Next Article in Journal
Detraining Effects on Muscle Quality in Older Men with Osteosarcopenia. Follow-Up of the Randomized Controlled Franconian Osteopenia and Sarcopenia Trial (FrOST)
Next Article in Special Issue
Dietary Fibre Modulates the Gut Microbiota
Previous Article in Journal
A Comparison of Sugar Intake between Individuals with High and Low Trait Anxiety: Results from the NutriNet-Santé Study
Previous Article in Special Issue
Healthy Lifestyle Intervention and Weight Loss Improve Cardiovascular Dysfunction in Children with Obesity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dietary Selenium Regulates microRNAs in Metabolic Disease: Recent Progress

1
Beijing Advanced Innovation Center for Food Nutrition and Human Health, Department of Nutrition and Health, China Agricultural University, Beijing 100083, China
2
Key Laboratory of Precision Nutrition and Food Quality, Department of Nutrition and Health, Ministry of Education, China Agricultural University, Beijing 100083, China
3
College of Veterinary Medicine, China Agricultural University, Beijing 100083, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this paper.
Nutrients 2021, 13(5), 1527; https://doi.org/10.3390/nu13051527
Submission received: 5 April 2021 / Revised: 24 April 2021 / Accepted: 27 April 2021 / Published: 1 May 2021
(This article belongs to the Special Issue Nutrition and Diet for Metabolic Health)

Abstract

:
Selenium (Se) is an essential element for the maintenance of a healthy physiological state. However, due to environmental and dietary factors and the narrow safety range of Se, diseases caused by Se deficiency or excess have gained considerable traction in recent years. In particular, links have been identified between low Se status, cognitive decline, immune disorders, and increased mortality, whereas excess Se increases metabolic risk. Considerable evidence has suggested microRNAs (miRNAs) regulate interactions between the environment (including the diet) and genes, and play important roles in several diseases, including cancer. MiRNAs target messenger RNAs to induce changes in proteins including selenoprotein expression, ultimately generating disease. While a plethora of data exists on the epigenetic regulation of other dietary factors, nutrient Se epigenetics and especially miRNA regulated mechanisms remain unclear. Thus, this review mainly focuses on Se metabolism, pathogenic mechanisms, and miRNAs as key regulatory factors in Se-related diseases. Finally, we attempt to clarify the regulatory mechanisms underpinning Se, miRNAs, selenoproteins, and Se-related diseases.

1. Introduction

Selenium (Se) is a metalloid element that fulfills important physiological functions within the necessary dose, but human health is also vulnerable to selenium deficiency or selenium excess [1]. Daily food can meet people’s demand for selenium, comprising a balanced selection of meat and plant products [2]. In the natural environment, rock and soil composition are believed to determine Se distribution characteristics [3]. Populations in the United States, Mexico, Colombia, India, and Iceland easily attain their recommended daily Se dose. However, populations in Northern Europe, Australia, New Zealand, and China have poor Se-containing soils, potentially leading to Se deficiencies [4].
Recently, microRNAs (miRNAs) and their role in Se-related inflammation and diseases have attracted considerable attention [5,6,7,8]. MiRNAs are non-coding endogenous single-stranded RNA molecules that consist of 20–23 nucleotides [9]. They play central roles in cell differentiation, proliferation, and survival by binding to complementary target messenger RNA (mRNA), leading to mRNA translation, inhibition, and/or degradation [10]. Therefore, they can be regarded as key gene expression regulators that can control physiological and pathological processes, including the development of cancer [9]. Studies have confirmed miRNA dysregulation is causal in many cancers [11,12,13,14,15], with miRNAs acting as tumor suppressors or oncogenes. Similarly, miRNA mimics and molecules targeting miRNAs have shown promise in preclinical studies [12].
Keshan disease (KD) and Kashin–Beck disease (KBD) are related to the Se deficiency, but the actual mechanisms that are behind these diseases are still not precisely understood [16]. Microarray and proteomics analysis revealed the genes and pathways that may be involved in these diseases. Nineteen Se- and three zinc-associated proteins were identified among 105 differentially-expressed proteins. The proteins involved in hypoxia-inducible factor-1α and apoptosis pathways may play significant roles in the pathogenesis of KD [17]. There are numerous functional pathways and cellular systems associated with the differentially expressed genes and proteins; the TCA Cycle II (Eukaryotic) pathway and NADP repair pathway may also participate in the pathogenesis of KD [18]. Thirty-four nutrients associated with differentially expressed genes and ten significant pathways have been identified for juvenile KBD, which are mainly related to metabolism, cell apoptosis, extracellular matrix, and other functions which consist of pathological changes of KBD [19]. One hundred and twenty-four miRNAs had lower expression levels in the subchondral bone sampled from KBD patients showed by miRNA array profiling [20]. These differential genes or proteins may become new targets for studying microRNAs in the two diseases.
In this review, we discussed recent studies related to miRNA changes in cardiovascular diseases, cancer, and other metabolic diseases caused by Se deficiency or excess. The prospect of miRNA as a potential target for selenium-related diseases is also pointed out in the article.

2. Selenium Uptake and Metabolism

Se levels in a given food product does not mean an organism will derive its correct Se quota—instead, this depends on the bioavailability, bioaccessibility, and/or bioactivity of a given Se compound [2]. Human dietary Se forms mainly include organic and inorganic Se. These forms are typically absorbed without any regulatory processes and have a high bioavailability in the body [2,21]. Se is primarily absorbed in the duodenum and caecum after active transport via a sodium pump, but this process is different depending on the chemical form [22]. After intestinal absorption, different Se forms enter the bloodstream and are transported into the liver via the portal vein, where they are metabolized, transported, and distributed to different tissues [23] (Figure 1, adapted from [24,25,26,27] and drawn with https://app.biorender.com/ access on 1 March 2021). In normal diets, tissue Se concentrations in the body range from the highest to the lowest in the following organs: Kidney, liver, spleen, pancreas, heart, brain, lung, bone, and skeletal muscle [28].
SeMet accounts for 90% of total Se in plants. Some SeMet is randomly incorporated into proteins at methionine positions [29], whereas other SeMet quantities are metabolized to selenocysteine (SeCys) via methionine cycle and transsulfuration pathways in the liver [26]. Furthermore, Se-methylselenocysteine and γ-glutamyl-Se-methylselenocysteine (believed to exert anticancer effects) are also found in plants, such as garlic, onions, and broccoli, and they are metabolized to methyl selenol [25,30]. SeCys occurs at much lower levels than SeMet in plants. When SeCys is absorbed, free SeCys does not appear to generate concentrations for efficient attachment to cysteine transfer RNA (tRNA). But once incorporated into proteins, SeCys predisposes these proteins to degradation processes [31]. SeCys is the main Se source in animal products, however, highly reactive free SeCys is maintained at very low concentrations in tissues [32]. Inorganic Se is the main form of Se supplementation as it promotes selenoprotein biosynthesis [26]. Selenate must be reduced to selenite before further metabolism. Then, interactions with the tripeptide and glutathione ensure this selenite is reduced to selenide (H2Se), which is a central gateway molecule for Se utilization and excretion [25]. Furthermore, all seleno-compounds must be metabolized to selenide for incorporation into selenoproteins [29]. After initial Se metabolism, H2Se is converted to selenophosphate, which is used to convert phosphoseryl-tRNA[Ser]Sec to SeCys-tRNA[Ser]Sec. Then the SeCys-tRNA[Ser]Sec reads the UGA codon and integrates SeCys into the amino acid sequence to form a selenoprotein [33]. Selenoprotein P, which is mainly produced in the liver, transports Se from the liver to extrahepatic tissues and organs, where it is metabolized to prevent oxidative damage [26]. Methyl selenol is demethylated to H2Se in the equilibrium reaction, where it and its precursors (SeMet and CH3SeCys) may be used as Se sources for selenoprotein synthesis [34]. The oxidation of excess H2Se leads to superoxides and other active oxygen species, often with toxic effects [35]. Se excess detoxification occurs via sequential methylation into dimethyl selenide, and is excreted via the breath, whereas Se-sugars and trimethyl selenonium are excreted in the urine [27]. Although all Se forms are excreted from the body at some stage, only Se-sugars are bioavailable. It is not only an excretion metabolite of Se, but also may transport selenium from liver cells to other cells in the body [36].

3. Selenium Related Pathogenic Mechanisms and Diseases

In general, when Se plasma levels are less than 85 µg/L, Se deficiency becomes evident in the body [37]. Se deficiency is caused by poor Se dietary intake, and may be induced or aggravated by nutritional, chemical, and infectious stresses. Several Se deficiency animal diseases are related to the co-existing vitamin E deficiencies [38]. Se deficiency causes heart disease (e.g., cardiomyopathy, arrhythmias), infertility, neuronal or neuromuscular diseases, and increased susceptibility to cancer, infection, and heavy metal toxicity [39,40,41,42,43]. The maximum harmless Se concentration is less than 400 μg per day in adults [44]. Excess dietary Se causes adverse effects (selenosis), including acute food poisoning symptoms, such as vomiting, nausea, and diarrhea, as well as chronic toxicity manifested by the hair and nail brittleness and loss, gastrointestinal disturbances, infertility, and nervous system abnormalities [45]. In addition, when Se is excessive, the toxicity of inorganic Se is much lower than that of organic selenomethionine (SeMet) [38].
Both Se excess and deficiency lead to Se-related disease. A recent epidemiological analysis showed that taking 300 µg/d selenium for 5 consecutive years increased all-cause mortality after 10 years in countries with moderately low selenium levels [46]. Dietary Se functions mainly depend on the selenoprotein form that exerts biological effects in the body. Almost all tissues are affected by changes in Se status or selenoprotein expression [47,48,49]. For instance, embryonic lethality caused by trsp gene deletion, which encodes Sec-tRNA for translation, also reflects the importance of selenoproteins to the body [50]. Currently, 25 genes in the human genome have been identified as encoding selenoproteins, with most exhibiting antioxidant activities [51]. Other specific processes include the biosynthesis of deoxyribonucleoside triphosphates for DNA, the reduction of oxidized proteins and membranes, redox regulation of transcription factors, apoptosis regulation, immunomodulation, thyroid hormone regulation, Se transport and storage, protein folding, and the degradation of misfolded proteins in the endoplasmic reticulum [27]. Therefore, it appears the physiological and pathological changes or diseases caused by Se deficiency are primarily mediated by a selenoprotein imbalance [28,52,53,54,55]. Correspondingly, excess Se generates toxicity via several mechanisms [56]: (1) CH3Se− formation, which either enters a redox cycle and generates superoxide and oxidative stress, or generates free radicals that bind to and inhibit key enzymes and proteins, (2) SeCys excess, which inhibits Se methylation metabolism, results in hydrogen selenide (intermediate metabolite) accumulation eventually leading to hepatotoxicity and other Se-related adverse effects, (3) excess Se analogs of sulfur-containing enzymes and structural proteins also play roles in avian teratogenesis. Equally, aquatic organisms exposed to high Se doses are at risk of organ damage and genomic mutations, which potentially pose a threat to human food chains [57,58].

4. Current Progress in Nutrient Regulation of miRNAs

4.1. miRNAs

MiRNAs are derived from intergenic or intragenic (exon and intron) genomic regions [59]. They are usually transcribed by RNA polymerase II from miRNA genes, first forming a ‘primary miRNA transcript’ (pri-miRNA). This transcript is cleaved by a microprocessor complex, comprising the double-stranded RNase III enzyme, DROSHA, and its essential cofactor, the DiGeorge syndrome critical region 8 (DGCR8) protein, generating a short sequence, the ‘miRNA precursor’ (pre-miRNA), which displays a hairpin-like secondary structure [60]. The pre-miRNA is exported to the cytoplasm and processed by DICER, a ribonuclease III enzyme that produces the mature miRNA for final incorporation into an RNA-induced silencing complex (RISC) [61]. Under most conditions, mature RISC represses gene expression post-transcriptionally by binding to 3′ untranslated regions of specific mRNAs, and mediates degradation, destabilization, or translational inhibition, based on target sequence complementarity [62]. MiRNAs are abundant in all cells, are found in extracellular body fluids (e.g., serum, plasma, saliva, and urine), and are implicated in several pathological conditions [63,64]. During some biological processes, miRNAs regulate protein levels of key regulatory factors, or serve as switches to govern gene expression [65]. Post-transcriptional regulation of miRNA can improve the compliance, accuracy, and sensitivity of gene expression regulation [66]. Due to its small molecular size, each miRNA potentially targets hundreds of mRNA molecules [67]. Similarly, each mRNA may be targeted by multiple miRNAs to form complexes and multifaceted regulatory networks [68]. Moreover, miRNAs also regulate DNA methylation and histone modification [69]. It was previously reported that >60% of human coding genes are regulated by miRNAs [70], and >2800 mature miRNA sequences are described in the miRBase 22 repository (http://www.mirbase.org/ access on 5 February 2021). In addition, miRNAs are not only endogenously synthesized, but may be derived from the diet (e.g., milk and plants) [71].

4.2. Regulation Mechanisms of Vitamins and Minerals on miRNAs

Abnormal miRNA expression is driven by genetic and epigenetic factors, which are implicated in cancer occurrence and development, and this may be reversed by a variety of dietary components [72]. Considerable evidence has indicated that dietary factors modify miRNA expression and mRNA targets during cancer, including apoptosis, cell cycle regulation, differentiation, inflammation, angiogenesis and metastasis, and stress response pathways [59], however, this topic is outside the remit of our review. Vitamins and mineral nutrients induce miRNA expression by activating transcription factors/response elements, thereby changing gene expression by inducing mRNA degradation or inhibiting translation [73]. Furthermore, vitamins and minerals alter the function of classical epigenetic mechanisms, including DNA methyltransferases (DNMT) and histone-modifying enzymes (e.g., histone deacetylases (HDAC) and histone acetyltransferases), such enzyme regulation regulates gene expression, including miRNAs [73]. Minerals, such as magnesium ions, are located in the small RNA binding domain of the argonaute protein. RISC is a magnesium-dependent protein; therefore, magnesium ions facilitate the binding of miRNA and the argonaute protein, and also help to cleave miRNA targets and regulate argonaute stability [74]. In contrast, DGCR8 forms a highly stable and active complex with heme, therefore when heme is reduced to a ferrous state, the pri-miRNA processing abilities of the DGCR8 complex disappear [75]. In addition, a recent study observed that aluminum sulfate up-regulated miR-125b and miR-146a expression via NF-κB-dependent mechanisms suggesting these miRNAs were involved in astrocyte proliferation and inflammation [76]. It was recently proposed that the environment, including dietary factors, may induce epigenetic changes via three possible mechanisms: (1) Activation/inhibition of chromatin machinery, (2) activation of nuclear receptors by ligands, (3) membrane receptor signal transduction cascades [77], and (4) the involvement of key epigenetic regulatory enzymes including DNMT, DNA desorption methylase, histone acetylase, and HDAC [78]. MiRNAs are also regulated by DNA methylation status in cells—up to 33% of dysregulated miRNA loci exhibit consistent DNA methylation and H3K9 acetylation patterns [79].

4.3. Role of Mammalian Target of Rapamycin (mTOR) in Nutrient Regulation of miRNAs

mTOR is central to the nutrient-sensing signaling network [80]. Its activity is regulated by multiple nutrients, such as amino acids and glucose, influencing muscle cell proliferation, differentiation, autophagy, and metabolism [81]. Following cellular nutrient depletion, in particular leucine starvation, mTORC1 becomes inhibited, thereby inhibiting translation initiation and elongation [82]. Studies have reported that miR-1 expression was regulated by mTOR, mainly via protein stability generated by the myogenic transcription factor, MyoD 18, which was located in the upstream enhancer of multiple myogenic miRNAs [83]. Similarly, miR-133 and miR-206 expression patterns were similarly regulated by MyoD 20 [84]. However, the exact mechanism by which mTORC1 regulates the stability of the myogenic transcription factors remains unclear. In addition, mTOR activation also significantly down-regulated miRNA biogenesis by up-regulating Mdm2, which is an important E3 ligase for the ubiquitination of the miRNA processing enzyme, DROSHA 21 [85]. Also, nutrient starvation can induce autophagy by inhibiting mTORC1, to provide an important substrate source for extracellular energy production [86]. Therefore, the nutrient-mTOR-miRNA pathway appears to rely on the typical nutrient-sensing mTORC1 pathway to regulate autophagy [87]. Importantly, the specific regulatory mechanisms underpinning downstream mRNA-mediated alterations have not been fully elucidated.

4.4. miRNAs Mediated by Se Status Are Implicated in Disease Development and Progression

In this section, recent developments on miRNA involvement in Se-related diseases are described, including miRNA changes when Se is in deficient, moderate, or excessive status, regulatory miRNA targets, and miRNA roles in Se antioxidant damage or Se-related diseases (Table 1).
Se supplementation reportedly changed miRNA profiles in the intestinal cell line, Caco-2, differentially down-regulating 12 miRNAs under nutrient deficiency conditions [77]. The miRNAs most affected were miR-185, miR-625, miR-203, and miR-429, whereas pathway analyses identified arachidonic acid metabolism, glutathione metabolism, oxidative stress, and mitochondrial respiration as Se-sensitive pathways [88]. In a Se deficiency rat model, five miRNAs from harvested heart tissue (miR-374, miR-16, miR-199a-5p, miR-195 and miR-30e were up-regulated > 5-fold in the Se nutrient deficiency group, when compared with the Se-supplemented group, whereas three miRNAs were down-regulated (miR-3571, miR-675, and miR-450a [78]. Up-regulated miRNAs were involved in signal transduction, cell differentiation, and stress responses, suggesting roles in cardiac function and regulation [89]. A Se pro-longevity mechanism study reported that several miRNAs were altered in response to dietary Se in the mouse liver [3]. Expression levels of 38 miRNAs were altered by Se deficiency compared with Se sufficiency, and the study showed that selenoprotein regulation by miRNAs was not a direct effect [3]. The role of glutathione peroxidase regulation and related miRNAs has also been reported [90,91]. In an intervention study in elderly males given Se and coenzyme Q10 supplements for four years, significant expression differences were observed in >100 miRNAs, with up to 4-fold differences in combined Se and coenzyme Q10 supplementation experiments [92]. Such changes may contribute to underlying clinical mechanisms. Early reports indicated that cardiovascular mortality was reduced, cardiac function improved, and inflammation and oxidative stress indications decreased after Se intervention [92]. Se decreased inflammation by increasing miR-146a expression, decreasing mmu-miR-155, TLR2/6, NF-κB, and MAPK signaling pathway expression in mammary tissue from infected animals, and mammary epithelial cells [93,94]. Although these studies investigated miRNA-mediated Se deficiency or Se antioxidant damage, data for miRNAs implicated in Se excess are limited. When Se is in excess, it potentially increases the risk of metabolic syndrome [95,96].
Apart from the aforementioned transcription factors being implicated in miRNA regulation by Se, it remains unclear how Se precisely regulates miRNAs. Thus, similar to mechanisms involved in nutrition-gene interactions, it is reasonable to speculate that selenium regulates the expression of miRNAs by potentially affecting the epigenetic regulation mechanisms, including DNA methylation and histone modification. Se supplementation may modify global DNA methylation and specific gene regions, possibly via DNMT inhibition [97]. Additionally, dietary Se deficiency may decrease DNA methylation by enhancing trans-sulfonation pathways [98]. Se also alters histone modification via HDAC inhibition of the Se metabolism products, seleno-α-keto acids [97]. Taken together, the current evidence indicates that different DNA hypomethylation mechanisms occur at different Se levels, including (1) the redirection of homocysteine towards transsulfuration pathways and glutathione synthesis during Se deficiency, (2) excess Se competes with S-adenosylmethionine to use the methyl group required for selenium metabolism—consequently, S-adenosylmethionine levels are reduced for DNMT and methylation processes are similarly inhibited and (3) Se affects specific tumor suppressor gene methylation mechanisms, possibly in a sex-dependent manner. Importantly cancer phenotypes are often characterized by the altered methylation of selenoprotein-encoding genes, mainly glutathione peroxidase 3 [99].
Table 1. MiRNAs are regulated by selenium (deficiency, moderate, and excess).
Table 1. MiRNAs are regulated by selenium (deficiency, moderate, and excess).
Se StatusMicro RNATargetObserved EffectNoteReference
Se deficiency↑miR-181a-5P↓SBP2↓GPX1, GPX4, and SELENOS levelsIn C28/I2 human juvenile chondrocytes and DA rats[100]
↑Gga-let-7f-3p↓SELENOK↑Oxidative stress, ERS, and apoptosisIn chicken myoblasts and muscle[101]
↑miR-200a-5pTXNRD2, TXNRD3, SELENON, SELENOT, SELENOF and SELENOP↑Glucose metabolism disorder, cardiomyocyte hypertrophyIn chicken cardiomyocytes [102]
↓RNF11↑Oxidative stress and myocardial necroptosisIn chicken cardiac tissue and cardiomyocytes[103]
↑miR-138-5p↓SELENOM↑Apoptosis, oxidative stress, mitochondrial fissionIn chicken chondrocytes[104]
↑miR-544a↓SELENOKInterferes with SELENOK translationIn HepG2 and HuH-7 human hepatocarcinoma cells[105]
↑miR-196-5p↓NFκBIA (IκB-α)↑LPS-induced oxidative stress and inflammation, respiratory mucosal immune dysfunctionIn chicken trachea[106]
↑miR-193b-3p↓MAML1↑Hepatocyte apoptosisIn the liver tissues and primary hepatocytes from broilers[107]
↑miR-33-3p↓ADAM10↑Cell cycle arrest and apoptosisIn vivo and in vitro in the chicken kidney[108]
↓E4F1↑Oxidative stress, ERS, and apoptosisIn vein endothelial cells from broilers[109]
↑miR-328↓ATP2A2↑Intracellular Ca2+ and cell apoptosisIn H9c2 rat cardiac myoblasts[110]
↑miR-215-5p↓CTCF↑Mitochondrial biosynthesis imbalance, defects in myocardial developmentIn heart tissue and primary cardiomyocytes from chickens[111]
↓PI3K/AKT/TOR↑ROS, Myocardial autophagyIn cardiomyocytes of chicken[112]
↑miR-1594↓TNNT2↑Ca2+In heart and primary cardiomyocytes from chickens[113]
↑miR-2954↓PI3K↑Autophagy and apoptosisIn heart and primary cardiomyocytes from chickens [114]
↑miR-16-5p↓PI3K/AKT↑NecroptosisIn tracheal tissues and tracheal epithelial cells of chicken [115]
↑miR-128-1-5p↓CADM1↑Tight junction structural damage and cell cycle arrestedIn vein tissues and vein endothelial cells from broilers [116]
↑ miR-374, miR-16, miR-199a-5p, miR-195 and miR-30e
↓ miR-3571, miR-675a and miR-450a
↑ Wnt/β--catenin↑ Cardiac dysfunctionIn rat heart[89]
↓miR--185↑GPX2, SEPHS2↑Altered expression of 12 miRNA and 50 genesIn Caco-2 human intestinal cells[88]
↓miR-29a-3p↑TNFR1Altered expression of selenoprotein genes, ↑necrotic cells In the pig brain and IPEC-J2 pig intestinal epithelial cells[117]
↓miR-155↑TNFRSF1B↑Oxidative stress-induced apoptosisIn splenic cells and spleen of broilers[118]
↓miR-146a↑MAPKs↑ROS-induced inflammationIn the head kidney of carp[119]
↓miR-7↓SELENOPBoth are potential biomarkers of HCCIn HCC patients and HepG2 human hepatocarcinoma cells [120]
Se moderate↑miR-146a↓TLR2, TLR6, NF-κB and MAPKS. aureus-infected mastitis In mammary tissues and mammary epithelial cells from mouse[93]
↑miR-125a and miR-125b↓Bak and caspase-3↓Cd-induced apoptosisIn LLC-PK1 porcine renal epithelial cells[121]
↑ miR-29b-3p, miR-30e-5p and miR-19a-3p
↓ miR-199a-5p, miR-130a-3p and miR-191-5p
——↓ Risk of heart failureIn healthy elderly males[92]
↓mmu-miR-155↓TNF-α, IL-1β, IL-10, TLR2, NF-κB and MAPKs↓ S. aureus-infected mastitisIn mammary tissues and mammary epithelial cells from mouse[94]
↓miR-224ID1↓Pb-induced oxidative damage and restoring thyroid hormone disequilibriumIn thyroid tissues of male rats[122]
↓miR-16-5p↑PiK3R1 and IGF1R↓Pb-induced neutrophil apoptosisFrom chicken peripheral blood[123]
↓miR-216a↑PI3K/AKT↓Cd-triggered necrosis and apoptosisIn the splenic lymphocytes of common carp[124]
Se excess↑miR-122-5p↑BMI, SBP and DBP↑Risk of MetSIn male adults[95]
↑miR-454-3p and miR-584-5p
↓miR-375
A link between Se intake, vitamin D metabolism, and calcium homeostasis↑miR-375 as a potential biomarker of MetSIn obese women[96]
miRNAs, genes, or proteins down-regulated/inhibited (↓) or up-regulated/activated (↑). SBP2, SECIS binding protein 2; SELENOK, selenoprotein K; TXNRD2, thioredoxin reductase 2; TXNRD3, thioredoxin reductase 3; SELENON, selenoprotein N; SELENOT, selenoprotein T; SELENOF, selenoprotein F; SELENOP, selenoprotein P; RNF 11, ring finger protein 11; SELENOM, selenoprotein M; NFκBIA (IκB-α), IkappaB-alpha; MAML1, mastermind-like protein 1; ADAM 10, adisintegrin and metalloprotease domain 10; E4F1, E4F transcription factor 1; ATP2A2, sarcoplasmic/endoplasmic reticulum calcium ATPase 2; CTCF, CCCTC-binding factor; TNNT2, Troponin T Type 2; CADM1, cell adhesion molecule 1; GPX2, glutathione peroxidase 2; SEPHS2, selenophosphatesynthase 2; TNFR1, TNF receptor superfamily member 1A; TNFRSF1B, TNF receptor superfamily member 1B; TLR2, toll-like receptor 2; TLR6, toll-like receptor 6; ID1, Inhibitor of DNA binding 1; PIK3R1, phosphoinositide-3-kinase regulatory subunit 1; IGF1R, type 1 insulin-like growth factor receptor; BMI, body mass index; SBP, systolic pressure, and DBP, diastolic pressure.

5. Conclusions

These data indicate that miRNAs can interfere with many proteins, including selenoproteins, promoting Se-related diseases. Some miRNAs have shown clear associations with cardiomyopathy pathology, inflammation, and apoptosis. Actually, the effects of Se on epigenetic mechanisms, especially miRNAs, are poorly described and represent an interesting field of study.
As miRNAs are involved in the transcriptional regulation of genes, thus, acting on the maintenance of the functionality of numerous physiological processes. Se benefits may be mediated through its function as a component of small Se-containing metabolites or via its role in selenoprotein activity/regulation [125]. To clarify the mechanism of microRNA regulating Se-related diseases, we should focus not only on the influence of Se status on miRNAs and the regulation of miRNAs on selenoproteins and other key proteins, but also on the effect of miRNAs on Se status in target tissues. It would be interesting to see whether miRNAs interfering with Se incorporation (e.g., targeting SCLY, TRSP) would have a profound effect on disease.

Author Contributions

Conceptualization, X.H., Y.-L.D. and J.-Q.H.; investigation, X.H. and T.L.; visualization, W.X. and X.Z.; writing—original draft preparation, X.H.; writing—review and editing, X.H., Y.-L.D. and J.-Q.H.; supervision, P.-J.W.; funding acquisition, Y.-L.D. and J.-Q.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China 32002216, Chinese Universities Scientific Fund (2021TC048), and the 111 Project from the Education Ministry of China (No. B18053).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

GPX1glutathione peroxidase-1
GPX3glutathione peroxidase-3
GPX4glutathione peroxidase-4
HCChepatocellular carcinoma
miRNAmicroRNA
Seselenium
SeMetselenomethionine
SeCysselenocysteine
H2Seselenide
HSePO32−selenophosphate
SELENOPselenoprotein P
oncomiRsmiRNAs acting as tumour suppressors or oncogenes
dNTPsdeoxyribonucleoside triphosphates
pri-miRNAprimary miRNA transcript
DGCR8DiGeorge syndrome critical region 8
pre-miRNAmiRNA precursor
mRNAmessenger RNA
DNMTDNA methyltransferase
HATshistone acetylase
HDAChistone deacetylase
mTORMammalian target of rapamycin

References

  1. Kieliszek, M.; Błazejak, S. Selenium: Significance, and outlook for supplementation. Nutrition 2013, 29, 713–718. [Google Scholar] [CrossRef]
  2. Bodnar, M.; Szczyglowska, M.; Konieczka, P.; Namiesnik, J. Methods of selenium supplementation: Bioavailability and determination of selenium compounds. Crit. Rev. Food Sci. Nutr. 2016, 56, 36–55. [Google Scholar] [CrossRef]
  3. Yim, S.H.; Clish, C.B.; Gladyshev, V.N. Selenium deficiency is associated with pro-longevity mechanisms. Cell Rep. 2019, 27, 2785–2797.e3. [Google Scholar] [CrossRef] [Green Version]
  4. Bitterli, C.; Bañuelos, G.S.; Schulin, R. Use of transfer factors to characterize uptake of selenium by plants. J. Geochem. Explor. 2010, 107, 206–216. [Google Scholar] [CrossRef]
  5. Liang, Z.Z.; Zhu, R.M.; Li, Y.L.; Jiang, H.M.; Li, R.B.; Wang, Q.; Tang, L.Y.; Ren, Z.F. Differential epigenetic profiles induced by sodium selenite in breast cancer cells. J. Trace Elem. Med. Biol. 2021, 64, 126677. [Google Scholar] [CrossRef]
  6. Li, Y.; He, M.; Li, J.; Yao, Y.; Zhu, L.; Wu, B. Regulatory protein genes and microRNAs in response to selenium stimuli in Pueraria lobata (Willd.) Ohwi. Metallomics 2021, 13, mfaa004. [Google Scholar] [CrossRef] [PubMed]
  7. Zichan, H.; Linfei, J.; Jinliang, W.; Zhiqiang, S.; Yimei, C.; Shu, L. MicroRNA-294 regulates apoptosis of the porcine cerebellum caused by selenium deficiency via targeting iNOS. Biol. Trace Elem. Res. 2021. [Google Scholar] [CrossRef]
  8. Zhirong, Z.; Qiaojian, Z.; Chunjing, X.; Shengchen, W.; Jiahe, L.; Zhaoyi, L.; Shu, L. Methionine selenium antagonizes LPS-induced necroptosis in the chicken liver via the miR-155/TRAF3/MAPK axis. J. Cell. Physiol. 2020, 236, 4024–4035. [Google Scholar] [CrossRef] [PubMed]
  9. Winter, J.; Jung, S.; Keller, S.; Gregory, R.I.; Diederichs, S. Many roads to maturity: microRNA biogenesis pathways and their regulation. Nat. Cell Biol. 2009, 11, 228–234. [Google Scholar] [CrossRef] [PubMed]
  10. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  11. Vishnoi, A.; Rani, S. MiRNA biogenesis and regulation of diseases: An overview. In Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2017; Volume 1509, pp. 1–10. [Google Scholar]
  12. Rupaimoole, R.; Slack, F.J. MicroRNA therapeutics: Towards a new era for the management of cancer and other diseases. Nat. Rev. Drug Discov. 2017, 16, 203–221. [Google Scholar] [CrossRef]
  13. Lu, J.; Getz, G.; Miska, E.A.; Alvarez-Saavedra, E.; Lamb, J.; Peck, D.; Sweet-Cordero, A.; Ebert, B.L.; Mak, R.H.; Ferrando, A.A.; et al. MicroRNA expression profiles classify human cancers. Nature 2005, 435, 834–838. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, Z.; Shi, K.; Yang, S.; Liu, J.; Zhou, Q.; Wang, G.; Song, J.; Li, Z.; Zhang, Z.; Yuan, W. Effect of exosomal miRNA on cancer biology and clinical applications. Mol. Cancer 2018, 17, 147. [Google Scholar] [CrossRef]
  15. Peng, Y.; Croce, C.M. The role of microRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef] [Green Version]
  16. Lammi, M.J.; Qu, C. Selenium-Related transcriptional regulation of gene expression. Int. J. Mol. Sci. 2018, 19, 2665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Wang, S.; Lv, Y.; Wang, Y.; Du, P.; Tan, W.; Lammi, M.J.; Guo, X. Network Analysis of Se-and Zn-related Proteins in the Serum Proteomics Expression Profile of the Endemic Dilated Cardiomyopathy Keshan Disease. Biol. Trace Elem. Res. 2018, 183, 40–48. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, S.; Yan, R.; Wang, B.; Du, P.; Tan, W.; Lammi, M.J.; Guo, X. Prediction of co-expression genes and integrative analysis of gene microarray and proteomics profile of Keshan disease. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef] [Green Version]
  19. Ning, Y.; Wang, X.; Zhang, P.; Anatoly, S.V.; Prakash, N.T.; Li, C.; Zhou, R.; Lammi, M.; Zhang, F.; Guo, X. Imbalance of dietary nutrients and the associated differentially expressed genes and pathways may play important roles in juvenile Kashin-Beck disease. J. Trace Elem. Med. Biol. 2018, 50, 441–460. [Google Scholar] [CrossRef] [PubMed]
  20. Wu, W.; He, A.; Wen, Y.; Xiao, X.; Hao, J.; Zhang, F.; Guo, X. Comparison of microRNA expression profiles of Kashin-Beck disease, osteoarthritis and rheumatoid arthritis. Sci. Rep. 2017, 7, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Vindry, C.; Ohlmann, T.; Chavatte, L. Selenium metabolism, regulation, and sex differences in mammals. In Molecular and Integrative Toxicology; Springer: Cham, Switzerland, 2018; pp. 89–107. [Google Scholar]
  22. Mehdi, Y.; Hornick, J.L.; Istasse, L.; Dufrasne, I. Selenium in the environment, metabolism and involvement in body functions. Molecules 2013, 18, 3292–3311. [Google Scholar] [CrossRef] [Green Version]
  23. Ha, H.Y.; Alfulaij, N.; Berry, M.J.; Seale, L.A. From selenium absorption to selenoprotein degradation. Biol. Trace Elem. Res. 2019, 192, 26–37. [Google Scholar] [CrossRef]
  24. Wrobel, J.K.; Power, R.; Toborek, M. Biological activity of selenium: Revisited. IUBMB Life 2016, 68, 97–105. [Google Scholar] [CrossRef] [PubMed]
  25. Rayman, M.P.; Infante, H.G.; Sargent, M. Food-Chain selenium and human health: Spotlight on speciation. Br. J. Nutr. 2008, 100, 238–253. [Google Scholar] [CrossRef] [Green Version]
  26. Burk, R.F.; Hill, K.E. Regulation of selenium metabolism and transport. Annu. Rev. Nutr. 2015, 35, 109–134. [Google Scholar] [CrossRef] [PubMed]
  27. Roman, M.; Jitaru, P.; Barbante, C. Selenium biochemistry and its role for human health. Metallomics 2014, 6, 25–54. [Google Scholar] [CrossRef] [PubMed]
  28. Guillin, O.M.; Vindry, C.; Ohlmann, T.; Chavatte, L. Selenium, selenoproteins and viral infection. Nutrients 2019, 11, 2101. [Google Scholar] [CrossRef] [Green Version]
  29. Sonet, J.; Mosca, M.; Bierla, K.; Modzelewska, K.; Flis-Borsuk, A.; Suchocki, P.; Ksiazek, I.; Anuszewska, E.; Bulteau, A.L.; Szpunar, J.; et al. Selenized plant oil is an efficient source of selenium for selenoprotein biosynthesis in human cell lines. Nutrients 2019, 11, 1524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Rayman, M.P. Selenium and human health. Lancet 2012, 379, 1256–1268. [Google Scholar] [CrossRef]
  31. Sabbagh, M.; Van Hoewyk, D. Malformed selenoproteins are removed by the ubiquitin-proteasome pathway in stanleya pinnata. Plant Cell Physiol. 2012, 53, 555–564. [Google Scholar] [CrossRef] [Green Version]
  32. Suzuki, K.T. Metabolomics of selenium: Se metabolites based on speciation studies. J. Health Sci. 2005, 51, 107–114. [Google Scholar] [CrossRef] [Green Version]
  33. Berry, M.J.; Banu, L.; Chen, Y.; Mandel, S.J.; Kieffer, J.D.; Harney, J.W.; Larsen, P.R. Recognition of UGA as a selenocysteine codon in Type I deiodinase requires sequences in the 3′ untranslated region. Nature 1991, 353, 273–276. [Google Scholar] [CrossRef] [PubMed]
  34. Suzuki, K.T.; Doi, C.; Suzuki, N. Metabolism of 76Se-methylselenocysteine compared with that of 77Se-selenomethionine and 82Se-selenite. Toxicol. Appl. Pharmacol. 2006, 217, 185–195. [Google Scholar] [CrossRef]
  35. Combs, G.F. Selenium in global food systems. Br. J. Nutr. 2001, 85, 517–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Juresa, D.; Blanusa, M.; Francesconi, K.A.; Kienzl, N.; Kuehnelt, D. Biological availability of selenosugars in rats. Chem. Biol. Interact. 2007, 168, 203–210. [Google Scholar] [CrossRef] [PubMed]
  37. Zwolak, I.; Zaporowska, H. Selenium interactions and toxicity: A review. Cell Biol. Toxicol. 2012, 28, 31–46. [Google Scholar] [CrossRef]
  38. Boylan, L.M.; Spallholz, J.E. Selenium. In Sports Nutrition: Vitamins and Trace Elements, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2005; pp. 275–286. ISBN 9781420037913. [Google Scholar]
  39. Kieliszek, M.; Lipinski, B.; Błażejak, S. Application of sodium selenite in the prevention and treatment of cancers. Cells 2017, 6, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Allan, C.B.; Lacourciere, G.M.; Stadtman, T.C. Responsiveness of selenoproteins to dietary selenium. Annu. Rev. Nutr. 1999, 19, 1–16. [Google Scholar] [CrossRef] [PubMed]
  41. Rayman, M.P. The importance of selenium to human health. Lancet 2000, 356, 233–241. [Google Scholar] [CrossRef] [Green Version]
  42. Chen, J.; Berry, M.J. Selenium and selenoproteins in the brain and brain diseases. J. Neurochem. 2003, 86, 1–12. [Google Scholar] [CrossRef]
  43. Huang, J.Q.; Ren, F.Z.; Jiang, Y.Y.; Lei, X.G. Characterization of selenoprotein M and its response to selenium deficiency in chicken brain. Biol. Trace Elem. Res. 2016, 170, 449–458. [Google Scholar] [CrossRef]
  44. Ullah, H.; Liu, G.; Yousaf, B.; Ali, M.U.; Abbas, Q.; Munir, M.A.M.; Mian, M.M. Developmental selenium exposure and health risk in daily foodstuffs: A systematic review and meta-analysis. Ecotoxicol. Environ. Saf. 2018, 149, 291–306. [Google Scholar] [CrossRef]
  45. Kieliszek, M.; Błazejak, S. Current knowledge on the importance of selenium in food for living organisms: A review. Molecules 2016, 21, 609. [Google Scholar] [CrossRef] [Green Version]
  46. Rayman, M.P.; Winther, K.H.; Pastor-Barriuso, R.; Cold, F.; Thvilum, M.; Stranges, S.; Guallar, E.; Cold, S. Effect of long-term selenium supplementation on mortality: Results from a multiple-dose, randomised controlled trial. Free Radic. Biol. Med. 2018, 127, 46–54. [Google Scholar] [CrossRef]
  47. Zhang, X.; Xiong, W.; Chen, L.L.; Huang, J.Q.; Lei, X.G. Selenoprotein V protects against endoplasmic reticulum stress and oxidative injury induced by pro-oxidants. Free Radic. Biol. Med. 2020, 160, 670–679. [Google Scholar] [CrossRef]
  48. Chen, L.L.; Huang, J.Q.; Xiao, Y.; Wu, Y.Y.; Ren, F.Z.; Lei, X.G. Knockout of selenoprotein V affects regulation of selenoprotein expression by dietary selenium and fat Intakes in mice. J. Nutr. 2020, 150, 483–491. [Google Scholar] [CrossRef] [PubMed]
  49. Huang, J.Q.; Ren, F.Z.; Jiang, Y.Y.; Xiao, C.; Lei, X.G. Selenoproteins protect against avian nutritional muscular dystrophy by metabolizing peroxides and regulating redox/apoptotic signaling. Free Radic. Biol. Med. 2015, 83, 129–138. [Google Scholar] [CrossRef]
  50. Bösl, M.R.; Takaku, K.; Oshima, M.; Nishimura, S.; Taketo, M.M. Early embryonic lethality caused by targeted disruption of the mouse selenocysteine tRNA gene (Trsp). Proc. Natl. Acad. Sci. USA 1997, 94, 5531–5534. [Google Scholar] [CrossRef] [Green Version]
  51. Han, S.J.; Lee, B.C.; Yim, S.H.; Gladyshev, V.N.; Lee, S.R. Characterization of mammalian selenoprotein O: A redox-active mitochondrial protein. PLoS ONE 2014, 9, e95518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Hariharan, S.; Dharmaraj, S. Selenium and selenoproteins: It’s role in regulation of inflammation. Inflammopharmacology 2020, 28, 667–695. [Google Scholar] [CrossRef] [PubMed]
  53. Hu, S.; Rayman, M.P. Multiple nutritional factors and the risk of Hashimoto’s thyroiditis. Thyroid 2017, 27, 597–610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Short, S.P.; Pilat, J.M.; Williams, C.S. Roles for selenium and selenoprotein P in the development, progression, and prevention of intestinal disease. Free Radic. Biol. Med. 2018, 127, 26–35. [Google Scholar] [CrossRef]
  55. Zhang, X.; Liu, R.P.; Cheng, W.H.; Zhu, J.H. Prioritized brain selenium retention and selenoprotein expression: Nutritional insights into Parkinson’s disease. Mech. Ageing Dev. 2019, 180, 89–96. [Google Scholar] [CrossRef]
  56. Fan, T.W.M.; Teh, S.J.; Hinton, D.E.; Higashi, R.M. Selenium toxicity: Cause and effects in aquatic birds. Aquat. Toxicol. 2002, 57, 27–37. [Google Scholar] [CrossRef]
  57. Wang, N.; Tan, H.Y.; Li, S.; Xu, Y.; Guo, W.; Feng, Y. Supplementation of micronutrient selenium in metabolic diseases: Its role as an antioxidant. Oxidative Med. Cell. Longev. 2017, 2017, 7478523. [Google Scholar] [CrossRef] [PubMed]
  58. Chapman, P.M. Is selenium a global contaminant of potential concern? Integr. Environ. Assess. Manag. 2009, 5, 353–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ross, S.A.; Davis, C.D. MicroRNA, nutrition, and cancer prevention. Adv. Nutr. 2011, 2, 472–485. [Google Scholar] [CrossRef]
  60. Treiber, T.; Treiber, N.; Meister, G. Regulation of microRNA biogenesis and its crosstalk with other cellular pathways. Nat. Rev. Mol. Cell Biol. 2019, 20, 5–20. [Google Scholar] [CrossRef]
  61. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of microRNA biogenesis, mechanisms of actions, and circulation. Front. Endocrinol. 2018, 9, 402. [Google Scholar] [CrossRef] [Green Version]
  62. Quintanilha, B.J.; Reis, B.Z.; Silva Duarte, G.B.; Cozzolino, S.M.F.; Rogero, M.M. Nutrimiromics: Role of micrornas and nutrition in modulating inflammation and chronic diseases. Nutrients 2017, 9, 1168. [Google Scholar] [CrossRef]
  63. Lu, T.X.; Rothenberg, M.E. MicroRNA. J. Allergy Clin. Immunol. 2018, 141, 1202–1207. [Google Scholar] [CrossRef] [Green Version]
  64. Weber, J.A.; Baxter, D.H.; Zhang, S.; Huang, D.Y.; Huang, K.H.; Lee, M.J.; Galas, D.J.; Wang, K. The microRNA spectrum in 12 body fluids. Clin. Chem. 2010, 56, 1733–1741. [Google Scholar] [CrossRef]
  65. Lee, C.T.; Risom, T.; Strauss, W.M. Evolutionary conservation of microRNA regulatory circuits: An examination of microRNA gene complexity and conserved microRNA-target interactions through metazoan phylogeny. DNA Cell Biol. 2007, 26, 209–218. [Google Scholar] [CrossRef] [PubMed]
  66. Stark, A.; Brennecke, J.; Bushati, N.; Russell, R.B.; Cohen, S.M. Animal microRNAs confer robustness to gene expression and have a significant impact on 3′UTR evolution. Cell 2005, 123, 1133–1146. [Google Scholar] [CrossRef] [Green Version]
  67. Selbach, M.; Schwanhäusser, B.; Thierfelder, N.; Fang, Z.; Khanin, R.; Rajewsky, N. Widespread changes in protein synthesis induced by microRNAs. Nature 2008, 455, 58–63. [Google Scholar] [CrossRef]
  68. Brennecke, J.; Stark, A.; Russell, R.B.; Cohen, S.M. Principles of microRNA-target recognition. PLoS Biol. 2005, 3, e85. [Google Scholar] [CrossRef] [PubMed]
  69. Lynn, F.C. Meta-Regulation: microRNA regulation of glucose and lipid metabolism. Trends Endocrinol. Metab. 2009, 20, 452–459. [Google Scholar] [CrossRef]
  70. Bishop, K.S.; Ferguson, L.R. The interaction between epigenetics, nutrition and the development of cancer. Nutrients 2015, 7, 922–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Cui, J.; Zhou, B.; Ross, S.A.; Zempleni, J. Nutrition, microRNAs, and human health. Adv. Nutr. 2017, 8, 105–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Stefanska, B.; Karlic, H.; Varga, F.; Fabianowska-Majewska, K.; Haslberger, A.G. Epigenetic mechanisms in anti-cancer actions of bioactive food components—The implications in cancer prevention. Br. J. Pharmacol. 2012, 167, 279–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Beckett, E.L.; Yates, Z.; Veysey, M.; Duesing, K.; Lucock, M. The role of vitamins and minerals in modulating the expression of microRNA. Nutr. Res. Rev. 2014, 27, 94–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ma, Z.; Xue, Z.; Zhang, H.; Li, Y.; Wang, Y. Local and global effects of Mg2+ on Ago and miRNA-target interactions. J. Mol. Model. 2012, 18, 3769–3781. [Google Scholar] [CrossRef] [PubMed]
  75. Barr, I.; Smith, A.T.; Chen, Y.; Senturia, R.; Burstyn, J.N.; Guo, F. Ferric, not ferrous, heme activates RNA-binding protein DGCR8 for primary microRNA processing. Proc. Natl. Acad. Sci. USA 2012, 109, 1919–1924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Pogue, A.I.; Percy, M.E.; Cui, J.G.; Li, Y.Y.; Bhattacharjee, S.; Hill, J.M.; Kruck, T.P.A.; Zhao, Y.; Lukiw, W.J. Up-Regulation of NF-kB-sensitive miRNA-125b and miRNA-146a in metal sulfate-stressed human astroglial (HAG) primary cell cultures. J. Inorg. Biochem. 2011, 105, 1434–1437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Gabory, A.; Attig, L.; Junien, C. Sexual dimorphism in environmental epigenetic programming. Mol. Cell. Endocrinol. 2009, 304, 8–18. [Google Scholar] [CrossRef] [Green Version]
  78. Biswas, S.; Rao, C.M. Epigenetics in cancer: Fundamentals and beyond. Pharmacol. Ther. 2017, 173, 118–134. [Google Scholar] [CrossRef] [PubMed]
  79. Hulf, T.; Sibbritt, T.; Wiklund, E.D.; Bert, S.; Strbenac, D.; Statham, A.L.; Robinson, M.D.; Clark, S.J. Discovery pipeline for epigenetically deregulated miRNAs in cancer: Integration of primary miRNA transcription. BMC Genom. 2011, 12, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Rubio, B.; Mora, C.; Pintado, C.; Mazuecos, L.; Fernández, A.; López, V.; Andrés, A.; Gallardo, N. The nutrient sensing pathways FoxO1/3 and mTOR in the heart are coordinately regulated by central leptin through PPARβ/δ. Implications in cardiac remodeling. Metabolism 2021, 115, 154453. [Google Scholar] [CrossRef] [PubMed]
  81. Kim, S.G.; Buel, G.R.; Blenis, J. Nutrient regulation of the mTOR complex 1 signaling pathway. Mol. Cells 2013, 35, 463–473. [Google Scholar] [CrossRef] [Green Version]
  82. Ma, X.M.; Blenis, J. Molecular mechanisms of mTOR-mediated translational control. Nat. Rev. Mol. Cell Biol. 2009, 10, 307–318. [Google Scholar] [CrossRef] [PubMed]
  83. Sun, Y.; Ge, Y.; Drnevich, J.; Zhao, Y.; Band, M.; Chen, J. Mammalian target of rapamycin regulates miRNA-1 and follistatin in skeletal myogenesis. J. Cell Biol. 2010, 189, 1157–1169. [Google Scholar] [CrossRef] [Green Version]
  84. Rao, P.K.; Kumar, R.M.; Farkhondeh, M.; Baskerville, S.; Lodish, H.F. Myogenic factors that regulate expression of muscle-specific microRNAs. Proc. Natl. Acad. Sci. USA 2006, 103, 8721–8726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ye, P.; Liu, Y.; Chen, C.; Tang, F.; Wu, Q.; Wang, X.; Liu, C.G.; Liu, X.; Liu, R.; Liu, Y.; et al. An mTORC1-Mdm2-Drosha axis for miRNA biogenesis in response to glucose- and amino acid-deprivation. Mol. Cell 2015, 57, 708–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zoncu, R.; Efeyan, A.; Sabatini, D.M. MTOR: From growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol. 2011, 12, 21–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Zhang, Y.; Yu, B.; He, J.; Chen, D. From nutrient to microRNA: A novel insight into cell signaling involved in skeletal muscle development and disease. Int. J. Biol. Sci. 2016, 12, 1247–1261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Maciel-Dominguez, A.; Swan, D.; Ford, D.; Hesketh, J. Selenium alters miRNA profile in an intestinal cell line: Evidence that miR-185 regulates expression of GPX2 and SEPSH2. Mol. Nutr. Food Res. 2013, 57, 2195–2205. [Google Scholar] [CrossRef]
  89. Xing, Y.; Liu, Z.; Yang, G.; Gao, D.; Niu, X. MicroRNA expression profiles in rats with selenium deficiency and the possible role of the Wnt/β-catenin signaling pathway in cardiac dysfunction. Int. J. Mol. Med. 2015, 35, 143–152. [Google Scholar] [CrossRef]
  90. Matoušková, P.; Hanousková, B.; Skálová, L. MicroRNAs as potential regulators of glutathione peroxidases expression and their role in obesity and related pathologies. Int. J. Mol. Sci. 2018, 19, 1199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Huang, J.Q.; Zhou, J.C.; Wu, Y.Y.; Ren, F.Z.; Lei, X.G. Role of glutathione peroxidase 1 in glucose and lipid metabolism-related diseases. Free Radic. Biol. Med. 2018, 127, 108–115. [Google Scholar] [CrossRef]
  92. Alehagen, U.; Johansson, P.; Aaseth, J.; Alexander, J.; Wågsäter, D. Significant changes in circulating microRNA by dietary supplementation of selenium and coenzyme Q10 in healthy elderly males. A subgroup analysis of a prospective randomized double-blind placebo-controlled trial among elderly Swedish citizens. PLoS ONE 2017, e0174880. [Google Scholar] [CrossRef]
  93. Sun, W.; Wang, Q.; Guo, Y.; Zhao, Y.; Wang, X.; Zhang, Z.; Deng, G.; Guo, M. Selenium suppresses inflammation by inducing microRNA-146a in Staphylococcus aureus-infected mouse mastitis model. Oncotarget 2017, 8, 110949–110964. [Google Scholar] [CrossRef] [Green Version]
  94. Zhang, Z.B.; Guo, Y.F.; Li, C.Y.; Qiu, C.W.; Guo, M.Y. Selenium influences mmu-miR-155 to inhibit inflammation in: Staphylococcus aureus-induced mastitis in mice. Food Funct. 2019, 10, 6543–6555. [Google Scholar] [CrossRef]
  95. Guo, X.; Yang, Q.; Zhang, W.; Chen, Y.; Ren, J.; Gao, A. Associations of blood levels of trace elements and heavy metals with metabolic syndrome in Chinese male adults with microRNA as mediators involved. Environ. Pollut. 2019, 248, 66–73. [Google Scholar] [CrossRef]
  96. Reis, B.Z.; Duarte, G.B.S.; Vargas-Mendez, E.; Ferreira, L.R.P.; Barbosa, F.; Cercato, C.; Rogero, M.M.; Cozzolino, S.M.F. Brazil nut intake increases circulating miR-454-3p and miR-584-5p in obese women. Nutr. Res. 2019, 67, 40–52. [Google Scholar] [CrossRef] [PubMed]
  97. Speckmann, B.; Grune, T. Epigenetic effects of selenium and their implications for health. Epigenetics 2015, 10, 179–190. [Google Scholar] [CrossRef]
  98. Joven, J.; Micol, V.; Segura-Carretero, A.; Alonso-Villaverde, C.; Menéndez, J.A.; Aragonès, G.; Barrajón-Catalán, E.; Beltrán-Debón, R.; Camps, J.; Cufí, S.; et al. Polyphenols and the modulation of gene expression pathways: Can we eat our way out of the danger of chronic disease? Crit. Rev. Food Sci. Nutr. 2014, 54, 985–1001. [Google Scholar] [CrossRef] [PubMed]
  99. Jabłońska, E.; Reszka, E. Selenium and epigenetics in cancer: Focus on DNA methylation. In Advances in Cancer Research; Elsevier: Amsterdam, The Netherlands, 2017; Volume 136, pp. 193–234. [Google Scholar]
  100. Min, Z.; Guo, Y.; Sun, M.; Hussain, S.; Zhao, Y.; Guo, D.; Huang, H.; Heng, L.; Zhang, F.; Ning, Q.; et al. Selenium-Sensitive miRNA-181a-5p targeting SBP2 regulates selenoproteins expression in cartilage. J. Cell. Mol. Med. 2018, 22, 5888–5898. [Google Scholar] [CrossRef]
  101. Fan, R.F.; Cao, C.Y.; Chen, M.H.; Shi, Q.X.; Xu, S.W. Gga-let-7f-3p promotes apoptosis in selenium deficiency-induced skeletal muscle by targeting selenoprotein K. Metallomics 2018, 10, 941–952. [Google Scholar] [CrossRef] [PubMed]
  102. Yang, T.; Liu, T.; Cao, C.; Xu, S. MiR-200a-5p augments cardiomyocyte hypertrophy induced by glucose metabolism disorder via the regulation of selenoproteins. J. Cell. Physiol. 2019, 234, 4095–4103. [Google Scholar] [CrossRef] [PubMed]
  103. Yang, T.; Cao, C.; Yang, J.; Liu, T.; Lei, X.G.; Zhang, Z.; Xu, S. MiR-200a-5p regulates myocardial necroptosis induced by Se deficiency via targeting RNF11. Redox Biol. 2018, 15, 159–169. [Google Scholar] [CrossRef]
  104. Chi, Q.; Luan, Y.; Zhang, Y.; Hu, X.; Li, S. The regulatory effects of miR-138-5p on selenium deficiency-induced chondrocyte apoptosis are mediated by targeting SelM. Metallomics 2019, 11, 845–857. [Google Scholar] [CrossRef]
  105. Potenza, N.; Castiello, F.; Panella, M.; Colonna, G.; Ciliberto, G.; Russo, A.; Costantini, S.C. Human miR-544a modulates SELK expression in hepatocarcinoma cell lines. PLoS ONE 2016, 11, e0156908. [Google Scholar] [CrossRef]
  106. Qin, L.; Zhang, Y.; Wan, C.; Wang, Z.; Cong, Y.; Li, S. MiR-196-5p involvement in selenium deficiency-induced immune damage: Via targeting of NFκBIA in the chicken trachea. Metallomics 2020, 12, 1679–1692. [Google Scholar] [CrossRef] [PubMed]
  107. Liu, T.; Yang, T.; Xu, Z.; Tan, S.; Pan, T.; Wan, N.; Li, S. MicroRNA-193b-3p regulates hepatocyte apoptosis in selenium-deficient broilers by targeting MAML1. J. Inorg. Biochem. 2018, 186, 235–245. [Google Scholar] [CrossRef]
  108. Wan, N.; Xu, Z.; Chi, Q.; Hu, X.; Pan, T.R.; Liu, T.; Li, S. MicroRNA-33-3p involved in selenium deficiency-induced apoptosis via targeting ADAM10 in the chicken kidney. J. Cell. Physiol. 2019, 234, 13693–13704. [Google Scholar] [CrossRef]
  109. Zhang, Y.; Wan, N.; Pan, T.; Hu, X.; Liu, Q.; Li, S. MicroRNA-33-3p regulates vein endothelial cell apoptosis in selenium-deficient broilers by targeting E4F1. Oxidative Med. Cell. Longev. 2019, 2019, 6274010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Zheng, X.; Hu, X.; Ge, T.; Li, M.; Shi, M.; Luo, J.; Lai, H.; Nie, T.; Li, F.; Li, H. MicroRNA-328 is involved in the effect of selenium on hydrogen peroxide-induced injury in H9c2 cells. J. Biochem. Mol. Toxicol. 2017, 31, e21920. [Google Scholar] [CrossRef] [PubMed]
  111. Cai, J.; Yang, J.; Liu, Q.; Gong, Y.; Zhang, Y.; Zhang, Z. Selenium deficiency inhibits myocardial development and differentiation by targeting the mir-215-5p/CTCF axis in chicken. Metallomics 2019, 11, 415–428. [Google Scholar] [CrossRef] [PubMed]
  112. Cai, J.; Yang, J.; Liu, Q.; Gong, Y.; Zhang, Y.; Zheng, Y.; Yu, D.; Zhang, Z. Mir-215-5p induces autophagy by targeting PI3K and activating ROS-mediated MAPK pathways in cardiomyocytes of chicken. J. Inorg. Biochem. 2019, 193, 60–69. [Google Scholar] [CrossRef] [PubMed]
  113. Yang, J.; Gong, Y.; Cai, J.; Liu, Q.; Zhang, Z. lnc-3215 suppression leads to calcium overload in selenium deficiency-induced chicken heart lesion via the lnc-3215-miR-1594-TNN2 pathway. Mol. Ther. Nucleic Acids 2019, 18, 1–15. [Google Scholar] [CrossRef] [Green Version]
  114. Liu, Q.; Cai, J.; Gao, Y.; Yang, J.; Gong, Y.; Zhang, Z. MiR-2954 Inhibits PI3K signaling and induces autophagy and apoptosis in myocardium selenium deficiency. Cell. Physiol. Biochem. 2018, 51, 778–792. [Google Scholar] [CrossRef]
  115. Wang, L.; Shi, X.; Zheng, S.; Xu, S. Selenium deficiency exacerbates LPS-induced necroptosis by regulating miR-16-5p targeting PI3K in chicken tracheal tissue. Metallomics 2020, 12, 562–571. [Google Scholar] [CrossRef] [PubMed]
  116. Pan, T.; Hu, X.; Liu, T.; Xu, Z.; Wan, N.; Zhang, Y.; Li, S. MiR-128-1-5p regulates tight junction induced by selenium deficiency via targeting cell adhesion molecule 1 in broilers vein endothelial cells. J. Cell. Physiol. 2018, 233, 8802–8814. [Google Scholar] [CrossRef] [PubMed]
  117. Cui, J.; Liu, H.; Xu, S. Selenium-Deficient diet induces necroptosis in the pig brain by activating TNFR1: Via mir-29a-3p. Metallomics 2020, 12, 1290–1301. [Google Scholar] [CrossRef]
  118. Liu, C.; Sun, Z.; Xu, Z.; Liu, T.; Pan, T.; Li, S. Down-Regulation of microRNA-155 promotes selenium deficiency-induced apoptosis by tumor necrosis factor receptor superfamily member 1B in the broiler spleen. Oncotarget 2017, 8, 58513–58525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Gao, X.J.; Tang, B.; Liang, H.H.; Yi, L.; Wei, Z.G. Selenium deficiency inhibits micRNA-146a to promote ROS-induced inflammation via regulation of the MAPK pathway in the head kidney of carp. Fish Shellfish Immunol. 2019, 91, 284–292. [Google Scholar] [CrossRef]
  120. Tarek, M.; Louka, M.L.; Khairy, E.; Ali-Labib, R.; Zaky, D.Z.; Montasser, I.F. Role of microRNA-7 and selenoprotein P in hepatocellular carcinoma. Tumor Biol. 2017, 39. [Google Scholar] [CrossRef] [Green Version]
  121. Chen, Z.; Gu, D.; Zhou, M.; Shi, H.; Yan, S.; Cai, Y. Regulatory role of miR-125a/b in the suppression by selenium of cadmium-induced apoptosis via the mitochondrial pathway in LLC-PK1 cells. Chem. Biol. Interact. 2016, 243, 35–44. [Google Scholar] [CrossRef]
  122. Atteia, H.H.; Arafa, M.H.; Prabahar, K. Selenium nanoparticles prevents lead acetate-induced hypothyroidism and oxidative damage of thyroid tissues in male rats through modulation of selenoenzymes and suppression of miR-224. Biomed. Pharmacother. 2018, 99, 486–491. [Google Scholar] [CrossRef]
  123. Yin, K.; Cui, Y.; Sun, T.; Qi, X.; Zhang, Y.; Lin, H. Antagonistic effect of selenium on lead-induced neutrophil apoptosis in chickens via miR-16-5p targeting of PiK3R1 and IGF1R. Chemosphere 2020, 246, 125794. [Google Scholar] [CrossRef]
  124. Zhang, J.; Zheng, S.; Wang, S.; Liu, Q.; Xu, S. Cadmium-Induced oxidative stress promotes apoptosis and necrosis through the regulation of the miR-216a-PI3K/AKT axis in common carp lymphocytes and antagonized by selenium. Chemosphere 2020, 258, 127341. [Google Scholar] [CrossRef]
  125. Diwadkar-Navsariwala, V.; Prins, G.S.; Swanson, S.M.; Birch, L.A.; Ray, V.H.; Hedayat, S.; Lantvit, D.L.; Diamond, A.M. Selenoprotein deficiency accelerates prostate carcinogenesis in a transgenic model. Proc. Natl. Acad. Sci. USA 2006, 103, 8179–8184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The metabolism of Se dietary forms. SeMet, (selenomethionine); SeCys, (selenocysteine); H2Se, (dihydrogen selenide); SELENOP, (selenoprotein P); CH3SeH, (methyl selenol); (CH3)2Se, (dimethyl selenide); (CH3)3Se+, (trimethyl selenonium); SeO2, (selenium dioxide).
Figure 1. The metabolism of Se dietary forms. SeMet, (selenomethionine); SeCys, (selenocysteine); H2Se, (dihydrogen selenide); SELENOP, (selenoprotein P); CH3SeH, (methyl selenol); (CH3)2Se, (dimethyl selenide); (CH3)3Se+, (trimethyl selenonium); SeO2, (selenium dioxide).
Nutrients 13 01527 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, X.; Dong, Y.-L.; Li, T.; Xiong, W.; Zhang, X.; Wang, P.-J.; Huang, J.-Q. Dietary Selenium Regulates microRNAs in Metabolic Disease: Recent Progress. Nutrients 2021, 13, 1527. https://doi.org/10.3390/nu13051527

AMA Style

Huang X, Dong Y-L, Li T, Xiong W, Zhang X, Wang P-J, Huang J-Q. Dietary Selenium Regulates microRNAs in Metabolic Disease: Recent Progress. Nutrients. 2021; 13(5):1527. https://doi.org/10.3390/nu13051527

Chicago/Turabian Style

Huang, Xin, Yu-Lan Dong, Tong Li, Wei Xiong, Xu Zhang, Peng-Jie Wang, and Jia-Qiang Huang. 2021. "Dietary Selenium Regulates microRNAs in Metabolic Disease: Recent Progress" Nutrients 13, no. 5: 1527. https://doi.org/10.3390/nu13051527

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop