Open access peer-reviewed chapter

The Impact of Dietary Compounds in Functional Foods on MicroRNAs Expression

Written By

Wittaya Chaiwangyen

Submitted: 26 October 2020 Reviewed: 19 February 2021 Published: 08 April 2021

DOI: 10.5772/intechopen.96746

From the Edited Volume

Functional Foods - Phytochemicals and Health Promoting Potential

Edited by Muhammad Sajid Arshad and Muhammad Haseeb Ahmad

Chapter metrics overview

498 Chapter Downloads

View Full Metrics

Abstract

MicroRNAs (miRNAs) are a class of non-coding endogenous RNA molecules that are involved in post-transcriptional gene silencing via binding to their target messenger RNA, leading to mRNA degradation or translational repression. MicroRNAs can be modulated by several factors including hormones, transcription factors, and dietary compounds. These biologically active compounds have positive impact on the progression of human pathology including non-communicable diseases, which indicating that administration of diet may have potential as therapeutic agents in modulating the risk of chronic diseases. Interestingly, evidence emerging in recent years suggests that dietary miRNAs can be absorbed in human circulation, modulated human gene expression and biological functions. The exploitation of the miRNA functioning within different origins, cellular miRNAs and dietary miRNAs will help us to understand the molecular machinery as well as the regulatory mechanisms involved in fundamentally important biological processes. Therefore, this knowledge may be applied of natural bioactive compounds in preventive or therapeutic approaches.

Keywords

  • functional foods
  • microRNAs
  • dietary microRNAs
  • chronic diseases
  • non-communicable diseases

1. Introduction - Origin, biogenesis and functions of microRNAs

MicroRNAs or miRNAs are a class of small non-coding RNA approximately 21–25 nucleotides that modulate on gene expression post-transcriptionally via binding to the 3′ untranslated region (3′-UTR) of the target messenger RNA (mRNA), resulting in mRNA degradation or translational repression. The first miRNA, lin-4, was discovered by Ambro and his research group in 1993 and it was found to be related with larva development in Caenorhabditis elegans [1]. Up to date, almost 2000 miRNAs have been identified in humans (http://www.miRbase.org – 7.3.2019) [1]. It has been estimated that 1–4% of human genes expression can be regulated by miRNAs, which is the largest of genomic regulator [2]. In mammals, miRNAs have been associated with various cellular pathways with the regulation of cell differentiation, cell cycle, proliferation, apoptosis, hematopoiesis, and other cellular functions. Recent studies have highlighted the importance of mRNA regulation mechanism by validation and differential miRNA expression in a variety of human pathological conditions, including chronic diseases.

miRNAs are normally transcribed by RNA polymerase II from miRNA genes. This transcription leads to generate a primary miRNA transcript (pri-miRNA). Then, pri-miRNA is further cleaved by a microprocessor complex, which consists of Drosha, the double-stranded RNase III enzyme and DiGeorge syndrome critical region 8 (DGCR8), important cofactor, into a hairpin structure precursor miRNA (pre-miRNA) in the nucleus (Figure 1). The double strand pre-miRNAs with 70 nucleotides are then exported to the cytoplasm by the process of nuclear export factor exportin-5. The pre-miRNA is then processed by RNase III, Dicer, thereby generating a mature miRNA:miRNA duplex approximately 22 nucleotides in length and without a hairpin structure. The helicase enzyme cleaves miRNA duplexes into single-stranded miRNAs and incorporated into the Argonaute (AGO), TRBP and PACT proteins to form the RNA-induced silencing complex (RISC). Usually, other single strand called passenger strand or the star (*) strand will be degraded, while single strand mature miRNA is able to bind with its target mRNA and mediating translational inhibition or mRNA degradation, along with their sequence complementarity to the target [1, 3]. In plants, target mRNA will be degraded if miRNA has perfect or near-perfect complementarity to its target. In contrast to mammal, miRNAs bind to partially complementary sites in the 3′-UTRs of target mRNA, which leading to translational repression [4]. the target mRNA is either blocked (imperfect complementary) or degraded (perfect complementary) of the ribosomal translation, which sequentially impacts the cellular functions.

Figure 1.

miRNA biogenesis. miRNA gene is transcribed by RNA polymerase II and then forming the primary miRNA transcript (pri-miRNA), which is further cleaved by the Drosha/DGCR8 complex to generate the precursor miRNA (pre-miRNA). Pre-miRNA is then exported into the cytoplasm by exportin 5/RAN-GTP and further processed by dicer to create the mature miRNA, which is loaded into RISC, which contains AGO, PACT and TRBP proteins. Mature miRNA that binding to its target mRNA by perfect complementary binding and resulting in gene suppression by mRNA degradation. The partially complementary binding of miRNA and its target mRNA, which in turn inhibit the protein translation.

Phytochemicals are major plant-derived compounds that naturally found in vegetables, fruits, medicinal plants or other plants with medicinal properties including antioxidant, anti-diabetic, anti-inflammatory, antimicrobial, antidepressant, anticancer and prevention in other chronic non-communicable diseases [5, 6, 7]. Phenolic and flavonoid compounds are the most important group of bioactive compounds and second metabolites in plants which comprise of essential molecules of human diet [6, 8]. It has been shown that bioactive compounds can modulate the endogenous miRNAs expression [1, 9, 10, 11, 12]. Recently, some studies have revealed that plant-derived miRNAs (dietary miRNAs) as new bioactive compounds in plants can affect the synthesis of endogenous miRNAs [13, 14, 15]. Strikingly, miRNAs do not function only their origins but they are able to regulate the gene expression in cross-kingdom. Therefore, bioactive compounds present in functional foods are potentially regulate endogenous miRNAs expression.

Advertisement

2. Dietary compounds and endogenous miRNAs

Extensive studies have been performed to understand the molecular mechanism of bioactive compounds with a positive effect on chronic diseases or non-communicable diseases such as arthritis, cancer, cardiovascular diseases, diabetes and obesity [1, 16]. Emerging evidences confirm that alteration of endogenous miRNAs expression can be influenced by bioactive compounds in functional foods [16, 17] (Figure 2 and Table 1).

Figure 2.

Influences of bioactive compounds and dietary miRNAs on human non-communicable diseases. Ascending arrows represent up-regulated miRNAs and descending arrows represent down-regulated miRNAs by bioactive compounds. The green triangles show the positive impact of dietary miRNAs on human health.

Dietary compoundmiRNA expressionTarget of miRNADiseasesReferences
Up-regulationDown-regulation
Acetyl-11-Keto-β-Boswellic AcidmiR-27a miR-34aUnknownColorectal cancer[23]
miR-155SOCS-1Neuroinflammation[21]
miR-206ER-αBreast cancer[22]
ArctigeninmiR-16
miR-199a
UnknownNeuroinflammation[28]
miR-21
miR-19b miR-148a
UnknownProstate cancer[29]
Cinnamic acid derivativesmiR-143MAPK/Erk5Colon cancer[31]
miR-145UnknownGastric cancer[33]
CurcuminmiR-15a, miR-16, miR-34a, miR-146b-5p
miR-181b
miR-19a miR-19bUnknownBreast cancer[38]
miR-101, miR-200b, miR-200c, miR-141 miR-429miR-21UnknownColorectal cancer[39, 40]
miR-21Gastric cancer[41]
miR-145 miR-1275
miR-1908 miR-3127 miR-3178 miR-3198
miR-23b*, miR-183 miR-193b* miR-210 miR-222* miR-494 miR-664*
miR-671-5p
Oct4Prostate cancer[42]
miR-181bCXCL1 CXCL2Breast cancer[43]
miR-378p38glioblastoma[44]
miR-124 miR-155UnknownNeurodegenerative disorder[45]
3,3′-Diindolyl-methanelet-7
miR-34a miR-150-5p
EZH2, Notch1
AR
Ahr
Prostate cancer[46]
miR-200FoxM1Breast cancer[47]
miR-212/132 cluster
miR-21
Sox4 Cdc25ABreast cancer[48, 49]
let-7b, let-7c, let-7d, let-7e, and miR-200b/cZEB-1,
E-cadherin
Pancreatic cancer[50]
miR-146aUnknownPancreatic cancer[51]
(−)-Epigallocatechin-3-GallatemiR-296STAT3Nasopharyngeal carcinoma[57]
let-7a miR34ac-MycHepatocellular carcinoma[58]
miR-34amiR-93UnknownProstate cancer[59]
miR-29 miR-210miR-125b miR-203UnknownCervical cancer[60]
let-7HMGA2Melanoma cell[61]
miR-384Beclin-1Myocardial ischemia/
reperfusion
[62]
miR-140-3pUnknownOsteoarthritis[63]
miR-10b miR-181a miR-221UnknownLiver fibrosis[64]
GenisteinmiR-23bUnknownBreast cancer[66]
miR-1260bsRRP1 Smad4Prostate cancer[67]
miR-1260bsFRP1, Dkk2, Smad4Renal cancer[68]
miR-27aUnknownLung cancer[69]
miR-29bUnknownLung cancer[70]
miR-451UnknownChronic liver disease[72]
QuercetinmiR-200b-3pNotch1Pancreatic cancer[75]
miR-146aEGFRBreast cancer[76]
miR-16HOXA10Oral cancer[77]
miR-22WNT1/β-cateninOral cancer[78]
miR-97
miR-298 miR-2218 miR-1502 miR-2117
UnknownOxidative stress in pheochromocytoma[79]
miR-503-5p
miR-1283, miR-3714 miR-6867-5p
CCND1Endometriosis[80]
miR-122miR-21UnknownLiver fibrosis[81]
miR-199Sert1Hypoxia[82]
SilymarinmiR-203class 1 HDAC proteins and ZEB1Lung cancer[84]
miR-155UnknownRheumatoid arthritis[85]
miR-122UnknownLiver damage[86]
miR-122 miR-192 miR-194UnknownLiver damage[87]
β-Sitosterol-d-glucosidemiR-10aUnknownBreast cancer[89]
SulforaphanemiR-23b miR-92b miR-381 miR-382UnknownBreast cancer[92]
miR-616-5pGSK3β/β-cateninLung cancer[93]
miR-135b-5pmiR-30a-3pRASAL2
Cx43
Pancreatic cancer[94]
miR-200cUnknownOral cancer[96]
miR-9 miR-326UnknownGastric cancer[97]
miR-124-3pSTAT3Nasopharyngeal cancer[98]
miR-423-5pUnknownLiver fibrosis[99]
miR-155UnknownNeuroinflammation[100]

Table 1.

Summary of miRNAs bioactive compounds and miRNAs expression in human pathology.

2.1 Acetyl-11-keto-β-boswellic acid

3-acetyl-11-keto- β -boswellic acid (AKBA) is pentacyclic triterpene acids that mainly found in Boswellia serrata and it has been shown in medicinal properties for chronic diseases including anti-tumor, anti-inflammation, antioxidant, asthma, diabetes, atherosclerosis and analgesic [18, 19, 20]. AKBA showed the reduction of inflammatory miRNA expression, miR-155 and increased the expression of miR-155 target gene, suppressor of cytokine signaling-1 (SOCS-1) in neuroinflammatory mice model [21]. Therefore, AKBA might be used for treatment of neuroinflammatory disorders. AKBA also induced breast cancer cell cycle arrest, apoptosis and decreased the expression of estrogen receptor alpha (ER-α) via the up-regulation of miR-206 [22]. In addition, combination of AKBA and curcumin suppressed colorectal cancer growth through the down-regulated miR-27a and miR-34a expression [23].

2.2 Arctigenin

Arctigenin (AR) is a phenylpropanoid dizbenzylbutyrolactone lignin and was first identified in Arctium lappa L. Several studies showed anti-inflammatory, anti-cancer, anti-viral, immune modulatory activities of AR [24, 25, 26, 27]. The study demonstrated that AR upregulated miR-16 and miR-199a expression by decreasing upstream protein (IKKα and IKKβ) expression and inhibiting NF-κB signaling pathway activity, thereby reducing inflammatory cytokines production in neural cells [28]. The combination treatment of AR and quercetin significantly inhibited the oncogenic miRNAs expression including miR-19b, miR-21 and miR-148a in prostate cancer cells. AR and quercetin also showed anti-migration activity in prostate cancer cells [29].

2.3 Cinnamic acid derivatives

Cinnamic acid derivatives can occur naturally in plants and their structure composing of benzene ring and acrylic acid group. Several compounds of cinnamic acid derivatives have been identified including artepilin C, baccharin, drupanin, ferulic acid, curcumin, caffeic acid, p-hydroxycinnamic acid, coumaric and chlorogenic acids, etc. [30, 31]. Medicinal activities of cinnamic acid derivatives have been reported such as anti-inflammatory, anti-oxidant, anti-viral, anti-microbial, anti-diabetic, neuroprotective and anti-tumor activities [30, 31, 32]. Cinnamic acid derivatives from propolis significantly induced colon cancer cell apoptosis through TRAIL/DR4/5 and/or FasL/Fas death-signaling pathways and via the upregulated miR-143 expression, resulting in decreased the target gene MAPK/Erk5 expression and its downstream target c-Myc [31]. Moreover, Li et al. demonstrated that cinnamic acid derivatives decreased gastric cancer cell proliferation through the up-regulation of miR-145 and down-regulation P13K/Akt signaling pathway [33]. Therefore, cinnamic acid derivatives have a potential as therapeutic agents for cancer.

2.4 Curcumin

Curcumin[(1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptane-3,5-dione] is well known as natural polyphenol and derived from the rhizome of turmeric or Curcuma longa Linn [34, 35]. Curcumin has been shown to possess of several medicinal properties such as anti-inflammation, antioxidant, pro-apoptosis, chemoprevention, anti-proliferation, wound healing, anti-nociception, anti-parasite, anti-malaria, anti-diabetes, neuroprotection and anti-tumor [34, 36]. Numerous studies have been focused on curcumin as a novel anti-cancer drug due to the inhibition of NF-kB, Akt/PI3K, and MAPK pathways and enhancement of p53 by curcumin, thereby inhibited several cancer cells proliferation, migration, invasion and induced apoptosis [35, 37]. Emerging data suggest that curcumin dysregulate oncogenic miRNAs and tumor suppressor miRNAs expression in various type of cancers such as lung cancer, prostate cancer, breast cancer, colorectal cancer, nasopharyngeal carcinoma, pancreatic cancer, ovarian cancer and etc. [35]. Curcumin have been shown to up-regulation of miR-15a, miR-16, miR-34a, miR-146b-5p and miR-181b and down-regulation of miR-19a and miR-19b upon treatment of several breast cancer cell lines with curcumin [38]. Curcumin but not 5-fluorouracil, upregulated the expression of miR-101, miR-200b, miR-200c, miR-141 and miR-429 and down-regulated oncogenic miR-21 in colorectal cancer cells [39, 40]. In addition, miR-21 was down-regulated in gastric cancer with curcumin treatment, resulting in inhibition of cell migration and invasion by regulation of the PTEN/PI3K/AKT pathway [41]. Lui et al. showed that curcumin up-regulated 6 miRNAs (miR-145, miR-1275, miR-1908, miR-3127, miR-3178, and miR-3198), whereas 8 miRNAs (miR-23b*, miR-183, miR-193b* miR-210, miR-222*, miR-494, miR-664*, miR-671-5p) were down-regulated when treated with curcumin in human prostate cancer stem cells (HuPCaSCs) [42]. Experimental confirmed of miR-145 function in HuPCaSCs revealed that miR-145 inhibited cell proliferation by targeting transcription factors Oct4 [42]. Another study also reported that miR-181b was up-regulated by curcumin and inhibited breast cancer cell proliferation, invasion and induced cell apoptosis by targeting CXCL1 and CXCL2 [43]. Inhibitory effect of curcumin on glioblastoma cell growth was observed and curcumin also up-regulated miR-378 expression and p38 was the target of miR-378 [44]. Curcumin and Pioglitazone combination have a potential as therapeutic applications for neurodegenerative disorders by increasing of miR-124 and miR-155 expression, thereby inhibiting the inflammatory cytokines TNF-α, IL-1β and IL-6 production and inflammation-associated enzymes COX-2, iNOS through inhibition of NF-κB activity in animal model [45].

2.5 3,3′-Diindolylmethane

3,3′-diindolylmethane (DIM) is a naturally active compound found in stomach, which derived from indole-3-carbinol (I3C) that present in cruciferous vegetables. DIM has been reported to regulate several miRNAs expression in cancer. Tumor suppressor miRNAs was upregulated by DIM in prostate cancer cells including let-7, miR-34a and miR-150-5p by targeting EZH2, Notch1 and AR and Ahr, respectively [46]. DIM also upregulated tumor suppressor miR-200, which led to inhibit the expression of FoxM1 in breast cancer cells [47]. miR-212/132 cluster and miR-21 were upregulated by DIM, which downregulated the expression of Sox4 and Cdc25A, respectively in breast cancer [48, 49]. Moreover, DIM upregulated let-7b, let-7c, let-7d, let-7e, and miR-200b/c expression, which led to inhibit the expression of ZEB-1, E-cadherin in pancreatic cancer cells [50]. It has been reported miR-146a was upregulated upon treated with DIM and suppressed the expression of MTA2, NF-κB, IRAK1, EGFR in pancreatic cancer cells [51].

DIM showed the modulation of miRNAs expression in other inflammatory diseases. The expression of miR-106a, miR-20b, and miR-125b-5p were increased after treatment with DIM and suppressed the expression of IRAK4 and TNF-α to limit responses to TLRs activated by LPS in acute liver failure (ALF) animal model [52]. DIM significantly upregulated miR-200c, miR-146a, miR-16, miR-93, and miR-22 in brain CD4+ T cells and inhibited the expression of cyclin E1 and B-cell lymphoma-2 in experimental autoimmune encephalomyelitis (EAE), a murine model of multiple sclerosis [53].

2.6 (−)-Epigallocatechin-3-gallate

(−)-Epigallocatechin-3-Gallate or EGCG is a major polyphenol compound in green tea (Camellia sinensis) and derivative of catechin [3, 16]. EGCG is a powerful antioxidant, anticancer and antiangiogenic properties, which has a potential role to influence human diseases [54, 55, 56]. EGCG suppressed nasopharyngeal carcinoma cell migration and invasion through a novel signaling axis of miR-296/STAT3 regulation [57]. Gold nanoparticles (AuNPs) have been used for drug delivery as their stability and increase drug bioavailability as well as accumulation of drug in cancer cells. EGCG-capped gold nanoparticles upregulated the tumor suppressor miRNAs, let-7a and miR34a, which consecutively their targeted gene, caspase-3 was upregulated, and c-Myc protein was decreased in hepatocellular carcinoma cells [58]. miR-34a is one of the tumor suppressor miRNAs that downregulated, whereas miR-93 is highly up-regulated in prostate cancer cells. Co-transfection of miR.34a mimic and miR.93 inhibitor along with EGCG significantly decreased androgen receptor (AR) and prostate-specific antigen (PSA) expression when compared to the co-transfection without EGCG [59]. In cervical carcinoma cells, Hela (HPV16/18+), EGCG inhibited cell growth and up-regulated miR-29 and miR-210 expression, while down-regulated the expression of miR-125b and miR-203 [60]. Up-regulation of let-7 was observed in EGCG treated melanoma cells, which led to inhibit the expression of high mobility group A2 (HMGA2) [61].

EGCG showed the protective effect against myocardial ischemia/reperfusion (I/R) injury through up-regulation of miR-384-mediated autophagy by targeting Beclin-1 via activating the PI3K/Akt signaling pathway [62]. EGCG also demonstrated the anti-arthritic effects by inhibited IL-1β-induced ADAMTS5 expression and up-regulated the expression of miR-140-3p in osteoarthritis chondrocytes [63]. EGCG treatment has potential role of preventing toxin-induced fibrosis by suppression of osteopontin expression and up-regulation of miR-10b, miR-181a and miR-221 in liver hepatocellular carcinoma cells [64].

2.7 Genistein

Genistein belongs to isoflavone family and presents in soybeans with antiangiogenic, anti-metastasis, anti-inflammatory, anti-oxidant, cell cycle arrest and induction of apoptosis effects [65]. Genistein can regulate the expression of miRNAs in several call types [65]. It has been reported that treatment of genistein up-regulated miR-23b and inhibited breast cancer cell growth [66]. Genistein also exhibited anti-tumor effect by down-regulated miR-1260b and targeting sRRP1 and Smad4 through DNA methylation or histone modifications in prostate cancer cells [67]. The same research group reported that miR-1260b was highly expressed in renal cancer cells and miR-1260b was down-regulated in genistein treated renal cancer cells [68]. The treatment of miR-1260b inhibitor inhibited the expression of its target genes, sFRP1, Dkk2, Smad4 [68]. Treatment with genistein induced non-small lung cancer cell apoptosis, caspase-3/9 activation and inhibited cell proliferation via up-regulation of miR-27a -mediated MET signaling [69]. Co-encapsulate miR-29b with genistein in hybrid nanoparticles (GMLHN) has been studied to treat effectively in non-small lung cancer cell and GMLHN showed the anti-proliferative effect by down-regulation of phosphorylated AKT (pAKT) and phosphorylated phosphoinositide-3 kinase (p-PI3K) [70].

Genistein promoted myoblast proliferation and differentiation through down-regulated miR-222 expression, resulting in increased expression of its target genes, MyoG, MyoD, and ERα [71]. Interestingly, genistein up-regulated miR-451 expression and inhibited IL1β expression and inflammation in chronic liver disease nonalcoholic steatohepatitis (NASH) mice model [72].

2.8 Quercetin

Quercetin is bioactive flavonoids that can be found in fruits and vegetables including onion, kale, apple, many berries, citrus fruits and tea [73]. Anti-cancer, anti-inflammatory, antioxidant, anti-diabetes, anti-atherosclerosis and anti-viral effects have been reported in different in vitro studies for quercetin [74]. Several studies have focused on quercetin and miRNAs modulation for therapeutic approaches. miR-200b-3p was up-regulated in pancreatic cancer cells when treated with quercetin, resulting in inhibition of self-renewal and decrease of proliferation through Notch1 signaling pathway [75]. Quercetin significantly inhibited breast cancer cell proliferation and invasion via up-regulated miR-146a expression and targeting EGFR [76]. Quercetin inhibited cell viability, migration and invasion by up-regulated miR-16 and targeting HOXA10 in oral cancer cells [77]. In addition, quercetin decreased oral cancer cell viability and increased cell apoptosis via miR-22/WNT1/β-catenin pathway [78].

Recently, quercetin modulated 34 miRNAs expression (5 upregulated and 29 downregulated) and novel miR-97, miR-298, miR-2218, miR-1502, and miR-2117 were identified in pheochromocytoma of the rat adrenal medulla that responded for protective effect against oxidative stress through PI3K-AKT signaling pathway [79]. Treatment of quercetin inhibited proliferation of endometriosis through up-regulated miR-503-5p, miR-1283, miR-3714 and miR-6867-5p by targeting CCND1 [80]. TGFβ1 is a fibrosis inducer and quercetin significantly down-regulated miR-21 and TGFβ1 and up-regulated miR-122 in liver fibrosis [81]. Protection of cardiomyocyte against hypoxia caused insults of quercetin has been reported by up-regulation of miR-199 mediated sirt1 expression and AMPK phosphorylation [82].

2.9 Silymarin

Silymarin is a flavonolignans extracted from the milk thistle Silybum marianum (L.) gaernt and recent studies have demonstrated the anti-cancer, anti-inflammatory, vascularization inhibitory, antioxidant, hepatoprotective, cardioprotective and anti-metastasis activities of silymarin [83]. Several miRNAs have been implicated in the invasive potential of cancer cells. Tumor suppressor miRNA, miR-203, was up-regulated and class 1 HDAC proteins and ZEB1 were repressed with silymarin treatment, resulted in inhibition of non-small cell lung cancer migration [84]. Silibinin, the major active constituent of silymarin extract, induced apoptosis and ERβ expression, inhibited cell proliferation, and reduced pro-inflammatory cytokines expression including IL-17 and TNF-α, through ERβ binding and down-regulated miR-155 in rheumatoid arthritis [85]. miR-122 is liver-specific miRNA and was down-regulation upon silymarin treatment in rat model for hepatoprotective and radio protective effects via increased superoxide dismutase (SOD), glutathione (GSH) and reduced lipid peroxidation (MDA) [86]. It has been reported the hepatoprotective activity of silymarin on thioacetamide-induced liver damage by restored miR-122, miR-192, and miR-194 expression levels [87].

2.10 β-Sitosterol-d-glucoside

β-Sitosterol-d-glucoside is bioactive compounds that has been isolated from Agave angustifolia and sweet potato [88, 89]. Pharmacological activity of β-Sitosterol-d-glucoside has been reported including immunomodulatory, anti-inflammatory, cytotoxic, and antiparasitic activities [88]. β-Sitosterol-d-glucoside exhibited cytotoxic effect in breast cancer cells by up-regulated miR-10a expression and decreased the PI3K/Akt signaling pathway [89]. Treatment of β-Sitosterol-d-glucoside can down-regulate miR-322-5p, miR-301a-3p, miR-129-5p, miR-322-3p, and miR-129-2-3p in neural stem cell and their targets are related to the regulation of proliferation [90]. Therefore, β-Sitosterol-d-glucoside could be developed for further therapeutic applications.

2.11 Sulforaphane

Sulforaphane is dietary compounds in broccoli (Brassica oleracea) and cruciferous plants. It has been demonstrated the capability of sulforaphane for anti-inflammatory, antiaging, antidiabetic, antioxidant, anti-tumor, hepatoprotective and cardioprotective effects [91]. Plant-derived phytochemicals including sulforaphane are potentially affected miRNAs expression. Sulforaphane inhibited breast cancer cell cycle arrest and senescence via down-regulation of miR-23b, miR-92b, miR-381 and miR-382 [92]. Anti-tumor effect of sulforaphane also reported in non-small cell lung cancer by down-regulation of miR-616-5p and targeting GSK3β/β-catenin signaling pathway [93]. Sulforaphane inhibited the progression of pancreatic cancer through down-regulated miR30a-3p with the increasing of its target, Cx43 expression and upregulated miR-135b-5p mediated RASAL2 expression [94, 95]. In addition, sulforaphane treatment significantly increased the expression of tumor suppressor miRNA, miR-200c, resulted in inhibited the cancer stemness and tumor-initiating properties in oral squamous cell carcinomas and cancer stem cells both in vitro and in vivo [96]. Anti-proliferative and apoptotic effects of sulforaphane have been reported in gastric cancer cells, which leading to alter the expression of miR-9 and miR-326 [97]. Up-regulation of miR-124-3p and inhibition of its target, STAT3 by sulforaphane treatment were observed and thereby induced apoptosis, inhibited proliferation and decreased the stemness of nasopharyngeal cancer cell [98].

Sulforaphane has potential to inhibit hepatic fibrosis by downregulating miR-423-5p in hepatic stellate cell [99]. Sulforaphane showed the protective effect in microglia-mediated neurotoxicity by inhibited LPS-induced expression of inflammatory miRNA, miR-155 [100].

Advertisement

3. Dietary miRNA and human gene regulation

Several evidences demonstrated the direct modulation of cellular signaling pathways by dietary compounds could decrease the risk of chronic diseases [101]. Interestingly, it has been reported that small non-coding RNA including miRNAs can be transferred across Kingdoms, for example dietary miRNAs have been found in human body fluids and these circulating miRNAs are likely to regulate human gene expression [15, 102, 103, 104, 105, 106, 107]. The uptake of plant derived miRNAs could be in the form of raw and cooked plants in capable of stability forms [107, 108]. Due to high temperature cooking process, low pH and enzymes in digestive tract as well as enzymes in blood circulation, miRNAs might be destroyed before their functions with target mRNAs [15]. Strikingly, GC base content, 2’-O-methylation on the 3′-terminal, unique nucleotide sequence of dietary miRNAs and extracellular vesicles (exosome and microvesicle) are preventive features of plant derived miRNAs in harmful conditions [109, 110, 111, 112, 113, 114].

There are numerous studies to support the functional roles of dietary miRNAs in cross kingdom gene regulation. Rice miR156a and miR168a were detected in human serum and miR168a down-regulated low-density lipoprotein receptor adapter protein 1 (LDLRAP1) expression, resulted in an increase of plasma LDL cholesterol level, Table 2 [105]. miR2910 from Populus euphratica was identified in human plasma and targeting Sprouty RTK Signaling Antagonist 4 (SPRY4) gene of the Janus kinase/signal transducers and activators of transcription (JAK–STAT) signaling pathway [115]. Based on the computationally predicted miRNAs from Camptotheca acuminate, 14 potential miRNAs were found to be regulated 152 target human genes such as miR4723–3p, miR5780d, and miR548d-3p targeting discs large MAGUK scaffoldprotein 2 (DLG2), NUMB endocytic adaptor protein (NUMB) and glycogen synthase kinase-3B (GSK3B) genes which were related to cancers such as breast cancer, lung cancer and leukemia [116]. Ocimum basilicum is a medicinal plant and its bioactive compounds have potential for therapeutic approaches. miRNA target prediction analysis revealed the target of O. basilicum miRNAs, miR156, miR531, miR160, miR529b, and miR1118 were 87 human target genes associated with the Ras-mitogen-activated protein kinase (Ras-MAPK) signaling pathway, Alzheimer disease, breast cancer, cardiomyopathy, HIV, lung cancer, and several neurological disorders [117].

PlantsPlant derived-miRNAsHuman target gene/ DiseaseReferences
Oryza sativaosa-miR156a
osa-miR166a
osa-miR168a
LDLRAP1[105]
Populus euphraticapeu-miR2910JAK–STAT pathway[115]
Camptotheca acuminata14 miRNAsCancer (breast, lung and leukemia)[116]
Ocimum basilicummiR156
miR531
miR160
miR529b
miR1118
Ras-MAPK signaling pathway, Alzheimer disease,
breast cancer, cardiomyopathy, HIV, lung cancer,
several neurological disorders
[117]
Curcuma longamiR14Rheumatoid arthritis[120]
cabbage, spinach and lettucemiR156aCardiovascular disease[118]
Oryza sativamiR156-5p
miR164-5p
miR168-5p
miR395-3p
miR396-3p
miR396-5p
miR444-3p
miR529-3p
miR1846-3p
miR2907-3p
Cancer, cardiovascular and neurodegenerative diseases[119]

Table 2.

Dietary miRNAs and human gene regulation.

The abundantly expressed miRNA in dietary green vegetable, miR156a which was detected in human serum and targeted the junction adhesion molecule-A (JAM-A) [118]. The JAM-A was up-regulated in atherosclerotic lesions from cardiovascular disease patients and miR156a could suppressed inflammatory cytokine-induced monocytes adhesion by targeting JAM-A [118]. The very recently report using a computational approach to predict the potential target of rice miRNAs including miR156-5p, miR164-5p, miR168-5p, miR395-3p, miR396-3p, miR396-5p, miR444-3p, miR529-3p, miR1846-3p, miR2907-3p, which can bind to the human mRNA [119]. Most of these target genes were associated with cancer, cardiovascular and neurodegenerative diseases [119]. miR14 derived from Curcuma longa was detected and remarkably stable in human serum for 48 h. The potential targets of miR14 were associated with inflammation in rheumatoid arthritis such as Phosphotidylinositol-specific-phospholipase C (PLCXD3), Adenylate cyclase 9 (ADCY9), and 3′ (2′), 5′-bisphosphate nucleotidase (BPNT1) [120].

Advertisement

4. Conclusion

It has been widely known that functional foods and their bioactive compounds have the capacity for human health benefits. To date, miRNAs have been shown a significant effect on gene expression and modulate the cellular biological functions in physiological and pathological conditions. There is emerging evidence suggesting that dietary bioactive compounds can be effective in human diseases as a result of altering miRNAs expression levels, resulting in modulation of cellular signaling pathway. Additional research the possibility of bioactive compounds for developing as novel drugs with less side effects is required in vitro and in vivo. Recently, it has been revealed in several studies that dietary derived-miRNAs are bioavailable and alter human gene expression. The cross-kingdom gene regulations of dietary miRNAs from plants to human have raised our expectations for evaluating the active therapeutic potential and dietary supplements.

Advertisement

Acknowledgments

This study has been supported by University of Phayao Research Grant (Grant no. FF64-RIM037) and School of Medical Sciences Research Grant (Grant no. MS 632001), University of Phayao.

References

  1. 1. Kura B, Parikh M, Slezak J, Pierce GN. The Influence of Diet on MicroRNAs that Impact Cardiovascular Disease. Molecules. 2019;24(8)
  2. 2. Calin GA, Sevignani C, Dumitru CD, Hyslop T, Noch E, Yendamuri S, et al. Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers. Proc Natl Acad Sci U S A. 2004;101(9):2999-3004
  3. 3. Quintanilha BJ, Reis BZ, Duarte GBS, Cozzolino SMF, Rogero MM. Nutrimiromics: Role of microRNAs and Nutrition in Modulating Inflammation and Chronic Diseases. Nutrients. 2017;9(11)
  4. 4. Pillai RS. MicroRNA function: multiple mechanisms for a tiny RNA? RNA. 2005;11(12):1753-61
  5. 5. Manach C, Hubert J, Llorach R, Scalbert A. The complex links between dietary phytochemicals and human health deciphered by metabolomics. Mol Nutr Food Res. 2009;53(10):1303-15
  6. 6. Singh D, Verma S, R. P. Investigations on Antioxidant Potential of Phenolic Acids and Flavonoids: The Common Phytochemical Ingredients in Plants. J Plant Biochem Physiol 2018;6(6):1-5
  7. 7. Koch W. Dietary Polyphenols-Important Non-Nutrients in the Prevention of Chronic Noncommunicable Diseases. A Systematic Review. Nutrients. 2019;11(5)
  8. 8. Ghasemzadeh A, Ghasemzadeh N. Flavonoids and phenolic acids: Role and biochemical activity in plants and human. J Med Plant Res. 2011;5(31):6697-703
  9. 9. Kocic H, Damiani G, Stamenkovic B, Tirant M, Jovic A, Tiodorovic D, et al. Dietary compounds as potential modulators of microRNA expression in psoriasis. Ther Adv Chronic Dis. 2019;10:2040622319864805
  10. 10. Cui J, Zhou B, Ross SA, Zempleni J. Nutrition, microRNAs, and Human Health. Adv Nutr. 2017;8(1):105-12
  11. 11. Ross SA, Davis CD. The emerging role of microRNAs and nutrition in modulating health and disease. Annu Rev Nutr. 2014;34:305-36
  12. 12. Gavrilas LI, Ionescu C, Tudoran O, Lisencu C, Balacescu O, Miere D. The Role of Bioactive Dietary Components in Modulating miRNA Expression in Colorectal Cancer. Nutrients. 2016;8(10)
  13. 13. Xie W, Weng A, Melzig MF. MicroRNAs as New Bioactive Components in Medicinal Plants. Planta Med. 2016;82(13):1153-62
  14. 14. Zempleni J, Baier SR, Howard KM, Cui J. Gene regulation by dietary microRNAs. Can J Physiol Pharmacol. 2015;93(12):1097-102
  15. 15. Li Z, Xu R, Li N. MicroRNAs from plants to animals, do they define a new messenger for communication? Nutr Metab (Lond). 2018;15:68
  16. 16. Kang H. MicroRNA-Mediated Health-Promoting Effects of Phytochemicals. Int J Mol Sci. 2019;20(10)
  17. 17. Son SW, Lee HY, Moeng S, Kuh HJ, Choi SY, Park JK. Participation of MicroRNAs in the Treatment of Cancer with Phytochemicals. Molecules. 2020;25(20)
  18. 18. Jin L, Yingchun W, Zhujun S, Yinan W, Dongchen W, Hui Y, et al. 3-acetyl-11-keto-beta-boswellic acid decreases the malignancy of taxol resistant human ovarian cancer by inhibiting multidrug resistance (MDR) proteins function. Biomed Pharmacother. 2019;116:108992
  19. 19. Roy NK, Parama D, Banik K, Bordoloi D, Devi AK, Thakur KK, et al. An Update on Pharmacological Potential of Boswellic Acids against Chronic Diseases. Int J Mol Sci. 2019;20(17)
  20. 20. Sun MX, He XP, Huang PY, Qi Q , Sun WH, Liu GS, et al. Acetyl-11-keto-beta-boswellic acid inhibits proliferation and induces apoptosis of gastric cancer cells through the phosphatase and tensin homolog /Akt/ cyclooxygenase-2 signaling pathway. World J Gastroenterol. 2020;26(38):5822-35
  21. 21. Sayed AS, Gomaa IEO, Bader M, El Sayed N. Role of 3-Acetyl-11-Keto-Beta-Boswellic Acid in Counteracting LPS-Induced Neuroinflammation via Modulation of miRNA-155. Mol Neurobiol. 2018;55(7):5798-808
  22. 22. Jiang X, Liu Y, Zhang G, Lin S, Yuan N, Wu J, et al. Acetyl-11-keto-beta-boswellic Acid Inhibits Precancerous Breast Lesion MCF-10AT Cells via Regulation of LINC00707/miR-206 that Reduces Estrogen Receptor-alpha. Cancer Manag Res. 2020;12:2301-14
  23. 23. Toden S, Okugawa Y, Buhrmann C, Nattamai D, Anguiano E, Baldwin N, et al. Novel Evidence for Curcumin and Boswellic Acid-Induced Chemoprevention through Regulation of miR-34a and miR-27a in Colorectal Cancer. Cancer Prev Res (Phila). 2015;8(5):431-43
  24. 24. He Y, Fan Q , Cai T, Huang W, Xie X, Wen Y, et al. Molecular mechanisms of the action of Arctigenin in cancer. Biomed Pharmacother. 2018;108:403-7
  25. 25. Wang P, Solorzano W, Diaz T, Magyar CE, Henning SM, Vadgama JV. Arctigenin inhibits prostate tumor cell growth in vitro and in vivo. Clin Nutr Exp. 2017;13:1-11
  26. 26. Gao Q , Yang M, Zuo Z. Overview of the anti-inflammatory effects, pharmacokinetic properties and clinical efficacies of arctigenin and arctiin from Arctium lappa L. Acta Pharmacol Sin. 2018;39(5):787-801
  27. 27. Hayashi K, Narutaki K, Nagaoka Y, Hayashi T, Uesato S. Therapeutic effect of arctiin and arctigenin in immunocompetent and immunocompromised mice infected with influenza A virus. Biol Pharm Bull. 2010;33(7):1199-205
  28. 28. Song J, Li N, Xia Y, Gao Z, Zou SF, Yan YH, et al. Arctigenin Confers Neuroprotection Against Mechanical Trauma Injury in Human Neuroblastoma SH-SY5Y Cells by Regulating miRNA-16 and miRNA-199a Expression to Alleviate Inflammation. J Mol Neurosci. 2016;60(1):115-29
  29. 29. Wang P, Phan T, Gordon D, Chung S, Henning SM, Vadgama JV. Arctigenin in combination with quercetin synergistically enhances the antiproliferative effect in prostate cancer cells. Mol Nutr Food Res. 2015;59(2):250-61
  30. 30. Ruwizhi N, Aderibigbe BA. Cinnamic Acid Derivatives and Their Biological Efficacy. Int J Mol Sci. 2020;21(16)
  31. 31. Kumazaki M, Shinohara H, Taniguchi K, Yamada N, Ohta S, Ichihara K, et al. Propolis cinnamic acid derivatives induce apoptosis through both extrinsic and intrinsic apoptosis signaling pathways and modulate of miRNA expression. Phytomedicine. 2014;21(8-9):1070-7
  32. 32. Hunke M, Martinez W, Kashyap A, Bokoskie T, Pattabiraman M, Chandra S. Antineoplastic Actions of Cinnamic Acids and Their Dimers in Breast Cancer Cells: A Comparative Study. Anticancer Res. 2018;38(8):4469-74
  33. 33. Li S, Hu S. Cinnamic hydroxamic acid inhibits the proliferation of gastric cancer cells via upregulation of miR 145 expression and down-regulation of P13K/Akt signaling pathway. Trop J Pharm Res. 2020;19(5):957-63
  34. 34. Gupta SC, Patchva S, Koh W, Aggarwal BB. Discovery of curcumin, a component of golden spice, and its miraculous biological activities. Clin Exp Pharmacol Physiol. 2012;39(3):283-99
  35. 35. Liu Y, Sun H, Makabel B, Cui Q , Li J, Su C, et al. The targeting of noncoding RNAs by curcumin: Facts and hopes for cancer therapy (Review). Oncol Rep. 2019;42(1):20-34
  36. 36. Su J, Zhou X, Wang L, Yin X, Wang Z. Curcumin inhibits cell growth and invasion and induces apoptosis through down-regulation of Skp2 in pancreatic cancer cells. Am J Cancer Res. 2016;6(9):1949-62
  37. 37. Mirzaei H, Masoudifar A, Sahebkar A, Zare N, Sadri Nahand J, Rashidi B, et al. MicroRNA: A novel target of curcumin in cancer therapy. J Cell Physiol. 2018;233(4):3004-15
  38. 38. Norouzi S, Majeed M, Pirro M, Generali D, Sahebkar A. Curcumin as an Adjunct Therapy and microRNA Modulator in Breast Cancer. Curr Pharm Des. 2018;24(2):171-7
  39. 39. Toden S, Okugawa Y, Jascur T, Wodarz D, Komarova NL, Buhrmann C, et al. Curcumin mediates chemosensitization to 5-fluorouracil through miRNA-induced suppression of epithelial-to-mesenchymal transition in chemoresistant colorectal cancer. Carcinogenesis. 2015;36(3):355-67
  40. 40. Mudduluru G, George-William JN, Muppala S, Asangani IA, Kumarswamy R, Nelson LD, et al. Curcumin regulates miR-21 expression and inhibits invasion and metastasis in colorectal cancer. Biosci Rep. 2011;31(3):185-97
  41. 41. Liu W, Huang M, Zou Q , Lin W. Curcumin suppresses gastric cancer biological activity by regulation of miRNA-21: an in vitro study. Int J Clin Exp Pathol. 2018;11(12):5820-9
  42. 42. Liu T, Chi H, Chen J, Chen C, Huang Y, Xi H, et al. Curcumin suppresses proliferation and in vitro invasion of human prostate cancer stem cells by ceRNA effect of miR-145 and lncRNA-ROR. Gene. 2017;631:29-38
  43. 43. Kronski E, Fiori ME, Barbieri O, Astigiano S, Mirisola V, Killian PH, et al. miR181b is induced by the chemopreventive polyphenol curcumin and inhibits breast cancer metastasis via down-regulation of the inflammatory cytokines CXCL1 and −2. Mol Oncol. 2014;8(3):581-95
  44. 44. Li W, Yang W, Liu Y, Chen S, Chin S, Qi X, et al. MicroRNA-378 enhances inhibitory effect of curcumin on glioblastoma. Oncotarget. 2017;8(43):73938-46
  45. 45. Abdulkader M, Zaky A, Kandeel K, Bassiouny A. Synergistic Neuroprotective Effect of Curcumin and Pioglitazone Against Intranigral LPS-Induced Sub-Acute Neurodegeneration in Rat. EC Neurology. 2019;11(12):111-28
  46. 46. Biersack B. 3,3’-Diindolylmethane and its derivatives: nature-inspired strategies tackling drug resistant tumors by regulation of signal transduction, transcription factors and microRNAs. Cancer Drug Resist 2020;3:1-12
  47. 47. Ahmad A, Ali S, Ahmed A, Ali AS, Raz A, Sakr WA, et al. 3, 3'-Diindolylmethane enhances the effectiveness of herceptin against HER-2/neu-expressing breast cancer cells. PLoS One. 2013;8(1):e54657
  48. 48. Hanieh H. Aryl hydrocarbon receptor-microRNA-212/132 axis in human breast cancer suppresses metastasis by targeting SOX4. Mol Cancer. 2015;14:172
  49. 49. Jin Y, Zou X, Feng X. 3,3'-Diindolylmethane negatively regulates Cdc25A and induces a G2/M arrest by modulation of microRNA 21 in human breast cancer cells. Anticancer Drugs. 2010;21(9):814-22
  50. 50. Li Y, VandenBoom TG, 2nd, Kong D, Wang Z, Ali S, Philip PA, et al. Up-regulation of miR-200 and let-7 by natural agents leads to the reversal of epithelial-to-mesenchymal transition in gemcitabine-resistant pancreatic cancer cells. Cancer Res. 2009;69(16):6704-12
  51. 51. Li Y, Vandenboom TG, 2nd, Wang Z, Kong D, Ali S, Philip PA, et al. miR-146a suppresses invasion of pancreatic cancer cells. Cancer Res. 2010;70(4):1486-95
  52. 52. Tomar S, Nagarkatti M, Nagarkatti PS. 3,3'-Diindolylmethane attenuates LPS-mediated acute liver failure by regulating miRNAs to target IRAK4 and suppress Toll-like receptor signalling. Br J Pharmacol. 2015;172(8):2133-47
  53. 53. Rouse M, Rao R, Nagarkatti M, Nagarkatti PS. 3,3′-diindolylmethane ameliorates experimental autoimmune encephalomyelitis by promoting cell cycle arrest and apoptosis in activated T cells through microRNA signaling pathways. J Pharmacol Exp Ther. 2014;350(2):341-52
  54. 54. Sharifi-Rad M, Pezzani R, Redaelli M, Zorzan M, Imran M, Ahmed Khalil A, et al. Preclinical Pharmacological Activities of Epigallocatechin-3-gallate in Signaling Pathways: An Update on Cancer. Molecules. 2020;25(3)
  55. 55. Huang YJ, Wang KL, Chen HY, Chiang YF, Hsia SM. Protective Effects of Epigallocatechin Gallate (EGCG) on Endometrial, Breast, and Ovarian Cancers. Biomolecules. 2020;10(11)
  56. 56. Negri A, Naponelli V, Rizzi F, Bettuzzi S. Molecular Targets of Epigallocatechin-Gallate (EGCG): A Special Focus on Signal Transduction and Cancer. Nutrients. 2018;10(12)
  57. 57. Lin CH, Wang HH, Chen TH, Chiang MC, Hung PH, Chen YJ. Involvement of MicroRNA-296 in the Inhibitory Effect of Epigallocatechin Gallate against the Migratory Properties of Anoikis-Resistant Nasopharyngeal Carcinoma Cells. Cancers (Basel). 2020;12(4)
  58. 58. Mostafa SM, Gamal-Eldeen AM, Maksoud NAE, Fahmi AA. Epigallocatechin gallate-capped gold nanoparticles enhanced the tumor suppressors let-7a and miR-34a in hepatocellular carcinoma cells. An Acad Bras Cienc. 2020;92(4):e20200574
  59. 59. Mokhtari H, Yaghmaei B, Sirati-Sabet M, Jafari N, Mardomi A, Abediankenari S, et al. Epigallocatechin-3-gallate Enhances the Efficacy of MicroRNA-34a Mimic and MicroRNA-93 Inhibitor Co-transfection in Prostate Cancer Cell Line. Iran J Allergy Asthma Immunol 2020;19(6):612-23
  60. 60. Zhu Y, Huang Y, Liu M, Yan Q , Zhao W, Yang P, et al. Epigallocatechin gallate inhibits cell growth and regulates miRNA expression in cervical carcinoma cell lines infected with different high-risk human papillomavirus subtypes. Exp Ther Med. 2019;17(3):1742-8
  61. 61. Yamada S, Tsukamoto S, Huang Y, Makio A, Kumazoe M, Yamashita S, et al. Epigallocatechin-3-O-gallate up-regulates microRNA-let-7b expression by activating 67-kDa laminin receptor signaling in melanoma cells. Sci Rep. 2016;6:19225
  62. 62. Zhang C, Liang R, Gan X, Yang X, Chen L, Jian J. MicroRNA-384-5p/Beclin-1 As Potential Indicators For Epigallocatechin Gallate Against Cardiomyocytes Ischemia Reperfusion Injury By Inhibiting Autophagy Via PI3K/Akt Pathway. Drug Des Devel Ther. 2019;13:3607-23
  63. 63. Rasheed Z, Rasheed N, Al-Shaya O. Epigallocatechin-3-O-gallate modulates global microRNA expression in interleukin-1beta-stimulated human osteoarthritis chondrocytes: potential role of EGCG on negative co-regulation of microRNA-140-3p and ADAMTS5. Eur J Nutr. 2018;57(3):917-28
  64. 64. Arffa ML, Zapf MA, Kothari AN, Chang V, Gupta GN, Ding X, et al. Epigallocatechin-3-Gallate Upregulates miR-221 to Inhibit Osteopontin-Dependent Hepatic Fibrosis. PLoS One. 2016;11(12):e0167435
  65. 65. Tuli HS, Tuorkey MJ, Thakral F, Sak K, Kumar M, Sharma AK, et al. Molecular Mechanisms of Action of Genistein in Cancer: Recent Advances. Front Pharmacol. 2019;10:1336
  66. 66. Avci CB, Susluer SY, Caglar HO, Balci T, Aygunes D, Dodurga Y, et al. Genistein-induced mir-23b expression inhibits the growth of breast cancer cells. Contemp Oncol (Pozn). 2015;19(1):32-5
  67. 67. Hirata H, Hinoda Y, Shahryari V, Deng G, Tanaka Y, Tabatabai ZL, et al. Correction: Genistein downregulates onco-miR-1260b and upregulates sFRP1 and Smad4 via demethylation and histone modification in prostate cancer cells. Br J Cancer. 2018;119(3):388
  68. 68. Hirata H, Ueno K, Nakajima K, Tabatabai ZL, Hinoda Y, Ishii N, et al. Genistein downregulates onco-miR-1260b and inhibits Wnt-signalling in renal cancer cells. Br J Cancer. 2013;108(10):2070-8
  69. 69. Yang Y, Zang A, Jia Y, Shang Y, Zhang Z, Ge K, et al. Genistein inhibits A549 human lung cancer cell proliferation via miR-27a and MET signaling. Oncol Lett. 2016;12(3):2189-93
  70. 70. Sacko K, Thangavel K, Shoyele SA. Codelivery of Genistein and miRNA-29b to A549 Cells Using Aptamer-Hybrid Nanoparticle Bioconjugates. Nanomaterials (Basel). 2019;9(7)
  71. 71. Gan M, Yang D, Fan Y, Du J, Shen L, Li Q , et al. Bidirectional regulation of genistein on the proliferation and differentiation of C2C12 myoblasts. Xenobiotica. 2020;50(11):1352-8
  72. 72. Gan M, Shen L, Fan Y, Tan Y, Zheng T, Tang G, et al. MicroRNA-451 and Genistein Ameliorate Nonalcoholic Steatohepatitis in Mice. Int J Mol Sci. 2019;20(23)
  73. 73. Kim DH, Khan H, Ullah H, Hassan STS, Smejkal K, Efferth T, et al. MicroRNA targeting by quercetin in cancer treatment and chemoprotection. Pharmacol Res. 2019;147:104346
  74. 74. Akbari Kordkheyli V, Khonakdar Tarsi A, Mishan MA, Tafazoli A, Bardania H, Zarpou S, et al. Effects of quercetin on microRNAs: A mechanistic review. J Cell Biochem. 2019;120(8):12141-55
  75. 75. Nwaeburu CC, Abukiwan A, Zhao Z, Herr I. Quercetin-induced miR-200b-3p regulates the mode of self-renewing divisions in pancreatic cancer. Mol Cancer. 2017;16(1):23
  76. 76. Tao SF, He HF, Chen Q . Quercetin inhibits proliferation and invasion acts by up-regulating miR-146a in human breast cancer cells. Mol Cell Biochem. 2015;402(1-2):93-100
  77. 77. Zhao J, Fang Z, Zha Z, Sun Q , Wang H, Sun M, et al. Quercetin inhibits cell viability, migration and invasion by regulating miR-16/HOXA10 axis in oral cancer. Eur J Pharmacol. 2019;847:11-8
  78. 78. Zhang C, Hao Y, Sun Y, Liu P. Quercetin suppresses the tumorigenesis of oral squamous cell carcinoma by regulating microRNA-22/WNT1/beta-catenin axis. J Pharmacol Sci. 2019;140(2):128-36
  79. 79. Zhang Z, Yi P, Yi M, Tong X, Cheng X, Yang J, et al. Protective Effect of Quercetin against H2O2-Induced Oxidative Damage in PC-12 Cells: Comprehensive Analysis of a lncRNA-Associated ceRNA Network. Oxid Med Cell Longev. 2020;2020:6038919
  80. 80. Park S, Lim W, Bazer FW, Whang KY, Song G. Quercetin inhibits proliferation of endometriosis regulating cyclin D1 and its target microRNAs in vitro and in vivo. J Nutr Biochem. 2019;63:87-100
  81. 81. Nozari E, Moradi A, Samadi M. Effect of Atorvastatin, Curcumin, and Quercetin on miR-21 and miR-122 and their correlation with TGFbeta1 expression in experimental liver fibrosis. Life Sci. 2020;259:118293
  82. 82. Guo G, Gong L, Sun L, Xu H. Quercetin supports cell viability and inhibits apoptosis in cardiocytes by down-regulating miR-199a. Artif Cells Nanomed Biotechnol. 2019;47(1):2909-16
  83. 83. Kim SH, Choo GS, Yoo ES, Woo JS, Han SH, Lee JH, et al. Silymarin induces inhibition of growth and apoptosis through modulation of the MAPK signaling pathway in AGS human gastric cancer cells. Oncol Rep. 2019;42(5):1904-14
  84. 84. Singh T, Prasad R, Katiyar SK. Therapeutic intervention of silymarin on the migration of non-small cell lung cancer cells is associated with the axis of multiple molecular targets including class 1 HDACs, ZEB1 expression, and restoration of miR-203 and E-cadherin expression. Am J Cancer Res. 2016;6(6):1287-301
  85. 85. Dupuis ML, Conti F, Maselli A, Pagano MT, Ruggieri A, Anticoli S, et al. The Natural Agonist of Estrogen Receptor beta Silibinin Plays an Immunosuppressive Role Representing a Potential Therapeutic Tool in Rheumatoid Arthritis. Front Immunol. 2018;9:1903
  86. 86. Abdelmageed Marzook E, Abdel-Aziz AF, Abd El-Moneim AE, Mansour HA, Atia KS, Salah NA. MicroRNA-122 expression in hepatotoxic and γ-irradiated rats pre-treated with naringin and silymarin. J Radiat Res Appl Sci. 2019;13(1):47-55
  87. 87. Teksoy O, Sahinturk V, Cengiz M, Inal B, Ayhanci A. The Protective Effects of Silymarin on Thioacetamide-Induced Liver Damage: Measurement of miR-122, miR-192, and miR-194 Levels. Appl Biochem Biotechnol. 2020;191(2):528-39
  88. 88. Lopez-Salazar H, Camacho-Diaz BH, Avila-Reyes SV, Perez-Garcia MD, Gonzalez-Cortazar M, Arenas Ocampo ML, et al. Identification and Quantification of beta-Sitosterol beta-d-Glucoside of an Ethanolic Extract Obtained by Microwave-Assisted Extraction from Agave angustifolia Haw. Molecules. 2019;24(21)
  89. 89. Xu H, Li Y, Han B, Li Z, Wang B, Jiang P, et al. Anti-breast-Cancer Activity Exerted by beta-Sitosterol-d-glucoside from Sweet Potato via Upregulation of MicroRNA-10a and via the PI3K-Akt Signaling Pathway. J Agric Food Chem. 2018;66(37):9704-18
  90. 90. Jiang LH, Yang NY, Yuan XL, Zou YJ, Jiang ZQ , Zhao FM, et al. Microarray Analysis of mRNA and MicroRNA Expression Profile Reveals the Role of beta -Sitosterol-D-glucoside in the Proliferation of Neural Stem Cell. Evid Based Complement Alternat Med. 2013;2013:360302
  91. 91. Rafiei H, Ashrafizadeh M, Ahmadi Z. MicroRNAs as novel targets of sulforaphane in cancer therapy: The beginning of a new tale? Phytother Res. 2020;34(4):721-8
  92. 92. Lewinska A, Adamczyk-Grochala J, Deregowska A, Wnuk M. Sulforaphane-Induced Cell Cycle Arrest and Senescence are accompanied by DNA Hypomethylation and Changes in microRNA Profile in Breast Cancer Cells. Theranostics. 2017;7(14):3461-77
  93. 93. Wang DX, Zou YJ, Zhuang XB, Chen SX, Lin Y, Li WL, et al. Sulforaphane suppresses EMT and metastasis in human lung cancer through miR-616-5p-mediated GSK3beta/beta-catenin signaling pathways. Acta Pharmacol Sin. 2017;38(2):241-51
  94. 94. Yin L, Xiao X, Georgikou C, Luo Y, Liu L, Gladkich J, et al. Sulforaphane Induces miR135b-5p and Its Target Gene, RASAL2, thereby Inhibiting the Progression of Pancreatic Cancer. Mol Ther Oncolytics. 2019;14:74-81
  95. 95. Georgikou C, Yin L, Gladkich J, Xiao X, Sticht C, Torre C, et al. Inhibition of miR30a-3p by sulforaphane enhances gap junction intercellular communication in pancreatic cancer. Cancer Lett. 2020;469:238-45
  96. 96. Liu CM, Peng CY, Liao YW, Lu MY, Tsai ML, Yeh JC, et al. Sulforaphane targets cancer stemness and tumor initiating properties in oral squamous cell carcinomas via miR-200c induction. J Formos Med Assoc. 2017;116(1):41-8
  97. 97. Kiani S, Akhavan-Niaki H, Fattahi S, Kavoosian S, Babaian Jelodar N, Bagheri N, et al. Purified sulforaphane from broccoli (Brassica oleracea var. italica) leads to alterations of CDX1 and CDX2 expression and changes in miR-9 and miR-326 levels in human gastric cancer cells. Gene. 2018;678:115-23
  98. 98. Li X, Zhao Z, Li M, Liu M, Bahena A, Zhang Y, et al. Sulforaphane promotes apoptosis, and inhibits proliferation and self-renewal of nasopharyngeal cancer cells by targeting STAT signal through miRNA-124-3p. Biomed Pharmacother. 2018;103:473-81
  99. 99. Feng MH, Li JW, Sun HT, He SQ , Pang J. Sulforaphane inhibits the activation of hepatic stellate cell by miRNA-423-5p targeting suppressor of fused. Hum Cell. 2019;32(4):403-10
  100. 100. Eren E, Tufekci KU, Isci KB, Tastan B, Genc K, Genc S. Sulforaphane Inhibits Lipopolysaccharide-Induced Inflammation, Cytotoxicity, Oxidative Stress, and miR-155 Expression and Switches to Mox Phenotype through Activating Extracellular Signal-Regulated Kinase 1/2-Nuclear Factor Erythroid 2-Related Factor 2/Antioxidant Response Element Pathway in Murine Microglial Cells. Front Immunol. 2018;9:36
  101. 101. Marzano F, Caratozzolo MF, Consiglio A, Licciulli F, Liuni S, Sbisa E, et al. Plant miRNAs Reduce Cancer Cell Proliferation by Targeting MALAT1 and NEAT1: A Beneficial Cross-Kingdom Interaction. Front Genet. 2020;11:552490
  102. 102. Minutolo A, Potesta M, Gismondi A, Pirro S, Cirilli M, Gattabria F, et al. Olea europaea small RNA with functional homology to human miR34a in cross-kingdom interaction of anti-tumoral response. Sci Rep. 2018;8(1):12413
  103. 103. Pastrello C, Tsay M, McQuaid R, Abovsky M, Pasini E, Shirdel E, et al. Circulating plant miRNAs can regulate human gene expression in vitro. Sci Rep. 2016;6:32773
  104. 104. Link J, Thon C, Schanze D, Steponaitiene R, Kupcinskas J, Zenker M, et al. Food-Derived Xeno-microRNAs: Influence of Diet and Detectability in Gastrointestinal Tract-Proof-of-Principle Study. Mol Nutr Food Res. 2019;63(2):e1800076
  105. 105. Sanchita, Trivedi R, Asif MH, Trivedi PK. Dietary plant miRNAs as an augmented therapy: cross-kingdom gene regulation. RNA Biol. 2018;15(12):1433-9
  106. 106. Liang G, Zhu Y, Sun B, Shao Y, Jing A, Wang J, et al. Assessing the survival of exogenous plant microRNA in mice. Food Sci Nutr. 2014;2(4):380-8
  107. 107. Zhang L, Hou D, Chen X, Li D, Zhu L, Zhang Y, et al. Exogenous plant MIR168a specifically targets mammalian LDLRAP1: evidence of cross-kingdom regulation by microRNA. Cell Res. 2012;22(1):107-26
  108. 108. Chin AR, Fong MY, Somlo G, Wu J, Swiderski P, Wu X, et al. Cross-kingdom inhibition of breast cancer growth by plant miR159. Cell Res. 2016;26(2):217-28
  109. 109. Voinnet O. Origin, biogenesis, and activity of plant microRNAs. Cell. 2009;136(4):669-87
  110. 110. Zhao Y, Mo B, Chen X. Mechanisms that impact microRNA stability in plants. RNA Biol. 2012;9(10):1218-23
  111. 111. Liang H, Jiao Z, Rong W, Qu S, Liao Z, Sun X, et al. 3'-Terminal 2'-O-methylation of lung cancer miR-21-5p enhances its stability and association with Argonaute 2. Nucleic Acids Res. 2020;48(13):7027-40
  112. 112. Xiao J, Feng S, Wang X, Long K, Luo Y, Wang Y, et al. Identification of exosome-like nanoparticle-derived microRNAs from 11 edible fruits and vegetables. PeerJ. 2018;6:e5186
  113. 113. Woith E, Fuhrmann G, Melzig MF. Extracellular Vesicles-Connecting Kingdoms. Int J Mol Sci. 2019;20(22)
  114. 114. Rome S. Biological properties of plant-derived extracellular vesicles. Food Funct. 2019;10(2):529-38
  115. 115. Liu YC, Chen WL, Kung WH, Huang HD. Plant miRNAs found in human circulating system provide evidences of cross kingdom RNAi. BMC Genomics. 2017;18(Suppl 2):112
  116. 116. Kumar D, Kumar S, Ayachit G, Bhairappanavar SB, Ansari A, Sharma P, et al. Cross-Kingdom Regulation of Putative miRNAs Derived from Happy Tree in Cancer Pathway: A Systems Biology Approach. Int J Mol Sci. 2017;18(6)
  117. 117. Patel M, Mangukia N, Jha N, Gadhavi H, Shah K, Patel S, et al. Computational identification of miRNA and their cross kingdom targets from expressed sequence tags of Ocimum basilicum. Mol Biol Rep. 2019;46(3):2979-95
  118. 118. Hou D, He F, Ma L, Cao M, Zhou Z, Wei Z, et al. The potential atheroprotective role of plant MIR156a as a repressor of monocyte recruitment on inflamed human endothelial cells. J Nutr Biochem. 2018;57:197-205
  119. 119. Rakhmetullina A, Pyrkova A, Aisina D, Ivashchenko A. HUMAN GENES ARE IN SILICO POTENTIAL TARGETS FOR RICE miRNA bioRxiv preprint. 2020:1-26
  120. 120. Sharma A, Sahu S, Kumari P, Gopi SR, Malhotra R, Biswas S. Genome-wide identification and functional annotation of miRNAs in anti-inflammatory plant and their cross-kingdom regulation in Homo sapiens. J Biomol Struct Dyn. 2017;35(7):1389-400

Written By

Wittaya Chaiwangyen

Submitted: 26 October 2020 Reviewed: 19 February 2021 Published: 08 April 2021