed survivors more vulnerable to SARS-CoV-2: some similarity in their pathogenesis. Disaster Med. Public Health Prep. 1‐12. DOI:10.1017/dmp.2020.156.

  86. Dalamaga M, Karampela I, Mantzoros CS (2020) Commentary: Could iron chelators prove to be useful as an adjunct to COVID-19 treatment regimens? Metabolism. 108: 154260. DOI:10.1016/j.metabol.2020.154260.

  87. Messina G, Polito R, Monda V, Cipolloni L, Nunno ND, Mizio GD, Murabito P, Carotenuto M, Messina A, Pisanelli D, Valenzano A, Cibelli G, Scarinci A, Monda M, Sessa F (2020) Functional role of dietary intervention to improve the outcome of COVID-19: A hypothesis of work. Int. J. Mol. Sci. 21 (9): E3104. DOI:10.3390/ijms21093104.

  88. Shneider A, Kudriavtsev A, Vakhrusheva A (2020) Can melatonin reduce the severity of COVID-19 pandemic? Int. Rev. Immunol. 1‐10. DOI:10.1080/08830185.2020.1756284.

  89. Scriven P, Brown NJ, Pockley AG, Wyld L (2007) The unfolded protein response and cancer: a brighter future unfolding? J. Mol. Med. 85: 331-341. DOI: 10.1007/s00109-006-0150-5.

  90. Lin JH, Walter P, Yen TS (2008) Endoplasmic reticulum stress in disease pathogenesis. Annu. Rev. Pathol. 3: 399- 425.

  91. Yorimitsu T, Nair U, Yang Z, Klionsky DJ (2006) Endoplasmic reticulum stress triggers autophagy. J. Biol. Chem. 281 (40): 30299‐30304. DOI:10.1074/jbc.M607007200.

  92. Allen EA, Baehrecke EH (2020) Autophagy in animal development. Cell Death Differ. 27 (3): 903‐918. DOI:10.1038/s41418-020-0497-0.

  93. Horton R, Wileman T, Rushworth SA (2020) Autophagy driven extracellular vesicles in the leukaemic microenvironment Curr. Cancer Drug Targets 10.2174/1568009620666200428111051. DOI:10.2174/1568009620666200428111051.

  94. He C, Klionsky DJ (2009) Regulation mechanisms and signaling pathways of autophagy. Annu. Rev. Genet. 43: 67‐93. DOI:10.1146/annurev-genet-102808-114910.

  95. Kulkarni AS, Gubbi S, Barzilai N (2020) Benefits of metformin in attenuating the hallmarks of aging. Cell Metab. S1550-4131(20)30183-2. DOI:10.1016/j.cmet.2020.04.001.

  96. Yorimitsu T, Klionsky DJ (2007) Endoplasmic reticulum stress: a new pathway to induce autophagy. Autophagy 3 (2): 160‐162. DOI:10.4161/auto.3653.

  97. Wang D, Zhang P, Xu X, Wang J, Wang D, Peng P, Zheng C, Meng Q-J, Yang L, Luo Z (2019) Knockdown of cytokeratin 8 overcomes chemoresistance of chordoma cells by aggravating endoplasmic reticulum stress through PERK/eIF2α arm of unfolded protein response and blocking autophagy. Cell Death Dis. 10 (12): 887. DOI: 10.1038/s41419-019-2125-9.

  98. Maeyashiki C, Melhem H, Hering L, Baebler K, Cosin-Roger J, Schefer F, Weder B, Hausmann M, Scharl M, Rogler G, de Vallière C, Ruiz PA (2020) Activation of pH-sensing receptor OGR1 (GPR68) induces ER stress via the IRE1α/JNK pathway in an intestinal epithelial cell model. Sci. Rep. 10 (1): 1438. DOI:10.1038/s41598-020-57657-9.

  99. Qin L, Wang Z, Tao L, Wang Y (2010) ER stress negatively regulates AKT/TSC/mTOR pathway to enhance autophagy. Autophagy 6 (2): 239‐247. DOI:10.4161/auto.6.2.11062.

  100. Dong X and Levine B (2013) Autophagy and viruses: adversaries or allies? J. Innate Immun. 5: 480–493. DOI: 10.1159/000346388.

  101. Paulus GL, Xavier RJ (2015) Autophagy and checkpoints for intracellular pathogen defense. Curr. Opin. Gastroenterol. 31: 14‐23. DOI: 10.1097/MOG.0000000000000134.

  102. Orvedahl A, MacPherson S, Sumpter R Jr, Talloczy Z, Zou Z, Levine B (2010) Autophagy protects against Sindbis virus infection of the central nervous system. Cell Host. Microbe 7: 115‐127. DOI: 10.1016/j.chom.2010.01.007.

  103. Huang SC, Chang CL, Wang PS, Tsai Y, Liu HS (2009) Enterovirus 71‐induced autophagy detected in vitro and in vivo promotes viral replication. J. Med. Virol. 81: 1241‐1252. DOI: 10.1016/j.chom.2010.01.007.

  104. O'Donnell V, Pacheco JM, LaRocco M, Burrage T, Jackson W, Rodriguez LL, Borca MV, Baxt B (2011) Foot‐and‐mouth disease virus utilizes an autophagic pathway during viral replication. Virology 410: 142‐150. DOI: 10.1016/j.virol.2010.10.042. 

  105. Granato M, Santarelli R, Farina A, Gonnella R, Lotti LV, Faggioni A, Cirone M (2014) Epstein‐barr virus blocks the autophagic flux and appropriates the autophagic machinery to enhance viral replication. J. Virol. 88: 12715‐12726. DOI: 10.1128/JVI.02199-14.

  106. Casais R, Molleda LG, Machín A, del Barrio G, García Manso A, Dalton KP, Coto A, Martin Alonso JM, Prieto M, Parra F (2008) Structural and functional analysis of virus factories purified from rabbit vesivirus-infected Vero cells. Virus Res. 137 (1): 112‐121. DOI:10.1016/j.virusres.2008.06.009.

  107. McLean JE, Wudzinska A, Datan E, Quaglino D, Zakeri Z (2011) Flavivirus NS4A‐induced autophagy protects cells against death and enhances virus replication. J. Biol. Chem. 286: 22147‐22159. DOI: 10.1074/jbc.M110.192500.

  108. Jackson WT, Giddings TH Jr, Taylor MP, Mulinyawe S, Rabinovitch M, Kopito RR, Kirkegaard K (2005) Subversion of cellular autophagosomal machinery by RNA viruses. PLoS Biol. 3: e156. DOI: 10.1371/journal.pbio.0030156.

  109. Chen X, Wang K, Xing Y, Tu J, Yang X, Zhao Q, Li K, Chen Z (2014) Coronavirus membrane-associated papain-like proteases induce autophagy through interacting with Beclin1 to negatively regulate antiviral innate immunity. Protein Cell. 5: 912-27. DOI: 10.1007/s13238-014-0104-6.

  110. Gassen NC, Niemeyer D, Muth D, Corman VM, Martinelli S, Gassen A, Hafner K, Papies J, Mösbauer K, Zellner A, Zannas AS, Herrmann A, Holsboer F, Brack-Werner R, Boshart M, Müller-Myhsok B, Drosten C, Müller MA, Rein T (2019) SKP2 attenuates autophagy through Beclin1-ubiquitination and its inhibition reduces MERS-Coronavirus infection. Nat. Commun. 10 (1): 5770. DOI:10.1038/s41467-019-13659-4.

  111. Prentice E, McAuliffe J, Lu X, Subbarao K, Denison MR (2004a) Identification and characterization of severe acute respiratory syndrome coronavirus replicase proteins. J. Virol. 78: 9977-86.

  112. Prentice E, Jerome WG, Yoshimori T, Mizushima N, Denison MR (2004b) Coronavirus replication complex formation utilizes components of cellular autophagy. J. Biol. Chem. 279: 10136-10141. DOI: 10.1074/jbc.M306124200.

  113. Yang N, Shen HM (2020) Targeting the endocytic pathway and autophagy process as a novel therapeutic strategy in COVID-19. Int. J. Biol. Sci. 16 (10): 1724‐1731. DOI:10.7150/ijbs.45498.

  114. Shibutani ST, Saitoh T, Nowag H, Munz C, Yoshimori T (2015) Autophagy and autophagy-related proteins in the immune system. Nat. Immunol. 16: 1014–1024. DOI: 10.1038/ni.3273.

  115. Rathinam VA, Vanaja SK, Fitzgerald KA (2012) Regulation of inflammasome signaling. Nat. Immunol. 13: 333–342. DOI: 10.1038/ni.2237.

  116. Jang YJ, Kim JH, Byun S (2019) Modulation of autophagy for controlling immunity. Cells 8 (2): 138. DOI:10.3390/cells8020138.

  117. San-Miguel B, Crespo I, Vallejo D, Álvarez M, Prieto J, Gonzalez-Gallego J, Tuñón MJ (2014) Melatonin modulates the autophagic response in acute liver failure induced by the rabbit hemorrhagic disease virus. J. Pineal Res. 56 (3): 313–321. DOI:10.1111/jpi.12124.

  118. Vega-Naredo I, Caballero B, Sierra V, García-Macia M, de Gonzalo- Calvo D, Oliveira PJ, Rodríguez-Colunga MJ, Coto-Montes A (2012) Melatonin modulates autophagy through a redox-mediated action in female Syrian hamster Harderian gland controlling cell types and gland activity. J. Pineal Res. 52 (1): 80‐92. DOI:10.1111/j.1600-079X.2011.00922.x.

  119. Maiese K (2020) The mechanistic target of rapamycin (mTOR): Novel considerations as an antiviral treatment and possibilities for COVID-19 Curr. Neurovasc. Res. 10.2174/1567202617666200425205122. DOI:10.2174/1567202617666200425205122.

  120. Tan, DX, Hardeland R (2020) Potential utility of melatonin in deadly infectious diseases related to the overreaction of innate immune response and destructive inflammation: focus on COVID-19. Melatonin Res. 3 (1): 120-143. DOI:10.32794/mr11250052.

  121. Zhang R, Wang X, Ni L, Di X, Ma B, Niu S, Liu C, Reiter RJ (2020) COVID-19: Melatonin as a potential adjuvant treatment. Life Sci. 250: 117583. DOI:10.1016/j.lfs.2020.117583.

  122. Shneider A, Kudriavtsev A, Vakhrusheva A (2020) Can melatonin reduce the severity of COVID-19 pandemic? Int. Rev. Immunol. 1‐10. DOI:10.1080/08830185.2020.1756284

  123. Martín Giménez VM, Inserra F, Tajer CD, Mariani J, Ferder L, Reiter RJ, Manucha W (2020) Lungs as target of COVID-19 infection: Protective common molecular mechanisms of vitamin D and melatonin as a new potential synergistic treatment. Life Sci. 254: 117808. DOI:10.1016/j.lfs.2020.117808.

  124. Zhou Y, Hou Y, Shen J, Huang Y, Martin W, Cheng F (2020) Network-based drug repurposing for novel coronavirus 2019-nCoV/SARS-CoV-2. Cell Discov. 6: 14. DOI:10.1038/s41421-020-0153-3.

  125. Anderson G, Reiter RJ (2020) Melatonin: Roles in influenza, Covid-19, and other viral infections. Rev. Med. Virol. 30 (3): e2109. DOI:10.1002/rmv.2109.

  126. Reiter RJ, Sharma R, Ma Q, Dominquez-Rodriguez A, Marik PE, Abreu-Gonzalez P (2020) Melatonin inhibits COVID-19-induced cytokine storm by reversing aerobic glycolysis in immune cells: a mechanistic analysis. Med. Drug Discov. 6: 100044. DOI:10.1016/j.medidd.2020.100044.

CCBY-1.png


This work is licensed under a Creative Commons Attribution 4.0 International License