Role of Selenium in Viral Infections: Comparison
Please note this is a comparison between Version 2 by Bruce Ren and Version 1 by Sabrina Sales Martinez.

Viral infections have afflicted human health and despite great advancements in scientific knowledge and technologies, continue to affect our society today. The current coronavirus (COVID-19) pandemic has put a spotlight on the need to review the evidence on the impact of nutritional strategies to maintain a healthy immune system, particularly in instances where there are limited therapeutic treatments. Selenium, an essential trace element in humans, has a long history of lowering the occurrence and severity of viral infections. Much of the benefits derived from selenium are due to its incorporation into selenocysteine, an important component of proteins known as selenoproteins. Viral infections are associated with an increase in reactive oxygen species and may result in oxidative stress. Studies suggest that selenium deficiency alters immune response and viral infection by increasing oxidative stress and the rate of mutations in the viral genome, leading to an increase in pathogenicity and damage to the host. 

  • selenium
  • selenoproteins
  • virus
  • viral
  • infection
  • reactive oxygen species
  • antioxidant
  • HIV
  • HCV
  • HBV
  • coxsackie virus
  • influenza
  • glutathione peroxidase
  • thioredoxin reductase

1. Introduction

Viral infections have afflicted human health despite great advancements in scientific knowledge and technologies [1,2,3][1][2][3]. Most recently, the novel severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has infected over 200 million individuals during 2019–August 2021 and has led to over 4.4 million deaths globally [4]. Selenium (Se), an essential trace element in humans, has a long history of lowering the occurrence and severity of viral infections [5,6,7,8,9][5][6][7][8][9]. Se deficiency impacts immune function [10], viral expression [8], selenoprotein expression [11], and alters antioxidant response [12], allowing for greater susceptibility to severe viral and bacterial infections [13]. Supplementing the diet with Se has demonstrated positive effects on enhancing immunity against viral attacks [5]. Much of the benefits derived from Se are due to its incorporation into selenocysteine, an important component of the antioxidant defense systems, including the regulation of glutathione peroxidase (GPXs) and thioredoxin reductase (TXNRD) activities [14]. Low levels of Se can lead to more severe forms of viral infections and adequate selenium levels may provide a protective effect toward the host response by affecting both immune response and oxidative stress [13,15][13][15]. Severe pathology in Se deficiency is evidenced by more frequent and graver symptoms, higher viral loads, declining levels of antioxidant enzymes such as GPX, and mutations to the viral genome. Studies conducted by Beck et al. described in this review, demonstrate that Se-deficiency is capable of increasing the virulence of a benign coxsackie virus through viral mutations and these mutations have led to a reduction in GPX activity, therefore, resulting in oxidative stress [13,15][13][15].

2. Viral Infections and Selenium

2.1. Coxsackie Virus

Several decades of research have provided sufficient evidence to demonstrate a relationship between Se deficiency and Keshan disease, a grave cardiomyopathy. This cardiomyopathy is believed to be caused by infection with Coxsackie B virus, a nonenveloped single-stranded RNA virus pertaining to the Picornaviridae family, and exclusively found within China [10,45][10][16]. It was later discovered in the 1970s and 1980s that much of the Se levels in the soil, water, food, and human circulating fluids in areas affected by Keshan disease were deficient compared to other neighboring Chinese providences [46][17]. Sodium selenite was provided to the population and a prospective study showed that it prevented Keshan disease. Keshan disease was eradicated from endemic areas after the government enacted a Se supplementation policy, therefore, demonstrating that Keshan disease occurred due to two factors, infection with Coxsackie B virus and Se deficiency [47,48,49][18][19][20].
Animal studies conducted by Beck et al. confirmed the relationship with Se in mice by infection with a non-cardio-virulent strain of Coxsackie B virus (CVB3/0) and a myocarditic strain (CVB3/20). Heart damage was only observed in the mice fed a Se-deficient diet compared to mice fed a Se-sufficient diet for 4 weeks, and the typical human pathology was also observed [28,50][21][22]. These studies illustrated that Se deficiency caused a virus that was non-virulent to contribute towards the development of myocarditis in the host, and also increased its pathogenicity as the cardiovirulent strain under Se deficiency produced greater symptoms [28,51][21][23].
Additional observations by Beck and colleagues showed higher viral loads in the Se-deficient mice infected with both CVB3/0 and CVB3/20. The Se-deficient mice were found to have reduced T-cell expansion and diminished mRNA levels of cytokines compared to Se-adequate mice [15]. Subsequent studies led to the finding that Se deficiency was responsible for a change in the genotype of the benign coxsackie virus CVB3/0 that caused it to become virulent. Specifically, six nucleotides were modified that mimicked other virulent strains of CVB3 viruses. Due to these mutations, the virus now had the possibility to become pathogenic even in a Se-adequate host [45][16]. It was then hypothesized that a reduction in GPX activity was responsible for the viral mutations. Therefore, subsequent studies were conducted to demonstrate the protective effect of GPX1 in developing heart damage when infected with a benign strain of Coxsackie B virus (CVB3/0) [52][24]. Mice with a disrupted gpx1 gene infected with CVB3/0 compared with wild type mice with an intact gpx1 gene experienced myocarditis, and sequencing of the viruses from the mice with disrupted gpx1 gene showed seven nucleotide changes in the Coxsackie virus. Interestingly, six of the seven nucleotide changes in the genome of the virus from the mice with disrupted gpx1 genes matched the changes found in the Se-deficient mice previously [52][24]. These classic experiments exhibit how nutritional status as it pertains to Se, and its ability to protect antioxidant systems and immunity may impact the potential evolution of viruses to become more virulent.

2.2. Influenza

Influenza viruses, known to cause the flu, are enveloped, single-stranded RNA viruses within the Orthomyxoviridae family. Selenium deficiency has been associated with poor selenoprotein expression [11] and altered antioxidant response in viral influenza A infection [12]. The elegant in vitro [13] and animal experiments conducted by Dr. Beck et al. [28,45,46,50,52,53][21][16][17][22][24][25] were the first to demonstrate the detrimental effects of Se deficiency in influenza A virulence, which occurred due to changes in the viral genome [54][26]. Se deficiency in mice infected with a highly virulent Influenza A strain (Influenza A/PR/8/34), however, had higher levels of IL-2 expression followed by a higher level of IL-4 expression in the lung, and higher survival compared to Se-adequate mice. These studies demonstrated the essential role of Se in mounting an immune response to influenza A, by changing its virulence and altering the host’s immune response [55][27].
These in vitro and animal studies suggested that in vivo Se supplementation might have a beneficial effect in humans, especially in the elderly, as the immune response is compromised by age. To test this hypothesis, Ivory et al. [56][28] conducted a 12-week randomized, double-blinded, placebo-controlled clinical trial in six groups of individuals with suboptimal Se status or plasma Se levels < 110 ng/mL to observe the response after the flu vaccine was provided. Four groups were given daily capsules of yeast: 20 participants were given 0 μg Se/day (placebo), 18 participants were given 50 μg Se/day, 21 participants received 100 μg Se/day, and 23 received 200 μg Se/day. Two groups were given onion-containing meals, 17 participants received < 1 μg Se/day (unenriched onions), and 18 participants received 50 μg Se/day (Se enriched onions). After 10 weeks of supplementation, all participants were administered the flu vaccine. Selenium supplementation compared to placebo had beneficial and detrimental effects on the cell immunity response to the flu vaccine that was dependent on the type of Se, and dose administered [56][28]. Se-yeast dose of 200 µg/day demonstrated enhanced IL-10 secretion and lower granzyme B content, a cytotoxic protease that induces apoptosis of target cells, within a cluster of differentiation 8 (CD8) cells, while 50 µg/day of Se through the enriched onion meal increased granzyme content and perforin in CD8 cells and reduced natural killer T-cells.
The effectiveness of antiviral agents such as amantadine (AM) [57][29], oseltamivir (OTV) [58][30], β-thujaplicin (TP) [59][31], and ribavirin (RBV) [60][32] to combat viral influence has been limited by the emergence of drug-resistant viruses. Biological Se nanoparticles are increasingly used as an agent to diminish drug resistance by “decorating” the nanoparticles with antiviral drugs to increase effectiveness, such as Se@AM, Se@OTV, Se@TP, Se@RBV. Selenium nanoparticles have been found to decrease oxidative stress, induce apoptosis of infected cells, and reduce lung cell damage during influenza infection, in addition to having low toxicity and increased drug activity in murine [59][31] and in vitro models [60][32].

2.3. Human Immunodeficiency Virus (HIV)

It is estimated that over 37 million people globally are living with HIV [61][33]. HIV is an enveloped, single-stranded RNA virus and without treatment causes a collapse of the immune system. The prevalence of Se deficiency in people living with HIV (PLWH) is reported to be around 7–66% and increases as HIV disease progresses over time [62,63,64,65][34][35][36][37]. Although antiretroviral therapy (ART) has allowed HIV disease to become a chronic disease, the immune system is still not fully reconditioned [66][38]. The rate of Se deficiency in PLWH in Sub-Saharan Africa is greater than that in the United States (U.S.A.) and the literature shows lower Se soil content in Sub-Saharan Africa [67][39]. Selenium deficiency in HIV disease is associated with disease progression and mortality, regardless of ART initiation [63,68,69,70,71][35][40][41][42][43]. Models of simian immunodeficiency virus also corroborate the relationship between Se deficiency and disease progression [72][44].
The relationship between HIV disease and increased oxidative stress [73,74,75,76,77,78][45][46][47][48][49][50] was recorded early in the disease, and the development of ROS and its association with HIV disease progression was documented in the very early stages of the emergence of the disease [79,80][51][52]. Lower GSH levels were found as HIV advances to acquired immunodeficiency syndrome (AIDS) [81][53] and alterations in antioxidant defense systems (SOD, CAT, and GPX) have also been observed in PLWH [75,77,82][47][49][54]. Supplementation of 250 µg of L-selenomethionine (100 µg of Se) for one year led to increased GPX activity [83][55] and adequate dietary Se intake was also associated with lower oxidative stress in PLWH [84][56].
Studies in children and adults living with HIV have found associations with Se deficiency and adverse health outcomes including mortality. Countries with a high prevalence of HIV such as South Africa, have shown that Se intake in children is not adequate and the overall diet quality is low [85][57]. In studies conducted in Nigeria, children with HIV had significantly lower Se levels compared to matched HIV-non-infected children in the same region and a high rate (>70%) of Se deficiency [86,87][58][59]. In children living with HIV in the U.S.A., Se deficiency was associated with advanced immunodeficiency [88][60] and mortality [63][35]. In adult PLWH who were initiating antiretroviral therapy (ART) or were already taking ART, Se deficiency was associated with HIV disease progression and mortality [70,71][42][43]. Additionally, Se values have been found to be lower in adult PLWH than in adults without HIV [89[61][62],90], as well as in later stages of HIV [90][62].
Several Se supplementation trials have been conducted within the United States [91,92][63][64], Tanzania [93[65][66],94], Botswana [95][67], and Rwanda [96][68]. These trials have demonstrated that Se supplementation in the dose of 200 µg in PLWH who are ART naïve or on ART may delay HIV disease progression through maintenance of cluster of differentiation 4 (CD4) cell counts. Hurwitz et al. [92][64] demonstrated that supplementation with Se resulted in significantly suppressed HIV viral load along with improved CD4 cell count. Trials using Se as part of a formula in combination with other micronutrients have not been able to discern the benefits of Se from the other components. We [95][67] concluded that supplementation with multivitamins and Se was safe and statistically significantly reduced the risk of immune decline and morbidity. Discrepancies between supplementation studies include the ART status of the participants, the baseline CD4 cell counts, and the length of time that the participants were supplemented and followed [97][69]. A Cochrane review of micronutrient supplementation and HIV concluded that additional trials with single nutrients were needed to build the evidence base for adults and establish long-term benefits [98][70].

2.4. Hepatitis B and C Viruses (HBV and HCV)

The World Health Organization (WHO) estimates that 257 million people and 71 million people were infected with hepatitis B virus (HBV) and hepatitis C virus (HCV), respectively [99][71]. Both HBV and HCV can cause acute and chronic hepatitis which can develop into cirrhosis and hepatocellular carcinoma (HCC). In 2015, there were 720,000 and 470,000 deaths from hepatic cirrhosis and HCC, respectively [99][71]. Though both HBV and HCV are hepatotropic, HCV belongs to the Flaviviridae family, whereas HBV is a member of the Hepadnaviridae family [100][72]. HBV is a partially double-stranded DNA virus that uses the host RNA polymerase II machinery to produce pre-genomic RNA, which is reverse transcribed into viral DNA [101][73]. HCV is an enveloped, single-stranded RNA virus, which exhibits extremely high mutation rates—up to one mutation per genome per generation cycle—since proofreading activity is lacking in RNA-dependent RNA polymerases required for its replication [102][74].
Selenium status determined by GPX3 activity and the concentration of serum/plasma Se and plasma SELENOP have been reported to be influenced in HBV and HCV patients in several studies. Serum Se concentrations are statistically significantly lower in HBV/HCV infected people when compared with the control group [103,104][75][76]. Selenium level is also associated with the severity and progression of the HBV/HCV disease [103,105,106][75][77][78]. Increased concentrations of aspartate aminotransferase and alanine aminotransferase (ALT) were independently associated with low Se concentration in chronic HBV patients with more hepatic damage [106][78]. The Se concentrations in plasma and erythrocytes are significantly lower in HCV-infected people than in controls and have an inversed correlation with HCV viral load [107][79]. Besides this, plasma Se level is statistically lower in people with HCV-induced cirrhosis with and without HCC when compared with HCV-infected people without liver cirrhosis or HCC [108][80].
Chronic HBV and HCV infection enhances ROS production and cause elevated oxidative stress and decreased antioxidant activity in liver cells [109,110,111,112][81][82][83][84]. ROS, produced as byproducts during cellular metabolism, have been implicated in several hepatic pathologies to maintain cellular homeostasis, including cell signaling, transcription, apoptosis, and immunomodulation [113,114,115,116][85][86][87][88]. Patients suffering from HBV or HCV infection show significant depletion of GSH and GPX when compared to non-infected participants [117,118][89][90]. As part of the antioxidant defense system, Se deficiency may be enhanced by the hepatic viral-induced oxidative stress and the requirement of selenoproteins during viral replication. An in vitro study showed that HCV can inhibit the expression of gastrointestinal-GPX (GPX2), a GPX that is also expressed in the liver, resulting in an increase in viral replication [7,119][7][91]. Nonstructural protein 5A (NS5A) of HCV, which is reported to enhance oxidative stress by perturbing Ca2+ homeostasis [120][92], also induces the expression and activity of GPX1 and GPX4 [121][93]. Besides, the GPX homology region overlaps the highly conserved NS4 gene in HCV, supporting that the NS4 gene is a functional GPX module [122][94]. Although the causes of Se deficiency in HBV and HCV are not fully understood, it is possible that the decreased level of circulating Se is related to the requirement of Se during viral replication.
The demand for Se during HBV and HCV infection causes the systemic deficiency of Se and can be compensated by supplementation. Supplementation of Se has shown to be protective against a wide range of different sources of oxidative stress and optimal immune responses [123,124][95][96]. Primary HCC incidence was reduced by 35.1% in Se supplemented people living with HBV as compared with non-supplemented people living with HBV [125][97]. However, when Se supplementation was stopped, primary HCC incidence began to increase [125][97]. Selenium also improves the rate and level of antibody response against the HBV vaccine in insulin-dependent diabetes mellitus cases that were on an accelerated vaccination schedule instead of a routine vaccine schedule [126][98]. A triple antioxidant combination of Se, alpha-lipoic acid, and silymarin supplementation in three chronic HCV-infected patients demonstrated an improvement in ALT [127][99]. However, a 6-month trial showed that those living with HCV supplemented with vitamins C, vitamin E, and 200 μg Se per day had an increase in antioxidant status with no beneficial effect on ALT, HCV viral load, or liver damage as compared with the non-supplemented individuals living with HCV [128][100].
As discussed in the previous paragraph, Se deficiency has been involved in the pathogenesis of HBV and HCV infection. In turn, the deficiency of Se leads to elevated oxidative stress, pathological changes, and inflammation in the liver [129][101]. Histological study shows hepatic sinus expansion, lymphocyte infiltration, and stripe-like hyperplasia in the liver with Se deficiency. Liver inflammation is initiated by Se deficiency as pro-inflammatory factors and molecules, such as IL-1β, IL-6, IL-12, NF-κβ, and NF-κβ p65, were all significantly higher in the Se-deficient group [129][101]. Hepatic antioxidant capacity is also influenced by Se deficiency as a decrease in both mRNA expression of selenoprotein genes (GPX1 and GPX3), as determined by quantitative real-time PCR and the level of selenoproteins (GPX1, GPX4, and TXNRD1), identified by global proteomics, are observed [129,130,131][101][102][103]. Interestingly, an in vitro study showed that Se deficiency can result in oxidative stress and apoptosis of non-HBV-infected hepatocytes, whereas HBV-infected hepatocytes gain a survival capacity and escape from the apoptosis consequence [132][104].

2.5. Poliovirus

Poliovirus is part of the Picornaviridae family of RNA viruses that are non-enveloped and may infect vertebrate animals [133][105]. Infection generates high levels of ROS and reactive nitrogen species as well as antioxidant enzymes being downregulated within cells that have been infected [134][106]. The supplementation of Se has been shown to improve the response of the vaccine for the poliovirus more in patients that have less optimal immune systems based on Se status, although the impact of supplementation on patients with optimal immune systems based on Se status is unclear [135][107]. Furthermore, the supplementation of Se did not affect all aspects of an individual’s immune response shown in the same trial where a live poliomyelitis vaccine was given to people with low Se status. This resulted in the increase of T cell and IL-10 production but did not affect the natural killer (NK) or B cell count, still resulting in the rapid removal of poliomyelitis from the patients supplemented [136][108]. Selenium also did not affect the levels of CD4+ T helper (Th) 1 cells to Th2 cells or the humoral immune response [135][107] in a different trial where patients were given a dose of the poliovirus vaccine and took either a placebo, 50 µg or 100 µg of Se. An increase in the antibody titers within all groups that were relatively equal was shown [135][107]. Se supplementation prior to the polio vaccine seemed to only enhance the cellular antiviral immune response.

2.6. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2)

The novel COVID-19 is caused by SARS-CoV-2, a single-stranded RNA coronavirus. The severity of the disease has been linked to aging and comorbidities such as hypertension, diabetes, obesity, cardiovascular disease, kidney disease, cancer, and pulmonary diseases [137,138][109][110]. Most of the people who test positive for COVID-19 develop mild or no symptoms, while others develop acute respiratory distress syndrome (ARDS), heart failure, blood clots, neurological complications, and elevated inflammatory response [137,139][109][111]. SARS-CoV-2 pathology has been associated with an increased immune response, leading to a release of cytokines and chemokines, also known as cytokine storm [140][112], as well as increased inflammatory markers such as D-dimer and ferritin [141,142][113][114]. This hyperactive inflammatory response may also bring about severe pathology in the brain [143][115]. SARS-CoV-2 may directly impact the central nervous system and enter the brain through various routes [144,145,146,147][116][117][118][119]. Increased systemic inflammation promoted by SARS-CoV-2 has the potential to disturb the blood-brain barrier and co-morbidities associated with severe cases of COVID-19 may enable the attack of the brain by SARS-CoV-2 [143,148][115][120].
It has been noted that there is a potential and developing relationship between Se levels and COVID-19 outcomes. Proposed mechanisms by which Se may act upon the SARS-CoV-2 virus based on previous research in RNA viruses include restoration of GPX and TXNRD thus reducing oxidative stress, reduction of viral-induced cell apoptosis, provision of Se for the host’s antioxidant needs, protection of endothelial cells, and reduced blood platelet aggregation [149,150][121][122]. COVID-19 is associated with a heightened level of oxidative stress and inflammation that are implicated in the pathogenesis of pulmonary disease [151][123]. GSH provides protection to the epithelial barrier within the lungs, and it has been suggested that improvement of GSH levels would be a strategy that may protect against inflammation and oxidant-related damage in the lungs [151][123]. A study conducted by Mahmoodpoor et al. [152][124] supplemented sodium selenite in patients with ARDS, often associated with severe cases of COVID-19, and found that it restored the antioxidant capability of the lungs, reduced inflammation, and improved respiratory mechanics. Lower total lymphocytes and CD4+ T, CD8+ T, B, and NK cells were found in COVID-19 patients and those with severe cases compared to mild cases of COVID-19 had lower lymphocyte subsets [153][125]. The function and differentiation of B and T cells may be affected by Se status [154][126]. Deficiency of Se in mice has been associated with lower T cell proliferation, while supplementation increased T cell activity and differentiation [155][127].
Clinical data investigating Se and COVID-19 are sparse; however, some reports from China and other countries globally have surfaced. In China, where there is a wide range of soil Se levels and thus a variation of Se daily intake, a linear association has been demonstrated between reported cure rates of COVID-19 and Se hair concentration data, dating from 2011 and older [156][128]. The same research group in China documented higher fatality risk in cities that had selenium-deficient levels in crops and topsoil compared to cities with non-deficient selenium levels in crops and topsoil [157][129]. Intake of Se varies worldwide, and China is known to be one of the most Se deficient countries in the world, with a wide range of levels that differs from lowest to highest in the world. COVID-19 fatality rate varies across different regions in China, suggesting that Se status may be related to COVID-19 outcomes [156,158][128][130]. In the city of Wuhan, where the SARS-CoV-2 virus was first discovered, and in other cities such as Suizhou and Xiaogan, low Se soil status was associated with the highest COVID-19 incidence [156][128]. In contrast, cities such as Enshi, Yichang, and Xiangyan, where high Se intake occurs, had the lowest COVID-19 incidence [156][128]. In contrast, in a retrospective study completed in Wuhan, China, with hospitalized COVID-19 patients, the severity of COVID-19 was associated with higher Se levels in urine [159][131]. The authors hypothesize that liver abnormalities due to the severity of the disease may have impacted the excess urinary Se found in severe COVID-19 patients [159][131].
Studies conducted in other parts of the world are showing similar relationships to those completed in China. In a study conducted in South Korea on hospitalized COVID-19 patients, 42% were found to be Se deficient and as the severity of disease increased, Se plasma levels decreased [160][132]. These patients also experienced additional nutritional deficiencies. COVID-19 patients compared to healthy controls in India, Iran and Russia had significantly lower plasma Se levels [161,162,163][133][134][135]. A greater rate of low plasma Se levels (<70 ng/mL) was found in COVID-19 patients (43%) compared to controls in India (20) [161][133]. Lung damage, as assessed by computer tomography, was inversely associated with Se levels in Russia [163][135].
COVID-19 patients may also experience increases in oxidative stress and increases in Se-related markers and lower Se levels have been documented in these patients. Moghaddam al. [164][136] observed an association between markers of Se status and COVID-19 outcomes from COVID-19 patients in Germany. Serum Se and SELENOP concentrations were lower in COVID-19 patients compared to a reference European population. A comparison of patients that survived compared to those who died from COVID-19 showed that the deceased had a significantly greater deficiency of serum Se and SELENOP concentrations than those who survived. In addition, those who died had significantly lower serum Se, SELENOP levels, and GPX compared with patients who survived. A study in Belgium using a convenience sample of patients hospitalized with severe COVID-19 pneumonia observed statistically lower GSH levels and higher GPX levels compared with reference intervals among other results showing elevated markers of oxidative stress and lower antioxidant status [165][137]. Recently, Polonikov [166][138] hypothesized that GSH deficiency plays a major role in augmenting SARS-CoV-2 oxidative damage, which leads to greater disease progression and mortality. This viewpoint was based on data showing lower GSH and higher ROS levels in COVID-19 patients with mild disease and increasing severity that included higher viral load with GSH deficiency [166][138] and work completed by Hurwitz et al. [167][139] that demonstrated improvement in dyspnea with high dose oral and IV GSH in two patients with underlying conditions who tested positive for COVID-19. These conclusions were based on very small samples and therefore require additional larger clinical studies to replicate the findings and eventual intervention studies. The evidence presented above suggests that Se availability contributes to resisting SARS-CoV-2 infection, corresponding with studies that show adequate levels of Se status maintains an appropriate immune response to viral infection [6,134,136][6][106][108].
There are no known published Se supplementation clinical trials in the context of COVID-19 at this time and one study is currently listed on clincaltrials.gov that will examine the efficacy of Se (selenious acid infusion also known as sodium selenite) for the treatment of moderately-ill, severely ill, and critically ill COVID-19 patients (Identifier: NCT04869579). Sodium selenite supplementation has been proposed for the prevention of COVID-19 infections and severe disease [149,168][121][140]. Sodium selenite is easily available, short-term toxicity is marginal and may cross the blood-brain barrier [149][121]. This chemical form may oxidize thiol groups located in the virus protein disulfide isomerase, which would interfere with its ability to infiltrate the cell membrane and produce an infection [168][140]. TXNRD activity increases quickly after supplementation with sodium selenite in cancer cell lines and critically ill patients [169,170][141][142] and has demonstrated reduced ROS production and viral-induced cell apoptosis in cell culture studies [171][143]. A common feature of COVID-19 is thrombotic complications and altered platelet function is believed to affect the sequelae of this infection [172][144]. Sodium selenite has also been shown to have an anti-aggregating effect through its reduction of thromboxane A2 formation, an important factor in blood platelet activation and formation [173][145]. The effectiveness of sodium selenite for the prevention and management of COVID-19 should be tested immediately as the COVID-19 pandemic continues to persist and threaten the health of individuals globally, thus necessitating rapidly accessible treatment strategies.
Since Se has pronounced therapeutic potential for the treatment of viral infections and other conditions such as cancer, Se nanomedicine has received a lot of attention. Se nanoparticles are known to have low toxicity with marked and selective cytotoxic effects with small quantities [174][146]. Additionally, Se nanoparticles have high effectiveness in the inhibition of oxidative damage [175,176,177][147][148][149]. Recently published data show that Se nanoparticles activate programmed cell death in target cancer tissue through calcium (Ca)2+ signaling pathways [178][150]. Immune cells also require calcium flux to generate oxidative stress [174][146]. Through chemical methods, Se nanoparticles may be produced with Se sources that include sodium selenite, selenious acid, and sodium selenosulfate [174][146]. Due to the developing relationship between Se and COVID-19, Se nanomedicine is being suggested as a tool in the fight against SARS-CoV-2 [179][151]. Currently, there are tremendous prospects of using nanomedicine in ARDS for the prevention, diagnosis, and treatment, which may have applicability for COVID-19 [180][152]. Jin et al. [181][153] discovered that an organic Se compound known as Ebselen, and a promising antioxidant drug, could inhibit SARS-CoV-2 by penetrating the cell membrane and displaying antiviral activity. Ebselen is known to have anti-inflammatory activity, mimic GPX activity, and should be considered for clinical studies [181,182][153][154].

References

  1. Krause, R.M. The Origin of Plagues: Old and New. Science 1992, 257, 1073–1078.
  2. Piret, J.; Boivin, G. Pandemics Throughout History. Front. Microbiol. 2021, 11, 3594.
  3. Parvez, M.K.; Parveen, S. Evolution and Emergence of Pathogenic Viruses: Past, Present, and Future. Intervirology 2017, 60, 1–7.
  4. WHO Coronavirus (COVID-19) Dashboard. WHO Coronavirus (COVID-19) Dashboard with Vaccination Data. Available online: https://covid19.who.int/ (accessed on 18 August 2021).
  5. Avery, J.C.; Hoffmann, P.R. Selenium, Selenoproteins, and Immunity. Nutrients 2018, 10, 1203.
  6. Guillin, O.M.; Vindry, C.; Ohlmann, T.; Chavatte, L. Selenium, Selenoproteins and Viral Infection. Nutrients 2019, 11, 2101.
  7. Hariharan, S.; Dharmaraj, S. Selenium and Selenoproteins: It’s Role in Regulation of Inflammation. Inflammopharmacology 2020, 28, 1.
  8. Rayman, M.P. Selenium and Human Health. Lancet 2012, 379, 1256–1268.
  9. Fairweather-Tait, S.J.; Bao, Y.; Broadley, M.R.; Collings, R.; Ford, D.; Hesketh, J.E.; Hurst, R. Selenium in Human Health and Disease. Antioxid. Redox Signal. 2011, 14, 1337–1383.
  10. Beck, M.A.; Levander, O.A.; Handy, J. Selenium Deficiency and Viral Infection. J. Nutr. 2003, 133, 1463S–1467S.
  11. Sheridan, D.; Zhong, N.; Carlson, B.; Perella, C.; Hatfield, D.; Beck, M. Decreased Selenoprotein Expression Alters the Immune Response during Influenza Virus Infection in Mice. J. Nutr. 2007, 137, 1466–1471.
  12. Stỳblo, M.; Walton, F.; Harmon, A.; Sheridan, P.; Beck, M. Activation of Superoxide Dismutase in Selenium-Deficient Mice Infected with Influenza Virus. J. Trace Elem. Med. Biol. 2007, 21, 52–62.
  13. Beck, M.; Handy, J.; Levander, O. Host Nutritional Status: The Neglected Virulence Factor. Trends Microbiol. 2004, 12, 417–423.
  14. Labunskyy, V.M.; Hatfield, D.L.; Gladyshev, V.N. Selenoproteins: Molecular Pathways and Physiological Roles. Physiol. Rev. 2014, 94, 739.
  15. Beck, M. Selenium and Host Defence towards Viruses. Proc. Nutr. Soc. 1999, 58, 707–711.
  16. Beck, M.A.; Shi, Q.; Morris, V.C.; Levander, O.A. Rapid Genomic Evolution of a Non-Virulent Coxsackievirus B3 in Selenium-Deficient Mice Results in Selection of Identical Virulent Isolates. Nat. Med. 1995, 1, 433–436.
  17. GQ, Y.; JS, C.; ZM, W.; KY, G.; LZ, Z.; XC, C.; XS, C. The Role of Selenium in Keshan Disease. Adv. Nutr. Res. 1984, 6, 203–231.
  18. Xu, G.; Wang, S.; Gu, B.; Yang, Y.; Song, H.; Xue, W.; Liang, W.; Zhang, P. Further Investigation on the Role of Selenium Deficiency in the Aetiology and Pathogenesis of Keshan Disease. Biomed. Environ. Sci. 1997, 10, 316–326.
  19. Cheng, Y.Y.; Qian, P.C. The Effect of Selenium-Fortified Table Salt in the Prevention of Keshan Disease on a Population of 1.05 Million. Biomed. Environ. Sci. 1990, 3, 422–428.
  20. Wei-Han, Y. A Study of Nutritional and Bio-Geochemical Factors in the Occurrence and Development of Keshan Disease : The 6th Conference on Prevention for Rheumatic Fever and Rheumatic Heart Disease. JPN Circ. J. 1982, 46, 1201–1207.
  21. Beck, M.A.; Kolbeck, P.C.; Rohr, L.H.; Shi, Q.; Morris, V.C.; Levander, O.A. Benign Human Enterovirus Becomes Virulent in Selenium-Deficient Mice. J. Med. Virol. 1994, 43, 166–170.
  22. Beck, M.A.; Williams-Toone, D.; Levander, O.A. Coxsackievirus B3-Resistant Mice Become Susceptible in Se/Vitamin E Deficiency. Free Radic. Biol. Med. 2003, 34, 1263–1270.
  23. Beck, M.A.; Kolbeck, P.C.; Shi, Q.; Rohr, L.H.; Morris, V.C.; Levander, O.A. Increased Virulence of a Human Enterovirus (Coxsackievirus B3) in SeleniumDeficient Mice. J. Infect. Dis. 1994, 170, 351–357.
  24. Beck, M.A.; Esworthy, R.S.; Ho, Y.-S.; Chu, F.-F. Glutathione Peroxidase Protects Mice from Viral-Induced Myocarditis. FASEB J. 1998, 12, 1143–1149.
  25. Jaspers, I.; Zhang, W.; Brighton, L.E.; Carson, J.L.; Styblo, M.; Beck, M.A. Selenium Deficiency Alters Epithelial Cell Morphology and Responses to Influenza. Free Radic. Biol. Med. 2007, 42, 1826.
  26. Beck, M.A.; Nelson, H.K.; Shi, Q.; Dael, P.; van Schiffrin, E.J.; Blum, S.; Barclay, D.; Levander, O.A. Selenium Deficiency Increases the Pathology of an Influenza Virus Infection. FASEB J. 2001, 15, 1481–1483.
  27. Li, W.; Beck, M.A. Selenium Deficiency Induced an Altered Immune Response and Increased Survival Following Influenza A/Puerto Rico/8/34 Infection. Exp. Biol. Med. 2017, 232, 412–419.
  28. Ivory, K.; Prieto, E.; Spinks, C.; Armah, C.N.; Goldson, A.J.; Dainty, J.R.; Nicoletti, C. Selenium Supplementation Has Beneficial and Detrimental Effects on Immunity to Influenza Vaccine in Older Adults. Clin. Nutr. 2017, 36, 407.
  29. Li, Y.; Lin, Z.; Guo, M.; Zhao, M.; Xia, Y.; Wang, C.; Xu, T.; Zhu, B. Inhibition of H1N1 Influenza Virus-Induced Apoptosis by Functionalized Selenium Nanoparticles with Amantadine through ROS-Mediated AKT Signaling Pathways. Int. J. Nanomed. 2018, 13, 2005.
  30. Li, Y.; Lin, Z.; Guo, M.; Xia, Y.; Zhao, M.; Wang, C.; Xu, T.; Chen, T.; Zhu, B. Inhibitory Activity of Selenium Nanoparticles Functionalized with Oseltamivir on H1N1 Influenza Virus. Int. Nanomed. 2017, 12, 5733.
  31. Wang, C.; Chen, H.; Chen, D.; Zhao, M.; Lin, Z.; Guo, M.; Xu, T.; Chen, Y.; Hua, L.; Lin, T.; et al. The Inhibition of H1N1 Influenza Virus-Induced Apoptosis by Surface Decoration of Selenium Nanoparticles with β-Thujaplicin through Reactive Oxygen Species-Mediated AKT and P53 Signaling Pathways. ACS Omega 2020, 5, 30633.
  32. Lin, Z.; Li, Y.; Gong, G.; Xia, Y.; Wang, C.; Chen, Y.; Hua, L.; Zhong, J.; Tang, Y.; Liu, X.; et al. Restriction of H1N1 Influenza Virus Infection by Selenium Nanoparticles Loaded with Ribavirin via Resisting Caspase-3 Apoptotic Pathway. Int. J. Nanomed. 2018, 13, 5787.
  33. HIV/AIDS. Available online: https://www.who.int/data/gho/data/themes/hiv-aids (accessed on 25 August 2021).
  34. Baum, M.K. Role of Micronutrients in HIV-Infected Intravenous Drug Users. J. Acquir. Immune Defic. Syndr. 2000, 25 (Suppl. S1), S49–S52.
  35. Campa, A.; Shor-Posner, G.; Indacochea, F.; Zhang, G.; Lai, H.; Asthana, D.; Scott, G.B.; Baum, M.K. Mortality Risk in Selenium-Deficient HIV-Positive Children. J. Acquir. Immune Defic. Syndr. Hum. Retrovirol. 1999, 20, 508–513.
  36. Osuna-Padilla, I.A.; Briceño, O.; Aguilar-Vargas, A.; Rodríguez-Moguel, N.C.; Villazon-De la Rosa, A.; Pinto-Cardoso, S.; Flores-Murrieta, F.J.; Perichart-Perera, O.; Tolentino-Dolores, M.; Vargas-Infante, Y.; et al. Zinc and Selenium Indicators and Their Relation to Immunologic and Metabolic Parameters in Male Patients with Human Immunodeficiency Virus. Nutrition 2020, 70, 110585.
  37. Shivakoti, R.; Gupte, N.; Yang, W.T.; Mwelase, N.; Kanyama, C.; Tang, A.M.; Pillay, S.; Samaneka, W.; Riviere, C.; Berendes, S.; et al. Pre-Antiretroviral Therapy Serum Selenium Concentrations Predict WHO Stages 3, 4 or Death but Not Virologic Failure Post-Antiretroviral Therapy. Nutrients 2014, 6, 5061.
  38. Bloch, M.; John, M.; Smith, D.; Rasmussen, T.A.; Wright, E. Managing HIV-Associated Inflammation and Ageing in the Era of Modern ART. HIV Med. 2020, 21, 2–16.
  39. Hurst, R.; Siyame, E.W.P.; Young, S.D.; Chilimba, A.D.C.; Joy, E.J.M.; Black, C.R.; Ander, E.L.; Watts, M.J.; Chilima, B.; Gondwe, J.; et al. Soil-Type Influences Human Selenium Status and Underlies Widespread Selenium Deficiency Risks in Malawi. Sci. Rep. 2013, 3, 1425.
  40. Stone, C.A.; Kawai, K.; Kupka, R.; Fawzi, W.W. Role of Selenium in HIV Infection. Nutr. Rev. 2010, 68, 671–681.
  41. Kupka, R.; Msamanga, G.I.; Spiegelman, D.; Rifai, N.; Hunter, D.J.; Fawzi, W.W. Selenium Levels in Relation to Morbidity and Mortality among Children Born to HIV-Infected Mothers. Eur. J. Clin. Nutr. 2005, 59, 1250–1258.
  42. Baum, M.K.; Shor-Posner, G.; Lai, S.; Zhang, G.; Lai, H.; Fletcher, M.A.; Sauberlich, H.; Page, J.B. High Risk of HIV-Related Mortality Is Associated with Selenium Deficiency. J. Acquir. Immune Defic. Syndr. Hum. Retrovirol. 1997, 15, 370–374.
  43. Constans, J.; Pellegrin, J.L.; Sergeant, C.; Simonoff, M.; Pellegrin, I.; Fleury, H.; Leng, B.; Conri, C. Serum Selenium Predicts Outcome in HIV Infection. J. Acquir. Immune Defic. Syndr. Hum. Retrovirol. 1995, 10, 392.
  44. Xu, X.M.; Carlson, B.A.; Grimm, T.A.; Kutza, J.; Berry, M.J.; Arreola, R.; Fields, K.H.; Shanmugam, I.; Jeang, K.T.; Oroszlan, S.; et al. Rhesus Monkey Simian Immunodeficiency Virus Infection as a Model for Assessing the Role of Selenium in AIDS. J. Acquir. Immune Defic. Syndr. 2002, 31, 453–463.
  45. Repetto, M.; Reides, C.; Gomez Carretero, M.L.; Costa, M.; Griemberg, G.; Llesuy, S. Oxidative Stress in Blood of HIV Infected Patients. Clin. Chim. Acta 1996, 255, 107–117.
  46. Suresh, D.R.; Annam, V.; Pratibha, K.; Prasad, B.V.M. Total Antioxidant Capacity a Novel Early Bio-Chemical Marker of Oxidative Stress in HIV Infected Individuals. J. Biomed. Sci. 2009, 16, 61.
  47. Ogunro, P.S.; Ogungbamigbe, T.O.; Elemie, P.O.; Egbewale, B.E.; Adewole, T.A. Plasma Selenium Concentration and Glutathione Peroxidase Activity in HIV-1/AIDS Infected Patients: A Correlation with the Disease Progression. Niger. Postgrad. Med. J. 2006, 13, 1–5.
  48. Pace, G.W.; Leaf, C.D. The Role of Oxidative Stress in HIV Disease. Free Radic. Biol. Med. 1995, 19, 523–528.
  49. Yano, S.; Colon, M.; Yano, N. An Increase of Acidic Isoform of Catalase in Red Blood Cells from HIV(+) Population. Mol. Cell. Biochem. 1996, 165, 77–81.
  50. Gil, L.; Martínez, G.; González, I.; Tarinas, A.; Álvarez, A.; Giuliani, A.; Molina, R.; Tápanes, R.; Pérez, J.; León, O.S. Contribution to Characterization of Oxidative Stress in HIV/AIDS Patients. Pharmacol. Res. 2003, 47, 217–224.
  51. Papadopulos-Eleopulos, E. Reappraisal of Aids—Is the Oxidation Induced by the Risk Factors the Primary Cause? Med. Hypotheses 1988, 25, 151–162.
  52. Papadopulos-Eleopulos, E.; Hedland-Thomas, B.; Causer, D.A.; Dufty, A.N.P. An Alternative Explanation for the Radiosensitization of AIDS Patients. Int. J. Radiat. Oncol. Biol. Phys. 1989, 17, 695–697.
  53. Buhl, R.; Holroyd, K.J.; Mastrangeli, A.; Cantin, A.M.; Jaffe, H.A.; Wells, F.B.; Saltini, C.; Crystal, R.G. Systemic Glutathione Deficiency In Symptom-Free Hiv-Seropositive Individuals. Lancet 1989, 334, 1294–1298.
  54. Skurnick, J.; Bogden, J.; Baker, H.; Kemp, F.; Sheffet, A.; Quattrone, G.; Louria, D. Micronutrient Profiles in HIV-1-Infected Heterosexual Adults. J. Acquir. Immune Defic. Syndr. Hum. Retrovirol. 1996, 12, 75–83.
  55. Delmas-Beauvieux, M.C.; Peuchant, E.; Couchouron, A.; Constans, J.; Sergeant, C.; Simonoff, M.; Pellegrin, J.L.; Leng, B.; Conri, C.; Clerc, M. The Enzymatic Antioxidant System in Blood and Glutathione Status in Human Immunodeficiency Virus (HIV)-Infected Patients: Effects of Supplementation with Selenium or Beta-Carotene. Am. J. Clin. Nutr. 1996, 64, 101–107.
  56. McDermid, J.M.; Lalonde, R.G.; Gray-Donald, K.; Baruchel, S.; Kubow, S. Associations between Dietary Antioxidant Intake and Oxidative Stress in HIV-Seropositive and HIV-Seronegative Men and Women. J. Acquir. Immune Defic. Syndr. 2002, 29, 158–164.
  57. Shiau, S.; Webber, A.; Strehlau, R.; Patel, F.; Coovadia, A.; Kozakowski, S.; Brodlie, S.; Yin, M.T.; Kuhn, L.; Arpadi, S.M. Dietary Inadequacies in HIV-Infected and Uninfected School-Aged Children in Johannesburg, South Africa HHS Public Access. J. Pediatr. Gastroenterol. Nutr. 2017, 65, 332–337.
  58. Anyabolu, H.C.; Adejuyigbe, E.A.; Adeodu, O.O. Serum Micronutrient Status of Haart-Naïve, HIV Infected Children in South Western Nigeria: A Case Controlled Study. AIDS Res. Treat 2014, 2014, 351043.
  59. Ubesie, A.C.; Ibe, B.C.; Emodi, I.J.; Iloh, K.K. Serum Selenium Status of HIV-Infected Children on Care and Treatment in Enugu, Nigeria. SAJCH 2017, 11, 21–25.
  60. Bologna, R.; Indacochea, F.; Shor-Posner, G.; Mantero-Atienza, E.; Grazziutti, M.; Sotomayor, M.C.; Fletcher, M.A.; Cabrejos, C.; Scott, G.B.; Baum, M.K. Selenium and Immunity in HIV-1 Infected Pediatric Patients. J. Nutr. Immunol. 1994, 3, 41–49.
  61. Allavena, C.; Dousset, B.; May, T.; Dubois, F.; Canton, P.; Belleville, F. Relationship of Trace Element, Immunological Markers, and HIV1 Infection Progression. Biol. Trace Elem. Res. 1995, 47, 133–138.
  62. Look, M.; Rockstroh, J.; Rao, G.; Kreuzer, K.-A.; Barton, S.; Lemoch, H.; Sudhop, T.; Hoch, J.; Stockinger, K.; Spengler, U.; et al. Serum Selenium, Plasma Glutathione (GSH) and Erythrocyte Glutathione Peroxidase (GSH-Px)-Levels in Asymptomatic versus Symptomatic Human Immunodeficiency Virus-1 (HIV-1)-Infection. Eur. J. Clin. Nutr. 1997, 51, 266–272.
  63. Burbano, X.; Miguez-Burbano, M.J.; McCollister, K.; Zhang, G.; Rodriguez, A.; Ruiz, P.; Lecusay, R.; Shor-Posner, G. Impact of a Selenium Chemoprevention Clinical Trial on Hospital Admissions of HIV-Infected Participants. HIV Clin. Trials 2002, 3, 483–491.
  64. Hurwitz, B.E.; Klaus, J.R.; Llabre, M.M.; Gonzalez, A.; Lawrence, P.J.; Maher, K.J.; Greeson, J.M.; Baum, M.K.; Shor-Posner, G.; Skyler, J.S.; et al. Suppression of Human Immunodeficiency Virus Type 1 Viral Load with Selenium Supplementation: A Randomized Controlled Trial. Arch. Intern. Med. 2007, 167, 148–154.
  65. Kupka, R.; Mugusi, F.; Aboud, S.; Hertzmark, E.; Spiegelman, D.; Fawzi, W.W. Effect of Selenium Supplements on Hemoglobin Concentration and Morbidity among HIV-1-Infected Tanzanian Women. Clin. Infect. Dis. 2009, 48, 1475–1478.
  66. Kupka, R.; Mugusi, F.; Aboud, S.; Msamanga, G.I.; Finkelstein, J.L.; Spiegelman, D.; Fawzi, W.W. Randomized, Double-Blind, Placebo-Controlled Trial of Selenium Supplements among HIV-Infected Pregnant Women in Tanzania: Effects on Maternal and Child Outcomes. Am. J. Clin. Nutr. 2008, 87, 1802–1808.
  67. Baum, M.K.; Campa, A.; Lai, S.; Sales Martinez, S.; Tsalaile, L.; Burns, P.; Farahani, M.; Li, Y.; van Widenfelt, E.; Page, J.B.; et al. Effect of Micronutrient Supplementation on Disease Progression in Asymptomatic, Antiretroviral-Naive, HIV-Infected Adults in Botswana: A Randomized Clinical Trial. JAMA 2013, 310, 2154–2163.
  68. Kamwesiga, J.; Mutabazi, V.; Kayumba, J.; Tayari, J.C.K.; Uwimbabazi, J.C.; Batanage, G.; Uwera, G.; Baziruwiha, M.; Ntizimira, C.; Murebwayire, A.; et al. Effect of Selenium Supplementation on CD4R T-Cell Recovery, Viral Suppression and Morbidity of HIV-Infected Patients in Rwanda: A Randomized Controlled Trial. AIDS 2015, 29, 1045–1052.
  69. Muzembo, B.A.; Ngatu, N.R.; Januka, K.; Huang, H.L.; Nattadech, C.; Suzuki, T.; Wada, K.; Ikeda, S. Selenium Supplementation in HIV-Infected Individuals: A Systematic Review of Randomized Controlled Trials. Clin. Nutr. ESPEN 2019, 34, 1–7.
  70. Visser, M.E.; Durao, S.; Sinclair, D.; Irlam, J.H.; Siegfried, N. Micronutrient Supplementation in Adults with HIV Infection. Cochrane Database Syst. Rev. 2017, 2017, CD003650.
  71. Global Hepatitis Report. 2017. Available online: https://www.who.int/publications/i/item/global-hepatitis-report-2017 (accessed on 16 August 2021).
  72. Wieland, S.F.; Chisari, F.V. Stealth and Cunning: Hepatitis B and Hepatitis C Viruses. J. Virol. 2005, 79, 9369.
  73. Tsukuda, S.; Watashi, K. Hepatitis B Virus Biology and Life Cycle. Antiviral Res. 2020, 182, 104925.
  74. Lauring, A.S.; Frydman, J.; Andino, R. The Role of Mutational Robustness in RNA Virus. Nat. Rev. Microbiol. 2013, 11, 327.
  75. Khan, M.S.; Dilawar, S.; Ali, I.; Rauf, N. The Possible Role of Selenium Concentration in Hepatitis B and C Patients. Saudi J. Gastroenterol. 2012, 18, 106.
  76. Rauf, N.; Tahir, S.S.; Dilawar, S.; Ahmad, I.; Parvez, S. Serum Selenium Concentration in Liver Cirrhotic Patients Suffering from Hepatitis B and C in Pakistan. Biol. Trace Elem. Res. 2011, 145, 144–150.
  77. Himoto, T.; Yoneyama, H.; Kurokohchi, K.; Inukai, M.; Masugata, H.; Goda, F.; Haba, R.; Watababe, S.; Kubota, S.; Senda, S.; et al. Selenium Deficiency Is Associated with Insulin Resistance in Patients with Hepatitis C Virus–Related Chronic Liver Disease. Nutr. Res. 2011, 31, 829–835.
  78. Abediankenari, S.; Ghasemi, M.; Nasehi, M.M.; Abedi, S.; Hosseini, V. Determination of Trace Elements in Patients with Chronic Hepatitis B. Acta Med. Iran. 2011, 49, 667–669.
  79. Ko, W.S.; Guo, C.-H.; Yeh, M.-S.; Lin, L.Y.; Hsu, G.S.W.; Chen, P.C.; Luo, M.C.; Lin, C.Y. Blood Micronutrient, Oxidative Stress, and Viral Load in Patients with Chronic Hepatitis C. World J. Gastroenterol. 2005, 11, 4697.
  80. Bettinger, D.; Schultheiss, M.; Hennecke, N.; Panther, E.; Knüppel, E.; Blum, H.E.; Thimme, R.; Spangenberg, H.C. Selenium Levels in Patients with Hepatitis C Virus-Related Chronic Hepatitis, Liver Cirrhosis, and Hepatocellular Carcinoma: A Pilot Study. Hepatology 2013, 57, 2543–2544.
  81. Reshi, M.L.; Su, Y.-C.; Hong, J.R. RNA Viruses: ROS-Mediated Cell Death. Int. J. Cell Biol. 2014, 2014, 467452.
  82. Yu, D.Y. Relevance of Reactive Oxygen Species in Liver Disease Observed in Transgenic Mice Expressing the Hepatitis B Virus X Protein. Lab. Anim. Res. 2020, 36, 6.
  83. Yuan, K.; Lei, Y.; Chen, H.-N.; Chen, Y.; Zhang, T.; Li, K.; Xie, N.; Wang, K.; Feng, X.; Pu, Q.; et al. HBV-Induced ROS Accumulation Promotes Hepatocarcinogenesis through Snail-Mediated Epigenetic Silencing of SOCS3. Cell Death Differ. 2016, 23, 616.
  84. Jain, S.K.; Pemberton, P.W.; Smith, A.; McMahon, R.F.T.; Burrows, P.C.; Aboutwerat, A.; Warnes, T.W. Oxidative Stress in Chronic Hepatitis C: Not Just a Feature of Late Stage Disease. J. Hepatol. 2002, 36, 805–811.
  85. Su, L.J.; Zhang, J.-H.; Gomez, H.; Murugan, R.; Hong, X.; Xu, D.; Jiang, F.; Peng, Z.-Y. Reactive Oxygen Species-Induced Lipid Peroxidation in Apoptosis, Autophagy, and Ferroptosis. Oxid. Med. Cell. Longev. 2019, 2019, 5080843.
  86. Zhang, J.; Wang, X.; Vikash, V.; Ye, Q.; Wu, D.; Liu, Y.; Dong, W. ROS and ROS-Mediated Cellular Signaling. Oxid. Med. Cell. Longev. 2016, 2016, 4350965.
  87. Ohl, K.; Tenbrock, K. Reactive Oxygen Species as Regulators of MDSC-Mediated Immune Suppression. Front. Immunol. 2018, 9, 2499.
  88. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The Immunomodulatory and Anti-Inflammatory Role of Polyphenols. Nutrients 2018, 10, 1618.
  89. Razzaq, Z.; Malik, A. Viral Load Is Associated with Abnormal Serum Levels of Micronutrients and Glutathione and Glutathione-Dependent Enzymes in Genotype 3 HCV Patients. BBA Clin. 2014, 2, 72.
  90. Kundu, D.; Roy, A.; Mandal, T.; Bandyopadhyay, U.; Ghosh, E.; Ray, D. Oxidative Stress in Alcoholic and Viral Hepatitis. N. Am. J. Med. Sci. 2012, 4, 412.
  91. Morbitzer, M.; Herget, T. Expression of Gastrointestinal Glutathione Peroxidase Is Inversely Correlated to the Presence of Hepatitis C Virus Subgenomic RNA in Human Liver Cells. J. Biol. Chem. 2005, 280, 8831–8841.
  92. Dionisio, N.; Garcia-Mediavilla, M.; Sanchez-Campos, S.; Majano, P.L.; Benedicto, I.; Rosado, J.A.; Salido, G.M.; Gonzalez-Gallego, J. Hepatitis C Virus NS5A and Core Proteins Induce Oxidative Stress-Mediated Calcium Signalling Alterations in Hepatocytes. J. Hepatol. 2009, 50, 872–882.
  93. Brault, C.; Lévy, P.; Duponchel, S.; Michelet, M.; Sallé, A.; Pécheur, E.-I.; Plissonnier, M.-L.; Parent, R.; Véricel, E.; Ivanov, A.V.; et al. Glutathione Peroxidase 4 Is Reversibly Induced by HCV to Control Lipid Peroxidation and to Increase Virion Infectivity. Gut 2016, 65, 144–154.
  94. Zhang, W.; Cox, A.G.; Taylor, E.W. Hepatitis C Virus Encodes a Selenium-Dependent Glutathione Peroxidase Gene. Med. Klin 1999, 94, 2–6.
  95. Kiełczykowska, M.; Kocot, J.; Paździor, M.; Musik, I. Selenium—A Fascinating Antioxidant of Protective Properties. Adv. Clin. Exp. Med. 2018, 27, 245–255.
  96. Bentley-Hewitt, K.L.; Chen, R.K.-Y.; Lill, R.E.; Hedderley, D.I.; Herath, T.D.; Matich, A.J.; McKenzie, M.J. Consumption of Selenium-Enriched Broccoli Increases Cytokine Production in Human Peripheral Blood Mononuclear Cells Stimulated Ex Vivo, a Preliminary Human Intervention Study. Mol. Nutr. Food Res. 2014, 58, 2350–2357.
  97. Yu Yu, S.; Zhu, Y.J.; Li, W.G. Protective Role of Selenium against Hepatitis B Virus and Primary Liver Cancer in Qidong. Biol. Trace Elem. Res. 1997, 56, 117–124.
  98. Janbakhsh, A.; Mansouri, F.; Vaziri, S.; Sayad, B.; Afsharian, M.; Rahimi, M.; Shahebrahimi, K.; Salari, F. Effect of Selenium on Immune Response against Hepatitis B Vaccine with Accelerated Method in Insulin-Dependent Diabetes Mellitus Patients. Caspian J. Intern. Med. 2013, 4, 603.
  99. Berkson, B.M. A Conservative Triple Antioxidant Approach to the Treatment of Hepatitis C. Med. Klin. 1999, 94, 84–89.
  100. Groenbaek, K.; Friis, H.; Hansen, M.; Ring-Larsen, H.; Krarup, H.B. The Effect of Antioxidant Supplementation on Hepatitis C Viral Load, Transaminases and Oxidative Status: A Randomized Trial among Chronic Hepatitis C Virus-Infected Patients. Eur. J. Gastroenterol. Hepatol. 2006, 18, 985–989.
  101. Tang, C.; Li, S.; Zhang, K.; Li, J.; Han, Y.; Zhan, T.; Zhao, Q.; Guo, X.; Zhang, J. Selenium Deficiency-Induced Redox Imbalance Leads to Metabolic Reprogramming and Inflammation in the Liver. Redox Biol. 2020, 36, 101519.
  102. Zhang, Y.; Yu, D.; Zhang, J.; Bao, J.; Tang, C.; Zhang, Z. The Role of Necroptosis and Apoptosis through the Oxidative Stress Pathway in the Liver of Selenium-Deficient Swine. Metallomics 2020, 12, 607–616.
  103. Burk, R.F.; Hill, K.E.; Nakayama, A.; Mostert, V.; Levander, X.A.; Motley, A.K.; Freeman, M.L.; Austin, L.M. Selenium Deficiency Activates Mouse Liver Nrf2-ARE but Vitamin E Deficiency Does Not. Free Radic. Biol. Med. 2008, 44, 1617.
  104. Irmak, M.; Ince, G.; Ozturk, M.; Cetin-Atalay, R. Acquired Tolerance of Hepatocellular Carcinoma Cells to Selenium Deficiency: A Selective Survival Mechanism? Cancer Res. 2003, 63, 6707–6715.
  105. Burrill, C.P.; Westesson, O.; Schulte, M.B.; Strings, V.R.; Segal, M.; Andino, R. Global RNA Structure Analysis of Poliovirus Identifies a Conserved RNA Structure Involved in Viral Replication and Infectivity. J. Virol. 2013, 87, 11670–11683.
  106. Steinbrenner, H.; Al-Quraishy, S.; Dkhil, M.A.; Wunderlich, F.; Sies, H. Dietary Selenium in Adjuvant Therapy of Viral and Bacterial Infections. Adv. Nutr. 2015, 6, 73–82.
  107. Broome, C.S.; McArdle, F.; Kyle, J.A.M.; Andrews, F.; Lowe, N.M.; Hart, C.A.; Arthur, J.R.; Jackson, M.J. An Increase in Selenium Intake Improves Immune Function and Poliovirus Handling in Adults with Marginal Selenium Status. Am. J. Clin. Nutr. 2004, 80, 154–162.
  108. Hoffmann, P.R.; Berry, M.J. The Influence of Selenium on Immune Responses. Mol. Nutr. Food Res. 2008, 52, 1273–1280.
  109. Zhou, F.; Yu, T.; Du, R.; Fan, G.; Liu, Y.; Liu, Z.; Xiang, J.; Wang, Y.; Song, B.; Gu, X.; et al. Clinical Course and Risk Factors for Mortality of Adult Inpatients with COVID-19 in Wuhan, China: A Retrospective Cohort Study. Lancet 2020, 395, 1054.
  110. Dai, M.; Liu, D.; Liu, M.; Zhou, F.; Li, G.; Chen, Z.; Zhang, Z.; You, H.; Wu, M.; Zheng, Q.; et al. Patients with Cancer Appear More Vulnerable to SARS-CoV-2: A Multicenter Study during the COVID-19 Outbreak. Cancer Discov. 2020, 10, 783.
  111. Wiersinga, W.J.; Rhodes, A.; Cheng, A.C.; Peacock, S.J.; Prescott, H.C. Pathophysiology, Transmission, Diagnosis, and Treatment of Coronavirus Disease 2019 (COVID-19): A Review. JAMA 2020, 324, 782–793.
  112. Song, P.; Li, W.; Xie, J.; Hou, Y.; You, C. Cytokine Storm Induced by SARS-CoV-2. Clin. Chim. Acta 2020, 509, 280.
  113. Mehta, P.; McAuley, D.F.; Brown, M.; Sanchez, E.; Tattersall, R.S.; Manson, J.J.; Collaboration, H.A.S. UK COVID-19: Consider Cytokine Storm Syndromes and Immunosuppression. Lancet 2020, 395, 1033.
  114. Li, Y.; Zhao, K.; Wei, H.; Chen, W.; Wang, W.; Jia, L.; Liu, Q.; Zhang, J.; Shan, T.; Peng, Z.; et al. Dynamic Relationship between D-dimer and COVID-19 Severity. Br. J. Haematol. 2020, 190, e24–e27.
  115. Valenza, M.; Steardo, L., Jr.; Steardo, L.; Verkhratsky, A.; Scuderi, C. Systemic Inflammation and Astrocyte Reactivity in the Neuropsychiatric Sequelae of COVID-19: Focus on Autism Spectrum Disorders. Front. Cell. Neurosci. 2021, 15, 748136.
  116. Zhou, Z.; Kang, H.; Li, S.; Zhao, X. Understanding the Neurotropic Characteristics of SARS-CoV-2: From Neurological Manifestations of COVID-19 to Potential Neurotropic Mechanisms. J. Neurol. 2020, 267, 1.
  117. Nemoto, W.; Yamagata, R.; Nakagawasai, O.; Nakagawa, K.; Hung, W.Y.; Fujita, M.; Tadano, T.; Tan-No, K. Effect of Spinal Angiotensin-Converting Enzyme 2 Activation on the Formalin-Induced Nociceptive Response in Mice. Eur. J. Pharmacol. 2020, 872, 172950.
  118. Satarker, S.; Nampoothiri, M. Involvement of the Nervous System in COVID-19: The Bell Should Toll in the Brain. Life Sci. 2020, 262, 118568.
  119. Merad, M.; Martin, J.C. Pathological Inflammation in Patients with COVID-19: A Key Role for Monocytes and Macrophages. Nat. Rev. Immunol. 2020, 20, 1.
  120. Erickson, M.A.; Rhea, E.M.; Knopp, R.C.; Banks, W.A. Interactions of SARS-CoV-2 with the Blood–Brain Barrier. Int. J. Mol. Sci. 2021, 22, 2681.
  121. Hiffler, L.; Rakotoambinina, B. Selenium and RNA Virus Interactions: Potential Implications for SARS-CoV-2 Infection (COVID-19). Front. Nutr. 2020, 7, 164.
  122. Liu, Q.; Zhao, X.; Ma, J.; Mu, Y.; Wang, Y.; Yang, S.; Wu, Y.; Wu, F.; Zhou, Y. Selenium (Se) Plays a Key Role in the Biological Effects of Some Viruses: Implications for COVID-19. Environ. Res. 2021, 196, 110984.
  123. Samir, D. Oxidative Stress Associated with SARS-CoV-2 (COVID-19) Increases the Severity of the Lung Disease—A Systematic Review. J. Infect. Dis. Epidemiol. 2020, 6, 121.
  124. Mahmoodpoor, A.; Hamishehkar, H.; Shadvar, K.; Ostadi, Z.; Sanaie, S.; Saghaleini, S.H.; Nader, N.D. The Effect of Intravenous Selenium on Oxidative Stress in Critically Ill Patients with Acute Respiratory Distress Syndrome. Immunol. Investig. 2019, 48, 147–159.
  125. Wang, F.; Nie, J.; Wang, H.; Zhao, Q.; Xiong, Y.; Deng, L.; Song, S.; Ma, Z.; Mo, P.; Zhang, Y. Characteristics of Peripheral Lymphocyte Subset Alteration in COVID-19 Pneumonia. J. Infect. Dis. 2020, 221, 1762–1769.
  126. Huang, Z.; Rose, A.H.; Hoffmann, P.R. The Role of Selenium in Inflammation and Immunity: From Molecular Mechanisms to Therapeutic Opportunities. Antioxid. Redox Signal. 2012, 16, 705.
  127. Hoffmann, F.K.W.; Hashimoto, A.C.; Shafer, L.A.; Dow, S.; Berry, M.J.; Hoffmann, P.R. Dietary Selenium Modulates Activation and Differentiation of CD4+ T Cells in Mice through a Mechanism Involving Cellular Free Thiols. J. Nutr. 2010, 140, 1155.
  128. Zhang, J.; Taylor, E.W.; Bennett, K.; Saad, R.; Rayman, M.P. Association between Regional Selenium Status and Reported Outcome of COVID-19 Cases in China. Am. J. Clin. Nutr. 2020, 111, 1297–1299.
  129. Zhang, H.Y.; Zhang, A.R.; Lu, Q.B.; Zhang, X.A.; Zhang, Z.J.; Guan, X.G.; Che, T.L.; Yang, Y.; Li, H.; Liu, W.; et al. Association between Fatality Rate of COVID-19 and Selenium Deficiency in China. BMC Infect. Dis. 2021, 21, 452.
  130. Cheng, C.; Chen, S.Y.; Geng, J.; Zhu, P.Y.; Liang, R.N.; Yuan, M.Z.; Wang, B.; Jin, Y.F.; Zhang, R.G.; Zhang, W.D.; et al. Preliminary Analysis on COVID-19 Case Spectrum and Spread Intensity in Different Provinces in China except Hubei Province. Zhonghua Liu Xing Bing Xue Za Zhi 2020, 41, 1601–1605.
  131. Zeng, H.-L.; Zhang, B.; Wang, X.; Yang, Q.; Cheng, L. Urinary Trace Elements in Association with Disease Severity and Outcome in Patients with COVID-19. Environ. Res. 2021, 194, 110670.
  132. Im, J.H.; Je, Y.S.; Baek, J.; Chung, M.-H.; Kwon, H.Y.; Lee, J.-S. Nutritional Status of Patients with COVID-19. Int. J. Infect. Dis. 2020, 100, 390.
  133. Majeed, M.; Nagabhushanam, K.; Gowda, S.; Mundkur, L. An Exploratory Study of Selenium Status in Healthy Individuals and in Patients with COVID-19 in a South Indian Population: The Case for Adequate Selenium Status. Nutrition 2021, 82, 111053.
  134. Younesian, O.; Khodabakhshi, B.; Abdolahi, N.; Norouzi, A.; Behnampour, N.; Hosseinzadeh, S.; Alarzi, S.S.H.; Joshaghani, H. Decreased Serum Selenium Levels of COVID-19 Patients in Comparison with Healthy Individuals. Biol. Trace Elem. Res. 2021, 1–6.
  135. Skalny, A.; Timashev, P.S.; Aschner, M.; Aaseth, J.; Chernova, L.N.; Belyaev, V.E.; Grabeklis, A.R.; Notova, S.V.; Lobinski, R.; Tsatsakis, A.; et al. Serum Zinc, Copper, and Other Biometals Are Associated with COVID-19 Severity Markers. Metabolites 2021, 11, 244.
  136. Moghaddam, A.; Heller, R.A.; Sun, Q.; Seelig, J.; Cherkezov, A.; Seibert, L.; Hackler, J.; Seemann, P.; Diegmann, J.; Pilz, M.; et al. Selenium Deficiency Is Associated with Mortality Risk from COVID-19. Nutrients 2020, 12, 2098.
  137. Pincemail, J.; Cavalier, E.; Charlier, C.; Cheramy–Bien, J.-P.; Brevers, E.; Courtois, A.; Fadeur, M.; Meziane, S.; Goff, C.L.; Misset, B.; et al. Oxidative Stress Status in COVID-19 Patients Hospitalized in Intensive Care Unit for Severe Pneumonia. A Pilot Study. Antioxidants 2021, 10, 257.
  138. Polonikov, A. Endogenous Deficiency of Glutathione as the Most Likely Cause of Serious and Death in COVID-19 Patients. ACS Infect. Dis. 2020, 6, 1558–1562.
  139. Horowitz, R.I.; Freeman, P.R.; Bruzzese, J. Efficacy of Glutathione Therapy in Relieving Dyspnea Associated with COVID-19 Pneumonia: A Report of 2 Cases. Respir. Med. Case Rep. 2020, 30, 101063.
  140. Kieliszek, M.; Lipinski, B. Selenium Supplementation in the Prevention of Coronavirus Infections (COVID-19). Med. Hypotheses 2020, 143, 109878.
  141. Berggren, M.; Gallegos, A.; Gasdaska, J.; Powis, G. Cellular Thioredoxin Reductase Activity Is Regulated by Selenium. Anticancer Res. 1997, 17, 3377–3380.
  142. Broman, L.M.; Bernardson, A.; Bursell, K.; Wernerman, J.; Fläring, U.; Tjäder, I. Serum Selenium in Critically Ill Patients: Profile and Supplementation in a Depleted Region. Acta Anaesthesiol. Scand. 2020, 64, 803–809.
  143. Baker, R.D.; Baker, S.S.; Rao, R. Selenium Deficiency in Tissue Culture: Implications for Oxidative Metabolism. J. Pediatr. Gastroenterol. Nutr. 1998, 27, 387–392.
  144. Kanth Manne, B.; Denorme, F.; Middleton, E.A.; Portier, I.; Rowley, J.W.; Stubben, C.; Petrey, A.C.; Tolley, N.D.; Guo, L.; Cody, M.; et al. Platelet Gene Expression and Function in Patients with COVID-19. Blood 2020, 136, 1317.
  145. Ersöz, G.; Yakaryilmaz, A.; Turan, B. Effect of Sodium Selenite Treatment on Platelet Aggregation of Streptozotocin-Induced Diabetic Rats. Thromb Res. 2003, 111, 363–367.
  146. Varlamova, E.G.; Turovsky, E.A.; Blinova, E.V. Therapeutic Potential and Main Methods of Obtaining Selenium Nanoparticles. Int. J. Mol. Sci. 2021, 22, 10808.
  147. Huang, B.; Zhang, J.; Hou, J.; Chen, C. Free Radical Scavenging Efficiency of Nano-Se in Vitro. Free Radic. Biol. Med. 2003, 35, 805–813.
  148. Khurana, A.; Tekula, S.; Saifi, M.A.; Venkatesh, P.; Godugu, C. Therapeutic Applications of Selenium Nanoparticles. Biomed. Pharmacother. 2019, 111, 802–812.
  149. Hosnedlova, B.; Kepinska, M.; Skalickova, S.; Fernandez, C.; Ruttkay-Nedecky, B.; Peng, Q.; Baron, M.; Melcova, M.; Opatrilova, R.; Zidkova, J.; et al. Nano-Selenium and Its Nanomedicine Applications: A Critical Review. Int. J. Nanomedicine 2018, 13, 2107.
  150. Turovsky, E.A.; Varlamova, E.G. Mechanism of Ca2+-Dependent pro-Apoptotic Action of Selenium Nanoparticles, Mediated by Activation of Cx43 Hemichannels. Biology 2021, 10, 743.
  151. He, L.; Zhao, J.; Wang, L.; Liu, Q.; Fan, Y.; Li, B.; Yu, Y.L.; Chen, C.; Li, Y.F. Using Nano-Selenium to Combat Coronavirus Disease 2019 (COVID-19)? Nano Today 2021, 36, 101037.
  152. Qiao, Q.; Liu, X.; Yang, T.; Cui, K.; Kong, L.; Yang, C.; Zhang, Z. Nanomedicine for Acute Respiratory Distress Syndrome: The Latest Application, Targeting Strategy, and Rational Design. Acta Pharm. Sin. B 2021, 11, 3060.
  153. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of Mpro from SARS-CoV-2 and Discovery of Its Inhibitors. Nature 2020, 582, 289–293.
  154. Azad, G.K.; Tomar, R.S. Ebselen, a Promising Antioxidant Drug: Mechanisms of Action and Targets of Biological Pathways. Mol. Biol. Rep. 2014, 41, 4865–4879.
More
Video Production Service