Milk exosomal microRNAs: friend or foe?—a narrative review
Review Article

Milk exosomal microRNAs: friend or foe?—a narrative review

Bodo C. Melnik1, Ralf Weiskirchen2, Gerd Schmitz3

1Department of Dermatology, Environmental Medicine and Health Theory, University of Osnabrück, Osnabrück, Germany; 2Institute of Molecular Pathobiochemistry, Experimental Gene Therapy and Clinical Chemistry (IFMPEGKC), RWTH University Hospital Aachen, Aachen, Germany; 3Institute for Clinical Chemistry and Laboratory Medicine, University Hospital of Regensburg, University of Regensburg, Regensburg, Germany

Contributions: (I) Conception and design: BC Melnik; (II) Administrative support: G Schmitz, R Weiskirchen; (III) Provision of study materials or patients: None; (IV) Collection and assembly of data: BC Melnik; (V) Data analysis and interpretation: All authors; (VI) Manuscript writing: All authors; (VII) Final approval of manuscript: All authors.

Correspondence to: Bodo C. Melnik. Department of Dermatology, Environmental Medicine and Health Theory, University of Osnabrück, Am Finkenhügel 7a, D-49076 Osnabrück, Germany. Email: melnik@t-online.de.

Background and Objective: Milk exosomes (MEX) and their inherent microRNAs (miRs) were recently promoted as promising effectors and drug carrier systems for the treatment of various chronic human diseases. This review intends to provide a comprehensive view on the potential beneficial and adverse health effects of MEX miRs.

Methods: English literature published between 1990 and 2022 were searched using the PubMed database focusing on publications including human and bovine MEX, milk extracellular vesicles (EVs), miRs, and reported beneficial and adverse effects of MEX miRs.

Key Content and Findings: MEX are regarded as signalosomes produced under control of the lactation genome to support species-specific growth of the offspring during the lactation period. Physiologically, mammals are not exposed to MEX miRs after the lactation period. MEX miRs are important for epigenetic postnatal programming and tissue maturation. Direct and translational evidence indicates that MEX miRs are involved in p53-mediated transcription, DNA methyltransferase-regulated gene silencing as well as Polycomb repressive complex-mediated gene repression and chromatin remodeling. MEX miRs are deficient in artificial formula, but may accidentally modify human gene expression by consumption of pasteurized cow milk. The continuous impact of bovine MEX miRs on the human epigenome and epitranscriptome remains a matter of concern.

Conclusions: MEX miRs are supportive for the growing neonate, whereas continuous exposure of adult humans to bovine MEX miRs may promote obesity, diabetes, cancer and neurodegeneration. Bovine MEX miRs should be removed from the human food chain of adults. The efficacy and long-term safety of MEX miRs should be carefully assessed in future studies before bovine MEX could be introduced for therapeutic purposes.

Keywords: Cancer; diabetes; milk exosomal microRNAs (miRs); transcription; health-risk evaluation


Received: 28 March 2022; Accepted: 18 August 2022; Published: 31 October 2022.

doi: 10.21037/exrna-22-5


Introduction

Milk contains abundant large extracellular vesicles (EVs) (1-5) and small nano-sized EVs (30–150 nm) known as milk exosomes (MEX), that transfer microRNAs (miRs), long noncoding RNAs (lncRNAs), circular RNAs (circRNAs), messenger RNAs (mRNAs) among others (6-22). MEX miRs are highly conserved between mammals (11,23), survive the harsh conditions of the gastrointestinal tract (24-26), are taken up by endocytosis (27,28), are bioavailable (29,30), reach the systemic circulation (29,31) and enter cells and peripheral tissues (30,32-36).

Due to the possibility of high yield preparations of bovine MEX (37-39) and their ability to overcome tissue boundaries (30,32-35), MEX are regarded as an ideal nanoplatform for the transfer of either intrinsic miRs (40-55) or as carrier systems for drug or anti-sense RNA transfer to target tissues (34,56-61) (Table 1).

Table 1

Potential therapeutic effects of MEX in experimental models

MEX source Disease—pathological condition Reference
Bovine Amelioration of experimental arthritis (40)
Bovine Inhibition of catabolic and inflammatory processes in cartilage from osteoarthritis patients (41)
Bovine Osteoprotective effect by increasing osteocyte numbers and targeting RANKL/OPG system in experimental models of bone loss (42)
Bovine Induction of proliferation and differentiation of osteoblasts and osteogenesis (43)
Porcine Protection against intestinal epithelial cell damage (44)
Porcine Inhibition LPS-induced intestinal epithelial cell apoptosis (45)
Bovine Enhanced goblet cell activity and prevention of the development of experimental necrotizing enterocolitis (46)
Bovine Attenuation of experimental ulcerative colitis (47)
Bovine Attenuation of experimental colitis (48)
Human, bovine Attenuation of experimental colitis (49)
Bovine Amelioration of cardiac fibrosis (50)
Bovine Potential attenuation of hepatic fibrosis (51)
Bovine Regeneration and acceleration of cutaneous wound healing (52)
Bovine Modulation of scar-free wound healing (53)
Bovine Acceleration of angiogenesis and diabetic wound healing (54)

MEX, milk exosome; LPS, lipopolysaccharide.

Despite these opportunities, the question remains: Is the transfer of MEX miRs beneficial or do they exert adverse health effects? To answer this question, the original biological purpose of MEX miRs and their physiological functions during the breastfeeding period have to be examined in more detail.


Methods

Literature research was performed using the PubMed database between 1990 and 2022 selecting original research papers published in English language including the search items: human and bovine milk exosome, milk extracellular vesicle, microRNA (miRNA, miR), gene expression, gene regulation, beneficial health effects, adverse health effects. Original data of reported biological effects of miRs were extracted and provided in the corresponding disease section. Literature data of human and bovine MEX-derived miRs were inspected and related to their potential health effects (Table 2).

Table 2

The search strategy summary

Items Specification
Date of search Jul 2022
Databases and other sources searched PubMed
Search terms used Milk exosome, extracellular vesicle, miR, gene regulation, gene expression, health effects
Timeframe 1990–2022
Inclusion and exclusion criteria Inclusion: original English research articles;
Exclusion: articles repeating published data
Selection process First author conducted paper selection

Outcome of literature research

Physiological versus artificial MEX miR signaling

The perception of milk as a food for infants (62) has changed to a complex maternal-neonatal signaling system promoting postnatal growth and appropriate metabolic, adipogenic, immunological and neuronal programming of the infant (12,63-72). MEX represent maternal signalosomes relaying gene-regulatory communication between the maternal lactation genome and the infant (69). They have an important impact on epigenetic postnatal programming (73,74) and early-life growth trajectories (75). Notably, miR-148a, the major miR of human and bovine MEX (7,23,76), exhibits an identical nucleotide sequence between Homo sapiens and Bos taurus (11). MEX miRs modify gene expression of MEX-receiving cells and tissues (11,33,50,77-81), although the life period and the magnitude of their gene-regulatory impact is a matter of debate (30,82,83). During the postnatal period, MEX function as epigenetic regulators (82,84). MEX miR-148a targets DNA methyltransferase 1 (DNMT1) and p53 (TP53), decreasing their expression (11,77,84-88).

In all mammals except humans, MEX miR exposure is restricted to the breastfeeding period and declines after weaning. About 7,000 to 10,000 years ago, Neolithic humans started to exploit milk from other mammalian species (89), but consumed preferentially fermented milk (89,90), which compared to raw milk contains diminished quantities of bioactive MEX miRs (91). Bovine MEX miRs are conserved by pasteurization of cow milk (72–78 °C, for >15 s), allowing their delivery into the human food chain (92,93).

To evaluate the risk-benefit relation of MEX miRs, five constellations of MEX miR exposure during human life have to be analyzed: (I) the impact of bovine MEX miR exposure during pregnancy and fetal development; (II) the physiological action of human MEX miRs during breastfeeding; (III) the absence of human MEX miRs by artificial formula feeding; (IV) the long-term influence of bovine MEX miRs on consumers of pasteurized cow milk; and (V) the action of intrinsic bovine MEX miRs and MEX miR transfer via drug-loaded MEX for the treatment of human diseases. It is important to be aware that the regulation of gene expression by miRs is highly complex, as a specific miR may regulate or repress the expression of hundreds of different mRNAs, which may act in concert and either stimulate (e.g., repression of an inhibitor) or inhibit (e.g., repression of an activator) a given cellular process.

Bovine MEX miR exposure during pregnancy

Consumption of unfermented milk during pregnancy is associated with fetal weight gain and higher birth weight (94-98), an indicator of fetal growth related to placental weight (99). In contrast to fermented milk with degraded MEX (93), raw and pasteurized milk deliver bioactive MEX including miR-21 and miR-148a (11,29,92,93,100,101). Following oral gavage of bovine MEX to mice, miR-21 and miR-30d accumulated in murine placenta and embryos (35). MIRNA30D knockout mice exhibit fetuses with smaller crown-rump length and fetal/placental weight ratio (102). Increased placental expression of miR-21 has been related with placental weight and fetal overgrowth (103,104). MEX miR-21 and miR-148a target phosphatase and tensin homolog (PTEN) and promote PI3K-AKT-mTORC1 signaling (77,105). Indeed, target gene prediction of human, bovine and porcine MEX miRs mainly concentrate on the PI3K-AKT-mTORC1 pathway (30,73,106). Increased trophoblast mTORC1 activity determines placental-fetal transfer of amino acids and glucose promoting fetal growth and birth weight (107-112). Thus, oral exposure of pregnant women to bovine MEX miRs may over-stimulate placental and fetal mTORC1 activation enhancing fetal growth.


Breastfeeding: the physiological MEX miR exposure of the infant

Intestinal maturation

Human, bovine and porcine MEX and their miRs improve intestinal maturation and proliferation (113-117), protect against epithelial damage (44,118-120) stabilize intestinal barrier function (46,121-123), exert anti-inflammatory (47-49,69,124,125) and anti-oxidative activities (126,127), stimulate intestinal and systemic immunity (6,7,64,128-130), antimicrobial defense (46,131), and appropriate conditioning of the gut microbiota (131,132). Notably, miR-148a inhibits TLR4/NF-κB and IL-6/STAT3 signaling (51,69,133-139) (Table 3). Experimental colitis models confirm that MEX attenuate TLR4 expression and NF-κB activation (45,118). Thus, MEX are of key importance for postnatal intestinal development and suitable agents for the prevention and treatment of necrotizing enterocolitis (46-49,69,86,140,141).

Table 3

Anti-inflammatory targets of miR-148a

Target genes Regulators of inflammatory responses References
TLR4 Toll-like receptor 4 (133)
CAMK2A Calcium/calmodulin-dependent protein kinase IIα (134,135)
CHUK Component of nuclear factor κ-B kinase complex (136)
IKBKB Inhibitor of nuclear factor κ-B kinase, subunit β (136-138)
IL6ST Interleukin 6 signal transducer (139)

miR, microRNA.

Immune tolerance

MEX that entered the systemic circulation may stimulate thymic maturation inducing regulatory T-cells (Tregs) (64). Admyre et al. (6) showed a dose-dependent expression of Forkhead box P3 (FoxP3) in CD4+ T-cells after addition of human MEX. FoxP3 is the master transcription factor of regulatory T (Treg) cells. Its expression is controlled by a Treg-specific demethylation region (TSDR) (142,143). DNMT1 and DNMT3B are associated with the FOXP3 locus in CD4+ T-cells. Whereas methylation of CpG residues represses FoxP3 expression, complete demethylation induces stable FoxP3 expression (144). Thus, MEX miR-148a-mediated suppression of DNMT1 and miR-29b-mediated suppression of DNMT3B may enhance Treg cell maturation that maintains immune tolerance, preventing allergy and autoimmunity (64,145,146).

Neuronal development

After oral gavage of bovine MEX to mice, MEX accumulated in the brain (32). In mice, MEX cross the blood-brain-barrier and promote dendritic complexity in the hippocampus, whereas dietary depletion of bovine MEX impaired sensorimotor gating and cognitive performance (147). In early postnatal life, developmental processes are critical for establishing proper neuronal connectivity. One protein functionally important for postnatal synaptic plasticity is α-synuclein (148-150). Hypomethylation of the SNCA promoter increases α-synuclein expression, which is regulated by DNMT1 (151). MEX miR-148a-mediated suppression of DNMT1 in neuronal cells may thus promote α-synuclein expression enhancing neuronal connectivity and cognitive function.

Adipose tissue development

Preterm infants receive high amounts of MEX miR-148a and miR-22 (115). Oxytocin, which is released during birth and lactation (152), induces the expression of MEX miR-148a and miR-30 in colostrum (153). Intriguingly, maternal obesity is negatively associated with the content of MEX miR-148a, miR-29a, miR-20b, miR-30b and miR-32 in human milk after one month of lactation (154). MEX miR-148a was negatively associated with infant weight, fat mass, and fat free mass, while miR-30b was positively associated with infant weight, percent body fat, and fat mass at 1 month (154). In mothers with gestational diabetes, levels of MEX miR-148a, miR-30b, let-7a and let-7d were also reduced (155). MEX miR-148a was negatively associated with infant weight, percentage of body fat, and fat mass, whereas MEX miR-30b was positively associated with infant weight and fat mass at 1 month of age. MEX miR-148a was negatively associated with infant weight at 6 months of age (155). Unfortunately, both studies did not differentiate between white adipose tissue (WAT), beige (BET) and brown adipose tissue (BAT), which is a critical limitation because MEX miRs may influence both WAT and BET/BAT development (69). Overexpression of miR-30b/c induces the expression of thermogenic genes such as uncoupling protein 1 (UCP1) in beige/brown adipocytes (156). MiR-30b/c target NRIP1, the gene encoding receptor-interacting protein 140 (RIP140) (156). RIP140 is essential for DNA and histone methylation to maintain gene repression and promotes the assembly of DNA and histone methyltransferases on the UCP1 enhancer (157). RIP140 directly interacts with DNMT1, DNMT3A, and DNMT3B and recruits H3K27me3 for the repression of UCP1. H3K27me3, which contains tri-methylation of lysine 27 on histone H3 protein, is associated with downregulation of nearby genes via the formation of heterochromatic regions. RIP140 represses the “brown-in-white” adipocyte program (158). MEX miR-30b/c-mediated suppression of RIP140 (156), MEX miR-148a-mediated suppression of DNMT1 (11,77), MEX miR-29b-mediated suppression of DNMT3A and DNMT3B (159,160) may thus increase BAT. Among the top 50 miRs abundantly expressed in bovine whey MEX is miR-489 (161), which also targets NRIP1 (162). Apparently, MEX miRs coordinate thermogenic adipose tissue development, which is important for the infant’s energy homeostasis.

Transcriptional activity and chromatin remodeling

The transcription factor p53, the guardian of the genome (163,164), interacts with approximately 1/10th of human gene promoters (165). p53 modifies the expression of target genes involved in cell cycle control (CDKN1A) (166), growth factor signaling (AR, IGF1R), translation and metabolism (PTEN, MDM2, mTORC1) (166-168), autophagy (ATG5, BECN1) (169,170) and apoptosis (FOXO1A, FOXO3A, TNFRSF10B, BIRC5) (166,171-173). Importantly, p53 is a direct target of human MEX miR-148a (86). Another highly conserved suppressor of p53 is miR-125b (174), which has been detected in human (7), bovine (24,175) and porcine MEX (9). Further negative regulators of p53 are miR-30d and miR-25 (176), which have been detected in human and porcine MEX (7,8,21,100,175). Thus, a network of MEX miRs synergistically attenuates p53 expression (45,86) (Figure 1). Furthermore, MEX miRs activate AKT (11,105,171) via inhibition of PTEN and may enhance p53 degradation via AKT-mediated phosphorylation of mouse double minute 2 (MDM2) (177,178). Mecocci et al. (79) identified MDM4 as a central node of transcriptomic regulation of cow, donkey and goat MEX RNAs. MDM4 restricts p53 transcriptional activity and facilitates MDM2’s E3 ligase activity toward p53 (179).

Figure 1 MEX-derived miRs attenuate the expression of p53, DNMT1, DNMT3A and DNMT3B. p53 suppression enhances cell cycle progression, growth factor signaling stimulating mTORC1-dependent cell proliferation, but reduces apoptosis and autophagy. MiR-148a-mediated suppression of p53 and DNMT1 and miR-30b-mediated suppression of RIP140 reduce HDAC activity opening chromatin structure enhancing transcription. MEX, milk exosome; miRs, microRNAs; DNMT1, DNA methyltransferase 1; HDAC, histone deacetylase.

MEX miR-148a/miR-21/miR-29b-mediated suppression of DNMT1, DNMT3A and DNMT3B (11,77,159,160) may reduce DNA methylation-dependent silencing of developmental genes (180) that are critically involved in postnatal programming (181). Notably, p53 and DNMT1 cooperate in gene silencing and interact on the promoter of BIRC5 with histone deacetylase 1 (HDAC1) (182) (Figure 1). MEX miR-mediated suppression of p53 and DNMT1 may disrupt the action of HDAC1 (182), thereby opening chromatin structure enhancing transcriptional activity (85,182). Functional analysis for the most abundant mRNAs of bovine MEX showed epigenetic functions such as histone modification, telomere maintenance, and chromatin remodeling (79). Liu et al. (106) identified methyl CpG binding protein 2 (MeCP2) as target gene of the most abundant miRs of porcine MEX. MeCP2 induces a repressive state of chromatin by recruiting histone deacetylases (HDACs), methyltransferases and other chromatin-silencing factors, which increase chromatin condensation and prevent transcription (183-186). DNMT1 directly associates with MeCP2 in order to perform maintenance of methylation in vivo (187). In addition, the methyl-CpG-binding domain of MeCP2 shows preferential interactions with H3K27me3 (188), which augments transcription repression (189,190). Of note, downregulation of H3K27me3 and upregulation of H3K27ac are involved in the activation of BAT thermogenic program (191,192).

The Polycomb repressive complex 2 (PRC2), composed of enhancer of zeste homolog 2 (EZH2), embryonic ectoderm development protein (EED), suppressor of zeste 12 homolog (SUZ12), and Jumonji and ARID-domain-containing protein 2 (JARID2) catalyzes the deposit of the repressive histone mark trimethyl lysine 27 of histone H3 (H3K27me3) at target gene promoters (193,194) (Figure 2). The methyltransferase EZH2 is the catalytic subunit of PRC2 (195). Importantly, the let-7 family, key miRs of human and bovine MEX (15,49), directly targets the 3'UTR of EZH2 and downregulate its expression (196,197). EZH2 directly regulates DNA methylation by serving as a recruitment platform for DNMT1, DNMT3A and DNMT3B through binding of DNMTs to several EZH2-repressed genes (198). In turn, EZH2 regulates the Lin28/let-7 pathway to restrict the activation of fetal gene signature in adult hematopoietic stem cells (199).

Figure 2 Illustration of predicted MEX miRs-mediated suppression of PRC2. The catalytically active enzyme of PRC2 is EZH2, which is a direct target of the let-7 family of miRs. The PRC2 activating component JARID2, which enhances HEK27me3 formation, is a target of miR-148a. Further suppressive effects of MEX miRs attenuate the interaction of MeCP2-DNMTs and EZH2-DNMTs, thus enhance transcription via reduced DNA and histone methylation. MEX miRs operate under control of the maternal lactation genome programmed to open chromatin structure and enhancing transcription, a reasonable support for cell growth and epigenetic programming of the newborn infant. Continued exposure of adult human cells to lactation signaling via MEX miRs is of critical concern with regard to the initiation and progression of cancer. MEX, milk exosome; miRs, microRNAs; PRC2, polycomb repressive complex 2; JARID2, JUMONJI, AT-rich interactive domain; DNMT, DNA methyltransferase; EZH2, enhancer of zeste homolog 2.

MiR-29a/b/c and miR-30b/c target EED (200,201). An inverse relationship between the miR-30b expression and the amount of trimethyl H3K27 has been reported (201). SUZ12 is a predicted target gene of miR-21 (202), another key miR of MEX. PRC2 forms a stable complex with JARID2, which binds to more than 90% of Polycomb group target genes (203). JARID2 methylated by PRC2, triggers allosteric activation of PRC2’s enzymatic activity (204). Inhibition of JARID2 expression leads to a major loss of Polycomb group binding and to a reduction of H3K27me3 levels on target genes (203). Remarkably, miR-148a, miR-155 and miR-29, pivotal miRs of MEX, are among the predicted miRs targeting JARID2 (205-207). EZH2-mediated H3K27me3 allows polycomb repressive complex 1 (PRC1) recruitment to chromatin, which establishes a higher repressive state of chromatin (193). Chromobox 2 (CBX2), a major subunit of canonical PRC1 recognizes H3K27me3 (208,209) and is a direct target of let-7 and miR-30 (210-212). The accumulation of let-7 in Lin28a–/– mice resulted in the reduction of PRC1 occupancy at the HOX cluster loci by targeting CBX2 (212). It is thus conceivable, that signature MEX miRs inhibit the activities of both PRC2 and PRC1 attenuating H3K27me3-mediated gene silencing (Table 4).

Table 4

Potential inhibitory impact of MEX miRs on formation of repressive H3K27me3

Target genes MEX miRNAs References
MECP2 porcine MEX miRs (106)
EZH2 let-7 family (196,197)
EED miR-29s, miR-30b, miR-30c (200,201)
SUZ12 miR-21 (202)
JARID2 miR-148a, miR-155, miR-29 (204,206,207)
CBX2 let-7, miR-30 (210-212)

MEX, milk exosome; miR, microRNA.

RIP140 is essential for both DNA and histone methylation to maintain gene repression (157,213). RIP140 mediates H3K9, H3K27 and DNA methylation for silencing of Polycomb group pre-marked genes (157). RIP140 directly recruits HDACs for gene silencing (214,215). MEX miR-30b/c mediated suppression of RIP140 may thus relieve RIP140-mediated gene silencing.

MEX miRs synergistically reduce the action of transcriptional repressors (p53, DNMTs, MeCP2, RIP140, EZH2, EED, JARID2, CBX2, H3K27me3) enhancing transcription and opening of chromatin structures, a meaningful epigenetic action of the lactation genome activating transcription for proliferation and infant growth.


Absence of MEX miRs during artificial formula feeding

Compared to raw cow milk, formula powdered milk for infants exhibits significant deficiencies in signature miRs, such as miR-148a, miR-30d and miR-21 (100). Leiferman et al. (216) reported that MEX miRs are not detectable in formulas. The levels of miR-148a and miR-125b levels in infant formula were only 1/500th and 1/100th of those in mature human milk, respectively (217). There is substantial concern that the absence of MEX miRs in artificial infant formula negatively affects appropriate postnatal epigenetic programming (84,85).

Pancreatic β-cell mTORC1/AMPK balance

Neonatal β-cells are immature and unable to secrete insulin appropriately in response to a glucose challenge (218). Adult β-cells repress a small set of housekeeping genes—such as those encoding lactate dehydrogenase A (LDHA), monocarboxylate transporter 1 (MCT1), and hexokinase 1 (HK1)—that would otherwise interfere with normal β-cell function. Dhawan et al. (219) elucidated a molecular mechanism involved in β-cell-specific repression of LDHA and HK1 that is mediated by induction of DNMT3A during the first weeks after birth. Failure to induce DNMT3A-dependent methylation disrupts normal glucose-induced insulin secretion (GSIS) in adult life. Recent evidence indicates that mTORC1 upregulates β-cell DNMT3A levels via translational control (220). MEX miRs may support appropriate mTORC1 signaling for postnatal β-cell maturation and mass expansion (221). Signature miRs of MEX enhance mTORC1 signaling via targeting PTEN [miR-148a, miR-155, miR-21 (77,222,223)], phosphatidylinositol 3-kinase-interacting protein 1 (PIK3IP1) [miR-148a (224)], PRKAA1 [miR-148a (225)], the catalytic subunit α1 of AMP-activated protein kinase (AMPK) and PRKAG2 [miR-148a (226)], the regulatory subunit γ2 of AMPK. Jaafar et al. (227) reported that the control of cellular signaling in β-cells fundamentally changed after weaning and switched from the nutrient sensor mTORC1 to the energy sensor AMPK, which was critical for functional β-cell maturation and GSIS.

Another signature miRNA of human milk is miR-375 (11), which was significantly more abundant in the colon of mice treated with human MEX (49). Bovine miR-375 also belongs to the most highly expressed miRs of bovine whey exosomes (161). Intriguingly, miR-375 is one of the most abundant miRs expressed in pancreatic β-cells and its overexpression suppresses GSIS, whereas its inhibition enhances insulin secretion (228). Importantly, miR-375 regulates the expression of genes involved in the control of β-cell mass and identity (229). Mice with genetic deletion of miR-375 exhibit decreased β-cell mass and function (230,231).

Exosomal miR-155 derived from adipose tissue macrophages of obese mice also exerts profound regulation on β-cells, leading to impaired GSIS and increased β-cell proliferation by repressing MAF bZIP transcription factor B (MAFB) (232). MAFB is a direct target of miR-155 (233), miR-148a (234) and miR-29 (235). It is thus conceivable that MEX miR-148a/miR-155/miR-29 via targeting MAFB enhance β-cell proliferation but impair GISIS, a meaningful mechanism during the breastfeeding period characterized by low variations in external glucose challenge and predominant dependence on amino acids for insulin secretion.

Thus, during breastfeeding, MEX miRs apparently maintain β-cells in a functionally immature state with restricted GSIS to use the developmental period of lactation for adequate mTORC1- and MEX miR-dependent β-cell growth and mass extension. The disappearance of MEX during weaning may be the critical signal for switching β-cells to functional maturation and GSIS. Absence of MEX miRs in the setting of artificial formula feeding may thus compromise adequate β-cell growth and mass extension increasing the risk for type 2 diabetes mellitus (T2DM) later in life.

Fat mass and obesity gene-dependent pathologies

Another critical gene involved in the development of adipogenesis and T2DM is the fat mass and obesity-associated gene (FTO), which encodes a mRNA N6-methyladenosine (m6A) demethylase (236-238). FTO targets key regulatory checkpoints enhancing adipogenesis (239-248) (Table 5). It is of critical concern that the levels of FTO expression in peripheral blood mononuclear cells (PBMCs) of formula fed infants is significantly higher (89.2±19.3) compared to breastfed infants (3.39±1.1) (249). Tews et al. (250) showed that FTO-deficient adipocytes exhibit a 4-fold higher mitochondrial expression of UCP1 compared with controls.

Table 5

FTO-mediated effects promoting adipogenesis

FTO targets Adipogenic functions References
SREBP1c FTO erases m6A marks on pre-mRNA of SREBP1c enhancing its expression (239,240)
Pre-adipocytes Differentiation of preadipocytes to adipocytes (241)
C/EBPβ Upregulation of C/EBPβ reducing UCP1 expression and formation of brown adipose tissue (242)
RUNX1T1 Generation of alternative splice variant of RUNX1T1 stimulating adipogenesis (243)
MIR130A Downregulation miR-130a enhancing PPARγ expression augmenting adipogenesis (244)
ANGPTL4 Reduction of angiopoietin-like protein 4 protein accelerating lipoprotein lipase release and extracellular triacylglycerol hydrolysis promoting triacylglycerol synthesis and formation of lipid droplets in adipocytes (245,246)
IRX3 Downregulation of hypothalamic IRX3 and inhibition of lipolysis in peripheral adipocytes (247)
CUX1 Interaction with CUX1 increasing retinitis pigmentosa GTPase regulator-interacting protein-1-like expression, which reduces leptin signal transduction and leptin-induced lipolysis in adipocytes (248)

FTO, fat mass- and obesity-associated gene.

MEX miRs appear to be critically involved in balancing appropriate levels of FTO expression. In fact, miR-21, miR-22, miR-30b, miR-150 and miR-155, which represent miR components of milk and MEX (7,9,15,18,100,115,251), directly target FTO mRNA (252-255). Intriguingly, miR-30b targets both RIP140 (156) and FTO (255), thus synergizes with different regulatory checkpoints known to enhance the expression of UCP1 and development of BAT. Suppression of miR-30b upregulates FTO expression resulting in aberrant FTO-mediated m6A methylation levels promoting lipid accumulation (255). Thus, deficient MEX miR signaling during formula feeding may explain the significant exacerbation of FTO expression in blood mononuclear cells of formula-fed infants compared to breastfed infants (249,256).

Increased expression of FTO is not only involved in the pathogenesis of obesity (236-241) but also of T2DM (257-259). Notably, m6A mRNA methylation plays a key role in human β-cell biology in physiological states and in T2DM (260). Recent transcriptome and m6A methylome analyses provided evidence for m6A-dependent mechanisms in regulating cell identity, insulin secretion, and proliferation in neonatal β-cells (261). Wang et al. (261) found an essential role of RNA methyltransferase-like 3/14 in neonatal murine β-cell development and functional maturation, both of which determined functional β-cell mass and glycemic control in adulthood. In addition, increased FTO expression is also associated with hypertension (262), cancer (239,263-265) and osteoporosis (266).

Insufficient MEX miR-mediated inhibition of DNMT1 and FTO via MEX-deficient formula feeding may negatively affect the maturation of Treg cells, as stable FoxP3 expression and T-cell homeostasis is epigenetically and epitranscriptionally controlled at the appropriate methylation level of DNA and mRNA (145,146,267-269). These epigenetic modes of action further support the association between formula feeding and increased risk for obesity and allergy, which is also related to increased mTORC1 activity (270). Recent evidence shows that Mettl3f/f; Foxp3Cre Treg cells lost their suppressive function over T cell proliferation (271). Obviously, m6A mRNA methylation sustains Treg cell suppressive functions (271). Formula-induced FTO overexpression may thus impair m6A-dependent Treg cell suppressive functions enhancing the risk of allergy and autoimmunity. Wood et al. (272) recently showed that breastfeeding compared to formula resulted in a twofold higher early neonatal Treg cell expansion.


Health risks associated with persistent bovine MEX miR exposure

Consumers of pasteurized cow milk are persistently exposed to bovine MEX miRs. During the last century, pasteurization of milk has been introduced to improve the microbial safety and quality of milk (273,274). At that time, the presence of MEX and their miRs in pasteurized milk were unknown. Recently, the bioavailability of bovine MEX miRs in pasteurized commercial milk has been confirmed (92,93,275,276). Potential adverse health effects of milk’s MEX and their intrinsic miRs are a matter of medical concern (277), because bovine MEX miRs can enter the human body and affect human gene expression (11,30,78-81,277).

Cancer risk

Bovine MEX transfer oncogenic miRs such as miR-21 (29,100), which is identical with human miR-21 (278). Common cancers of industrialized societies such as breast cancer (BCa) and prostate cancer (PCa), which show a risk association with milk consumption (279-282), exhibit exosomal overexpression of miR-21 in the circulation and tumor tissues (283-290). Thus, persistent bovine MEX miR-21 exposure may enhance mTORC1 signaling promoting tumor initiation and progression (291-296). Furthermore, MEX miR-mediated destabilization of p53 (86,88,174,176) may promote cancer development and disturb tumor immunity (297-299).

Intriguingly, bovine MEX orally administered to mice implanted with colorectal and BCa cells reduced the primary tumor burden but accelerated metastasis in BCa and pancreatic cancer mouse models (78). Upon treatment with MEX, epithelial-to-mesenchymal transition (EMT) has been observed in cancer cells (78). Increased expression of miR-148a has been detected in PCa tissue correlating with increased Gleason score (300). The androgen-responsive miR-148a promoted LNCaP prostate cell growth by repressing cullin-associated neddylation-dissociated protein 1 (CAND1) (301) and DNMT1 (11,77). Reduced expression of DNMT1 was associated with EMT induction and cancer stem cell phenotype, enhancing tumorigenesis and metastasis of PCa (302). PCa-derived exosomes, which contain oncogenic miR-21, miR-155 and miR-125b promote PCa progression (303). Notably, exposure of human breast cells with human MEX enhanced EMT-associated proteins related to transfer of MEX-derived TGF-β2 (304), a known inducer of EMT-promoting miR-155 (305-308). Bovine MEX transport TGF-β (288) and miR-155 and increase intracellular levels of miR-155 (126). Exosomal miR-125b was upregulated in highly invasive pancreatic cancer cells with increased migration, invasion, and EMT (309). Melanoma-secreted exosomal miR-155 can induce a proangiogenic switch of cancer-associated fibroblasts (310). Exosomal miR-21 has been shown to promote melanoma development and progression (311). Mastitis, a common chronic health problem of dairy cows, has been associated with increased expression of MEX miR-223 (312), another oncogenic miR associated with human cancers (313). Thus, persistent MEX miR exposure via consumption of pasteurized cow milk may synergize with cancer-derived exosomes enhancing the total burden of oncogenic exosome-mediated miR signaling.

Adipogenesis and satiety control

In adipose tissues from obese individuals and mice fed a high-fat diet, the expression of miR-148a is increased and silences the endogenous inhibitors of adipogenesis WNT1 and WNT10B (314-316). Increased expression of miR-148a via suppression of DNMT1 enhances adipocyte differentiation and precociously promoted adipocyte-specific gene expression and lipid accumulation (317). Inhibition of DNA methylation at late stage of pre-adipocyte differentiation promotes lipogenesis and adipocyte phenotype in 3T3-L1 cells. This is mediated by promoter demethylation of SREBF1c, which is upregulated during adipogenesis (318). MIR148A represents a domestication gene of dairy cows increasing milk yield (319,320) but is also an obesity risk gene in humans (321-323). Persistent transfer of bovine MEX miR-148a may thus enhance adipogenesis and SREBF1c-mediated lipid accumulation in adipocytes.

In addition, miR-148a directly inhibits the mRNA of the low density-lipoprotein (LDL) receptor (LDLR) (324,325), the pivotal regulator of cholesterol homeostasis and hepatic LDL clearance (326). Importantly, miR-148a also inhibits ATP-binding cassette transporter 1 (ABCA1) (324), the key player for HDL-mediated reverse cholesterol transport (326-328), thus promoting hypercholesterolemia and atherosclerosis.

Adipocyte-derived exosomes can regulate proopiomelanocortin (POMC) expression through hypothalamic mTORC1 signaling, thereby affecting body energy intake. Adipocytes of obese mice secreted MALAT1-containing exosomes, which increased appetite and weight when administered to lean mice (329). It is thus conceivable that MEX, which reach the brain (32,147), may also have an impact on hypothalamic appetite and satiety control. One of the gut hormones sending satiety signals to the hypothalamus is cholecystokinin (CCK), which is secreted from intestinal mucosa cells when the duodenum is filled with food (330). CCK binds to the CCK A receptor (CCKAR) and the CCK B receptor (CCKBR). CCKBR knock out mice developed obesity associated with hyperphagia (331). As shown in rodents, hypothalamic CCKBR mediates inhibition of food intake (331,332). CCKBR deletion was associated with increased body weight and hypothalamic neuropeptide Y content, explaining the increased food intake in CCKBR knockout mice (333). Notably, the gene expressing CCKBR is a direct target gene of miR-148a (334) exhibiting four different target sites for miR-148a on the mRNA of CCKBR (335). Further predicted miRs that target CCKBR are miR-29b, miR-30a, miR-30d, miR-223 (335), which are all miR components of MEX. Hypothalamic MEX miR signaling may thus attenuate satiety signals. This mode of action may be of advantage for postnatal growth and weight gain of the infant, but could promote obesity in adults persistently exposed to bovine orexigenic MEX miRs.

MiR-21 is another important miR which contributes to adipocyte differentiation (336-339) and promotes human adipose tissue-derived mesenchymal stem cell (MSC) differentiation towards adipocytes (336). Overexpression of miR-21 in MSCs enhances the expression of PPARγ and modulates ERK-MAPK activity by repressing Sprouty 2 (337), a known negative regulator of receptor tyrosine kinase pathways involved in ERK-MAPK signaling during MSC differentiation (337). Notably, miR-21 expression was twofold higher in WAT of patients with T2DM (339).

MiR-29b is another abundant miR of cow milk and MEX (27,275), which is detectable after pasteurization and homogenization of milk (276) and exhibits increased expression in plasma and PBMCs after milk consumption (275). MiR-29b promotes adipogenic differentiation of human adipose tissue-derived stromal cells (340). In bovine mammary epithelial cells, miR-29s enhance lactation performance and lipogenesis via suppression of DNMT3A and DNMT3B, whereas inhibition of miR-29s increased the methylation levels of promoters of lactation-related lipogenic genes, reducing the expression of PPARG and SREBF1 (341).

Overexpression of miR-155 in mice has been shown to reduce BAT mass (342), whereas thermogenic miR-30b promotes UCP1 expression and BAT formation (156).

Obviously, MEX miRs may synergize with key miRs involved in the regulation of adipogenesis (343).

T2DM

Only few epidemiological studies compared the risk of milk consumption versus fermented milk/products in relation to the risk of T2DM (344). The Lifeline Cohort Study identified positive associations of full-fat dairy products, non-fermented dairy products and milk with newly diagnosed T2DM (345). Unfortunately, no study related pasteurized versus ultraheat-treated (UHT) milk with a T2DM risk analysis. This is of importance because consumers of pasteurized milk are exposed to bioactive MEX, while MEX miRs are reduced by UHT (92,93). Kleinjan et al. (92) showed that UHT of milk resulted in a loss of EVs, whereas EV numbers after pasteurization were not affected. Although pasteurization and homogenization of commercial milk reduced total milk-EV-associated RNAs [from 40.2±3.4 ng/µL in raw milk to (17.7±5.4)–(23.3±10.0) ng/µL in processed milk], miR-148a, miR-21, miR-30d, although reduced, were still detectable in pasteurized and homogenized milk (92). Zhang et al. (93) recently confirmed that bioactive miRs in raw milk were lost after ultraheat treatment but not after pasteurization.

Observed exosome crosstalk between pancreatic β-cells and exosomes derived from different cell types (lymphocytes, adipocytes, and muscle cells) supports the concept that exosomal miRs communicate with β-cells (346-350). For instance, human T lymphocyte-derived exosomes, which transfer miR-155, induced apoptosis in β-cells and promoted type 1 diabetes mellitus (T1DM) in mice (348). Exosomes from insulin-resistant muscles influenced gene expression and proliferation in murine β-cells (349). Exosomes from obese adipose tissue were harmful for human β-cells (350). Exosomes of inflammatory adipocytes exhibited a fourfold increased expression of miR-155 (351).

Postnatal β-cell maturation is associated with islet-specific miR changes induced by nutrient shifts at weaning (352). It is thus conceivable that MEX miRs may participate in exosomal β-cell interactions supporting postnatal β-cell proliferation and mass expansion but suppress GSIS, which is activated after weaning (221). Pancreatic β-cells differentiate during fetal life, but only postnatally acquire the capacity for GSIS. Jaafar et al. (227) found that the control of cellular signaling in β-cells fundamentally switched from the nutrient sensor mTORC1 to the energy sensor AMPK, which was critical for functional maturation. Notably, T2DM is associated with a remarkable reversion of the normal AMPK-dependent adult β-cell signature to a more neonatal one, characterized by mTORC1 activation (227). Allowing mice to continue assimilating milk fat throughout their entry into adulthood was sufficient to maintain neonatal levels of β-cell mTORC1 activity, which was otherwise completely repressed in control mice (227). Recent evidence confirms that the shift from amino acid- to glucose-stimulated insulin secretion after birth is mediated by a transition in nutrient sensitivity of the mTORC1 pathway, which leads to intermittent mTORC1 activity (353).

The switch from mTORC1 to AMPK signaling is important to shift β-cell mitochondrial biogenesis to oxidative metabolism and functional maturation (227). AMPK is regarded as guardian of metabolism and mitochondrial homeostasis (354). AMPK upregulation inhibits β-cell apoptosis (355). AMPK is activated by the action of metformin (356), which inhibits mTORC1 (357) and suppresses miR-21 (358). The suppression of miR-21 enhances the expression of critical upstream activators of the AMPK including calcium-binding protein 39-like protein and sestrin-1 (358-361) leading to AMPK activation and inhibition of mTORC1 (358). Thus, metformin-induced mTORC1 inhibition and AMPK activation simulate the switch of mTORC1 to AMPK activation, a comparable mechanism found after weaning associated with the physiological termination of MEX miR signaling.

Importantly, human and bovine milk fat as well as human and bovine MEX are a rich source of miR-148a (11,101,362,363), which targets key regulatory components of AMPK (225,226,364), thereby attenuating AMPK’s inhibitory function on mTORC1 via AMPK-mediated phosphorylation of TSC2 and Raptor (365,366). Milk fat and MEX miR-148a may stimulate β-cell mTORC1 activity promoting β-cell growth and mass expansion (367). The disappearance of MEX miR-148a after weaning may induce the critical switch to enhanced AMPK activity promoting β-cell mitochondrial function and GSIS to respond to dietary challenges of varying glucose intake. Persistent exposure of humans to bovine MEX miR-148a may dedifferentiate β-cells back to postnatal conditions characterized by overactivated mTORC1, which promotes β-cell proliferation and in the long run induces early β-cell apoptosis (221,341).

This concept is further supported by the regulatory events of MAFB, which is required for the generation of functional β-cell populations by directly activating insulin gene transcription and key regulators of β-cell differentiation and function (368). Importantly, MAFB increases the expression of MAFA, which is important to maintain β-cell function in adults (369). MAFB is a direct target of miR-148a (234). Reduced expression of MAFB in murine and human β-cells has been associated with decreased GSIS (370). Loss of MAFB is associated with β-cell dedifferentiation. Loss of MAFA and/or MAFB represents an early indicator of β-cell inactivity and the subsequent deficit of more impactful NKX6.1 (and/or PDX1) resulting in overt dysfunction associated with T2DM. Notably, MAFA, MAFB, NKX6.1, and PDX1 expression levels are compromised in human β-cell in T2DM (371). The significance of MAFB to primate β-cells is supported by suppressed GSIS in the human EndoC-βH1 β-cell line by MAFB knockdown (372). Both, MAFA and MAFB mediate GSIS in human β-cells (373). The primary glucose transporter of human β-cells is GLUT1 (SLC2A1) (374,375). GLUT1 expression is activated by MAFB (372), which is required for formation of glucose-responsive β-cells (376). Remarkably, GLUT1 (SLC2A1) and MAFB are targets of miR-148a (234,377).

It has recently been demonstrated that overexpression of miR-21 in β-cells markedly reduced GSIS and led to reductions in mRNA expression of MAFA, NKX6.1, INS1, INS2 and GLUT2 associated with a loss of β-cell identity and increased markers of β-cell dedifferentiation (378). Increased expression of miR-21 reduced the expression of TGFB2 and SMAD2, direct targets involved in β-cell commitment (379). TGF-β signaling plays a key role for adult β-cell function and maturity (380). Induction of miR-21 in human islets was also associated with a dedifferentiated phenotype and reduced expression of miR-21 target mRNAs linked to β-cell identity (378). Furthermore, lentiviral overexpression of miR-21 in the β-cell line INS-1, increased proliferation, but also induced apoptosis, questioning the potential of miR-21 as a therapeutic agent to increase β-cell survival (381).

Metformin, which reduces the expression of miR-21, may thus counteract β-cell dedifferentiation (358). Bai et al. (382) reported that miR-21 acts as a bidirectional switch in the formation of insulin-producing cells by regulating the expression of target and downstream genes (SOX6, RPBJ and HES1).

In analogy to miR-21, miR-148a also targets TGFB2 and SMAD2 (383-386). Increased serum levels of miR-148a and miR-21 have been reported in patients with T1DM (387). Of note, SMAD2 deficiency impaired insulin secretion in response to glucose (388,389). SMAD2 disruption in mouse pancreatic β-cells leads to islet hyperplasia and impaired insulin secretion due to an attenuation of KATP channel activity (389), which plays a critical role in glucose homeostasis by linking glucose metabolism to electrical excitability and insulin secretion (390). TGFBR-SMAD2/3 signaling sustains functional maturation of neonatal β-cells (391). TGFBR–SMAD2/3 inhibition counteracted upregulation of CDKN2A, NEUROD1, UCN3 and ABCC8, the ATP-sensitive component of the KATP channel (391). Of note, TGFBR–SMAD2/3 signaling repressed aldolase B (ALDOB), a disallowed gene in mature β-cells and a marker of functionally immature as well as of diabetic β-cells (391). Inhibition of TGF-β signaling promotes human β-cell replication, whereas TGF-β signaling induces CDKN2A (INK4A) expression leading to replicative decline in β-cells through the recruitment of SMAD3 as a part of the recruitment of histone methyltransferase Mll1 complex (392). Furthermore, TGF-β has been reported to stimulate insulin secretion (393), insulin gene transcription, and islet function (394). TGF-β/SMAD pathway enhances the transcription of miR-375, miR-26a, and NGN3, thereby promoting β-cell differentiation (395). Notably, a fasting-mimicking diet in mice reduced PKA and mTORC1 activity and induced SOX2 and NGN3 expression and insulin production (396). MiR-375 and miR-26a induced insulin-producing cell differentiation from nestin-positive umbilical cord-derived MSCs by suppressing target genes including MTPN, SOX6, BHLHE22 and CCND1 (397).

Let-7 family members, major components of MEX (15,49), also suppress TGF-β signaling and decrease mRNA expression of TGFBR1, TGFBR3 and SMAD2 (398-400). Notably, the lncRNA H19, a sponge of let-7 (401), is profoundly downregulated during the postnatal period (402). It has been shown that let-7 promotes the expression of IRS2 and mTOR in β-cells (403), whereas suppression of let-7 expression in endothelical cells increased the expression of TGF-β and TGF-βR1 (404). Glucose-responsive genes are highly regulated by TGF-β signaling (405).

It is conceivable that MEX miR-148a/miR-21/let-7 signaling during the breastfeeding period via suppressing TGF-β signaling may promote β-cell proliferation and delays β-cell differentiation and GSIS, a physiological mechanism that fades after weaning.

Mice with specific inactivation of ABCA1 in β-cells exhibited impaired insulin secretion (406,407). Lack of β-cell ABCA1 leads to impaired exocytosis of insulin granules (408). ABCA1 as well is a direct target of miR-148a (324).

Another important target of MEX miR-148a is TP53 (86). Suppression of p53 activates mTORC1 (167) and inhibits CDKN1A, accelerating cell cycle progression (409). Loss of p53 function decreases lysosomal TSC2 and increases lysosomal Rheb, resulting in hyperactive mTORC1 (410). Survivin (BIRC5), which is inhibited by p53 (173), is critically involved in the regulation of β-cell mass after birth (411). Targeted deletion of survivin in the pancreas resulted in a significant decline in β-cell mass throughout the perinatal period, leading to glucose intolerance in the adult. Survivin-deficient islets showed decreased cell proliferation as a result of a delay in cell cycle progression with perturbations in cell cycle proteins (412). MEX miR-mediated suppression of p53 may thus promote survivin-induced β-cell mass expansion (Figure 3).

Figure 3 Illustrated β-cell model showing a β-cell exposed to MEX miR-148a signaling (A) during postnatal breastfeeding and (B) absent MEX miR-148a signaling after weaning. (A) MEX miR-148a suppresses TGFB2, SMAD2, p53, MAFB, GLUT1, PTEN, AMPK, PGC-1α and ABCA1 increasing mTORC1- and surviving-dependent β-cell proliferation but suppressing β-cell differentiation and GSIS. (B) After weaning, disappearance of MEX miR-148a enhances TGF-β signaling promoting β-cell differentiation with increased expression of MAFB, GLUT1, AMPK, ABCA1, ABCC8 activating GSIS but attenuating mTORC1-driven β-cell proliferation. Lack of MEX miR-148a in artificial formula compromises miR-148a-mediated β-cell proliferation and mass expansion. However, persistent exposure of humans with MEX-miR-148a/miR-21/let-7 of pasteurized cow milk may dedifferentiate the β-cell back to the postnatal hyperproliferative state with overactivated mTORC1, impaired AMPK and TGF-β signaling promoting ER stress and early β-cell apoptosis, key mechanism in the pathogenesis of type 2 diabetes mellitus. MEX, milk exosome; miR, microRNA; GSIS, glucose-stimulated insulin secretion; TGF, transforming growth factor.

Peroxisome proliferator-activated receptor-γ coactivator (PGC)-1α, a transcription coactivator that plays a central role in the regulation of cellular energy metabolism, stimulates mitochondrial biogenesis (413) and controls β-cell homeostasis (414). Of note, the susceptibility of β-cells to environmental programming continues into the neonatal period (415). Overexpression of PGC-1α in β-cells during fetal life in mice is sufficient to induce β-cell dysfunction in adults, leading to glucose intolerance (416,417). Ling et al. (418) demonstrated that downregulation of PPARGC1A expression in human islets by siRNA, reduced insulin secretion. Remarkably, key signature miRs of milk and MEX including the let-7 family and miR-148a target PPARGC1A (364,419,420) and may thus restrain insulin secretion during the breastfeeding period. Increased plasma levels of miR-148a have been associated with T2DM progression, increased HbA1c, HOMA-IR, and hyperinsulinemia (421) (Table 6). Melkman-Zehavi et al. (422) reported that miR-148 acts as a positive regulator of insulin transcription.

Table 6

Potential regulatory effects of MEX miR-148a on postnatal β-cell maturation

miRNA-148a targets Predicted regulatory effects of MEX miRNA-148a References
TP53 Increased expression of survivin (BIRC5) and suppression of p21 (CDKN1A) stimulates β-cell proliferation; decreased expression of TSC2 activates mTORC1; reduced sestrin (SESN1, SESN2) expression attenuates AMPK activity enhancing mTORC1 activation (86,166,167,173,360,382)
PTEN Reduced expression of PTEN activates PI3K and thus AKT-mTORC1 (77)
PIK3IP1 Reduced expression of PIK3IP1 activates PI3K and thus AKT-mTORC1 (224)
PRKAA1 Reduced expression of PRKAA1 reduces the catalytic activity of AMPK promoting mTORC1 activation and suppressing mitochondrial activity (225)
PRKAG2 Reduced expression of PRKAG2 reduces the activity of AMPK promoting mTORC1 activation and suppressing mitochondrial activity (226)
PPARGC1A Reduction of mitochondrial biogenesis and GSIS (364,392)
MAFB Reduced expression of the GLUT1 attenuates glucose sensing of β-cells; reduced expression of MAFA impairs the MAFA/MAFB-dependent maintenance function of β-cells (234,235,376)
SLC2A1 Reduced expression of GLUT1 attenuates glucose sensing of β-cells (377)
ABCA1 Impaired insulin granule function and insulin secretion (324,378-380)
TGFB2 Increased β-cell proliferation with decreased GSIS, delayed differentiation (381-384)
SMAD2 Increased β-cell proliferation with decreased GSIS, delayed differentiation (381-384)

MEX, milk exosome; miR, microRNA.

In accordance with miR-148a, miR-375 is overexpressed in T2DM (423-425). MiR-375 reduces GISIS and enhances β-cell proliferation (228-231,426,427) and plays a critical role for the differentiation of human embryonic stem cells into insulin-producing cells (428). Upregulation of miR-375 is associated with T2DM and apoptosis of pancreatic islet β-cells (429,430). Dietary exposure of adult milk consumers with MEX miR-375 may thus stimulate β-cell proliferation promoting early β-cell apoptosis.

MEX deliver miR-29s (27,275), which are considered as diabetogenic miRs promoting insulin resistance (431-435). Elevated plasma levels of branched-chain amino acids (BCAAs), crucial activators of mTORC1 (105), correlate with an increased risk of insulin resistance and T2DM (433-440). Intriguingly, miR-29b is critically involved in the regulation of cellular BCAA homeostasis (441). MiR-29b inhibits branched chain α-ketoacid dehydrogenase (BCKD) via targeting its core protein dihydrolipoamide branched-chain acyltransferase (441) thereby decreasing BCAA catabolism. This is a meaningful metabolic regulation for the newborn mammal protecting BCAAs from catabolism and saving them for the synthesis of functional and structural proteins as well as for BCAA-mediated mTORC1 activation (442,443), which is important for postnatal β-cell proliferation (221,341).

However, persistent overactivation of mTORC1 induces endoplasmic reticulum (ER) stress, which in the long run impairs β-cell autophagy, promotes β-cell apoptosis and β-cell failure (444,445). Notably, mTORC1 activity is markedly increased in islets from patients with T2DM (227,446).

Recent evidence indicates that β-cells control glucose homeostasis via the secretion of exosomal miR-29 family members (447). Notably, intravenous administration of exosomal miR-29a/b/c attenuated insulin sensitivity (447). In fact, β-cell-specific transgenic miR-29a/b/c mice are predisposed to develop glucose intolerance and insulin resistance when fed a high-fat diet (448). Expression of miR-29 in pancreatic β-cells promotes inflammation and diabetes via targeting tumor necrosis factor (TNF) receptor-associated factor 3 (TRAF3) to promote CCXL10 release from β-cells, which then attracts nearby circulating monocytes (448). Furthermore, miR-29 targets MAFB (235).

Taken together, key miRs donated via MEX promote β-cell proliferation and reduce GSIS, a meaningful metabolic regulation during the breastfeeding period to support β-cell mass expansion for insulin demands in adulthood. In contrast to the importance of MEX during the lactation period, in adult life MEX miRs may exert diabetogenic effects promoting β-cell dedifferentiation and loss of function (Figure 3).

Parkinson’s disease (PD)

Epidemiological studies associate consumption of unfermented milk with an increased risk of PD (449-456). PD is an α-synucleinopathy associated with deficient lysosomal clearance and aggregation of misfolded α-synuclein (456-458). Accumulating evidence supports a gut-brain axis in PD pathogenesis involving α-synuclein synthesis in intestinal enteroendocrine cells (EECs) and retrograde traffic of α-synuclein via the vagal nerve to the brain (459-462). Hypomethylation of the SNCA promoter increases α-synuclein expression, which is controlled by DNMT1 (151,351). In the chronic MPTP mouse model of PD increased expression of miR-148a reduced the expression of DNMT1, diminished α-synuclein methylation resulting in increased α-synuclein expression (463).

Uptake of MEX miR-148a by intestinal EECs may thus enhance intestinal α-synuclein synthesis increasing its retrograde transneuronal traffic to the brain and pancreatic islets (351). Although the major aggregating peptide in β-cells of T2DM patients is islet amyloid polypeptide (IAPP), α-synuclein in β-cells interacts with IAPP (464). Mucibabic et al. (465) showed that α-synuclein is a component of amyloid extracted from the pancreas of transgenic mice overexpressing human IAPP (denoted hIAPPtg mice) and from islets of T2DM individuals. Notably, α-synuclein promotes IAPP fibril formation in vitro and enhanced β-cell amyloid formation in vivo, whereas β-cell amyloid formation was reduced in mice on a SNCA−/− background (465). As bovine MEX are able to cross the blood-brain barrier and reach the brain (32,147), MEX miR-148a may stimulate neuronal α-synuclein in the substantia nigra, whereas MEX miR-21 may suppress lysosome-associated membrane protein type 2A (LAMP2A) (466,467), a critical effector protein of chaperone-mediated autophagy clearing neurotoxic α-synuclein (349).


Limitations

This narrative review focuses on MEX miRs and their potential beneficial and adverse effects on human health. However, it is important to consider that milk contains a large spectrum of EV subtypes whose components may exhibit a differential resistance to digestion and may have differential impacts on health (1,14,24). Furthermore, MEX contain a wide spectrum of very small, small, medium and long non-coding RNAs in addition to miRs (19,468), that may also contribute to biological effects. It has recently been discussed that exosomes provide unappreciated carrier effects that assist transfers of their miRs to targeted cells (130). In addition, MEX deliver other biomolecules such as lipids and proteins that may also determine biological outcomes of MEX-derived signaling (55), which are beyond the scope of this review. As long as risk assessments on a statistical basis are not yet available, statements of the potential hazardous nature of cited miRs based on translational and indirect evidence have to be judged with caution.


Conclusions

To approximate risks and benefits of MEX and their miRs, it is mandatory to appreciate their physiological inherent nature. MEX are signalosomes generated under control of the lactation genome to support growth and tissue maturation during the breastfeeding period. MEX miRs apparently interfere with p53-mediated transcriptional activity, DNMT-regulated gene silencing, histone methylation and chromatin remodeling. MEX miR-dependent effects are not provided by formula feeding potentially resulting in overexpression of FTO. The time period and dose of MEX miR exposure appears to be of critical importance for MEX miR-induced desired and adverse gene-regulatory effects. The persistent transfer of bioactive MEX miRs via consumption of pasteurized cow milk beyond the breastfeeding period is a recent insight (92,93), which has been neglected in all epidemiological studies correlating cow milk intake with human pathologies (90,277). The use of concentrated MEX either native or uploaded with drugs will expose the recipient to increased amounts of intrinsic MEX miRs that may exert desired gene-regulatory effects in the target tissue but may adversely affect other non-target tissues. Bovine MEX showed beneficial effects in inflammatory bowel diseases and states of fibrosis (47-51,469), but may promote hyperphagia, obesity, T2DM, cancer and neurodegeneration (88,90,221,277,296,351) (Table 7). Certainly, human MEX are supportive for the newborn infant and contribute to the superiority of breastfeeding compared to formula. However, long-term bovine MEX miR exposure via pasteurized cow milk may enhance the risk for diseases of civilization and should be terminated. MEX miR-mediated suppression of TGF-β signaling may be beneficial for the treatment of fibrosis (50-54), but may be deleterious for adult pancreatic β-cell homeostasis (386-393). Thus, there is no clear-cut answer to judge the benefit/risk evaluation of MEX miRs in general. Before systemic administration of bovine MEX miRs—either native or combined with uploaded drugs or anti-sense RNAs—is employed, long-term safety studies have to be performed, including tissues, which are not direct targets for therapeutic intervention. Future experimental investigations are required that determine bovine MEX miR delivery to cells and tissues of human cow milk consumers and their effects on gene expression not only in short term but especially in long-term studies to evaluate their physiological and pathological effects.

Table 7

Potential effects of bovine MEX miRs and related adult human pathologies

Target tissue Potential MEX-mediated effects Potential mode of action
Intestine Anti-inflammatory action of miR-148a in inflammatory bowel diseases (Crohn disease, Colitis ulcerosa); reduction of inflammation-induced colon cancer;
Stimulation of intestinal α-synuclein synthesis promoting Parkinson disease
miR-148a-mediated suppression of TLR4 and NF-κB signaling;
miR-148a/DNMT1-mediated stimulation of α-synuclein synthesis
Liver Attenuation of liver fibrosis and fibrosis-related liver cancer; stimulation of liver cancer Transfer of anti-fibrotic miR-148a and miR-29b;
Transfer of oncogenic miR-155 and miR-21
Pancreatic islet Increased β-cell proliferation resulting in early onset of β-cell apoptosis promoting type 2 diabetes mellitus; Impaired TGFB2/SMAD2 signaling promoting β-cell dedifferentiation miR-148a/miR-21/miR-375-mediated stimulation of β-cell proliferation and mTORC1 activity; MEX miR-mediated suppression of p53, MAFB, GLUT1; AMPK, PGC1α, ABCA1, TGFB2, SMAD2 and GISIS
White adipose tissue Promotion of white adipogenesis promoting obesity miR-148a-mediated inhibition of WNT1 and WNT10B; miR-21-mediated stimulation of adipocyte differentiation; miR-148a-mediated suppression of LDLR and ABCA1 compromising cholesterol homeostasis
Brown adipose tissue Promotion of beige/brown adipogenesis reducing obesity miR-30b/c-mediated stimulation of UCP1 expression
Hypothalamic centers Hyperphagia promoting obesity miR-148a-mediated suppression of CCKBR reducing satiety signaling
Brain Over-expression of α-synuclein and reduced CMA promoting neurodegenerative diseases miR-148a-DNMT1-mediated overexpression of α-synuclein, miR-21-mediated suppression of LAMP2A-dependent CMA
Malignant tumors Promotion of tumor growth and epithelial-mesenchymal transition miR-148a/miR-125b/miR-30-mediated suppression of p53; transfer of oncogenic miR-155 and miR-21; miR-30b/c-mediated suppression of RIP140;
MEX miR-mediated suppression of MeCP2, H3K27me3, opening of chromatin structure; increased permission of transcription

MEX, milk exosome; miR, microRNA; CMA, chaperone-mediated autophagy.


Acknowledgments

Funding: None.


Footnote

Provenance and Peer Review: This article was commissioned by the Guest Editors (Patrick Provost and Chenyu Zhang) for the series “Dietary MicroRNAs: Focus on Milk” published in ExRNA. The article has undergone external peer review.

Conflicts of Interest: All authors have completed the ICMJE uniform disclosure form (available at https://exrna.amegroups.com/article/view/10.21037/exrna-22-5/coif). The series “Dietary MicroRNAs: Focus on Milk” was commissioned by the editorial office without any funding or sponsorship. The authors have no other conflicts of interest to declare.

Ethical Statement: The authors are accountable for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Open Access Statement: This is an Open Access article distributed in accordance with the Creative Commons Attribution-NonCommercial-NoDerivs 4.0 International License (CC BY-NC-ND 4.0), which permits the non-commercial replication and distribution of the article with the strict proviso that no changes or edits are made and the original work is properly cited (including links to both the formal publication through the relevant DOI and the license). See: https://creativecommons.org/licenses/by-nc-nd/4.0/.


References

  1. Benmoussa A, Ly S, Shan ST, et al. A subset of extracellular vesicles carries the bulk of microRNAs in commercial dairy cow's milk. J Extracell Vesicles 2017;6:1401897. [Crossref] [PubMed]
  2. Blans K, Hansen MS, Sørensen LV, et al. Pellet-free isolation of human and bovine milk extracellular vesicles by size-exclusion chromatography. J Extracell Vesicles 2017;6:1294340. [Crossref] [PubMed]
  3. Bickmore DC, Miklavcic JJ. Characterization of Extracellular Vesicles Isolated From Human Milk Using a Precipitation-Based Method. Front Nutr 2020;7:22. [Crossref] [PubMed]
  4. Morozumi M, Izumi H, Shimizu T, et al. Comparison of isolation methods using commercially available kits for obtaining extracellular vesicles from cow milk. J Dairy Sci 2021;104:6463-71. [Crossref] [PubMed]
  5. Hu Y, Thaler J, Nieuwland R. Extracellular Vesicles in Human Milk. Pharmaceuticals (Basel) 2021;14:1050. [Crossref] [PubMed]
  6. Admyre C, Johansson SM, Qazi KR, et al. Exosomes with immune modulatory features are present in human breast milk. J Immunol 2007;179:1969-78. [Crossref] [PubMed]
  7. Zhou Q, Li M, Wang X, et al. Immune-related microRNAs are abundant in breast milk exosomes. Int J Biol Sci 2012;8:118-23. [Crossref] [PubMed]
  8. Gu Y, Li M, Wang T, et al. Lactation-related microRNA expression profiles of porcine breast milk exosomes. PLoS One 2012;7:e43691. [Crossref] [PubMed]
  9. Chen T, Xi QY, Ye RS, et al. Exploration of microRNAs in porcine milk exosomes. BMC Genomics 2014;15:100. [Crossref] [PubMed]
  10. Fernández-Messina L, Gutiérrez-Vázquez C, Rivas-García E, et al. Immunomodulatory role of microRNAs transferred by extracellular vesicles. Biol Cell 2015;107:61-77. [Crossref] [PubMed]
  11. Golan-Gerstl R, Elbaum Shiff Y, Moshayoff V, et al. Characterization and biological function of milk-derived miRNAs. Mol Nutr Food Res 2017; [Crossref] [PubMed]
  12. Zempleni J, Sukreet S, Zhou F, et al. Milk-Derived Exosomes and Metabolic Regulation. Annu Rev Anim Biosci 2019;7:245-62. [Crossref] [PubMed]
  13. Lönnerdal B. Human Milk MicroRNAs/Exosomes: Composition and Biological Effects. Nestle Nutr Inst Workshop Ser 2019;90:83-92. [Crossref] [PubMed]
  14. Benmoussa A, Laugier J, Beauparlant CJ, et al. Complexity of the microRNA transcriptome of cow milk and milk-derived extracellular vesicles isolated via differential ultracentrifugation. J Dairy Sci 2020;103:16-29. [Crossref] [PubMed]
  15. Chen Z, Xie Y, Luo J, et al. Milk exosome-derived miRNAs from water buffalo are implicated in immune response and metabolism process. BMC Vet Res 2020;16:123. [Crossref] [PubMed]
  16. Özdemir S. Identification and comparison of exosomal microRNAs in the milk and colostrum of two different cow breeds. Gene 2020;743:144609. [Crossref] [PubMed]
  17. Yun B, Kim Y, Park DJ, et al. Comparative analysis of dietary exosome-derived microRNAs from human, bovine and caprine colostrum and mature milk. J Anim Sci Technol 2021;63:593-602. [Crossref] [PubMed]
  18. Zeng B, Chen T, Luo J, et al. Exploration of Long Non-coding RNAs and Circular RNAs in Porcine Milk Exosomes. Front Genet 2020;11:652. [Crossref] [PubMed]
  19. Zeng B, Chen T, Luo JY, et al. Biological Characteristics and Roles of Noncoding RNAs in Milk-Derived Extracellular Vesicles. Adv Nutr 2021;12:1006-19. [Crossref] [PubMed]
  20. Cione E, Zambrini AS, Cannataro R. MicroRNAs and Extracellular Vesicles in Milk: RNA-Based Micronutrients? J Nutr 2021;151:1378-9. [Crossref] [PubMed]
  21. Tingö L, Ahlberg E, Johansson L, et al. Non-Coding RNAs in Human Breast Milk: A Systematic Review. Front Immunol 2021;12:725323. [Crossref] [PubMed]
  22. Shome S, Jernigan RL, Beitz DC, et al. Non-coding RNA in raw and commercially processed milk and putative targets related to growth and immune-response. BMC Genomics 2021;22:749. [Crossref] [PubMed]
  23. van Herwijnen MJC, Driedonks TAP, Snoek BL, et al. Abundantly Present miRNAs in Milk-Derived Extracellular Vesicles Are Conserved Between Mammals. Front Nutr 2018;5:81. [Crossref] [PubMed]
  24. Benmoussa A, Lee CH, Laffont B, et al. Commercial Dairy Cow Milk microRNAs Resist Digestion under Simulated Gastrointestinal Tract Conditions. J Nutr 2016;146:2206-15. [Crossref] [PubMed]
  25. Liao Y, Du X, Li J, et al. Human milk exosomes and their microRNAs survive digestion in vitro and are taken up by human intestinal cells. Mol Nutr Food Res 2017; [Crossref] [PubMed]
  26. Rani P, Vashisht M, Golla N, et al. Milk miRNAs encapsulated in exosomes are stable to human digestion and permeable to intestinal barrier in vitro. J Funct Foods 2017;34:431-9. [Crossref]
  27. Wolf T, Baier SR, Zempleni J. The Intestinal Transport of Bovine Milk Exosomes Is Mediated by Endocytosis in Human Colon Carcinoma Caco-2 Cells and Rat Small Intestinal IEC-6 Cells. J Nutr 2015;145:2201-6. [Crossref] [PubMed]
  28. Kusuma RJ, Manca S, Friemel T, et al. Human vascular endothelial cells transport foreign exosomes from cow's milk by endocytosis. Am J Physiol Cell Physiol 2016;310:C800-7. [Crossref] [PubMed]
  29. Wang L, Sadri M, Giraud D, et al. RNase H2-Dependent Polymerase Chain Reaction and Elimination of Confounders in Sample Collection, Storage, and Analysis Strengthen Evidence That microRNAs in Bovine Milk Are Bioavailable in Humans. J Nutr 2018;148:153-9. [Crossref] [PubMed]
  30. López de Las Hazas MC, Del Pozo-Acebo L, Hansen MS, et al. Dietary bovine milk miRNAs transported in extracellular vesicles are partially stable during GI digestion, are bioavailable and reach target tissues but need a minimum dose to impact on gene expression. Eur J Nutr 2022;61:1043-56. [Crossref] [PubMed]
  31. Lin D, Chen T, Xie M, et al. Oral Administration of Bovine and Porcine Milk Exosome Alter miRNAs Profiles in Piglet Serum. Sci Rep 2020;10:6983. [Crossref] [PubMed]
  32. Manca S, Upadhyaya B, Mutai E, et al. Milk exosomes are bioavailable and distinct microRNA cargos have unique tissue distribution patterns. Sci Rep 2018;8:11321. [Crossref] [PubMed]
  33. Aguilar-Lozano A, Baier S, Grove R, et al. Concentrations of Purine Metabolites Are Elevated in Fluids from Adults and Infants and in Livers from Mice Fed Diets Depleted of Bovine Milk Exosomes and their RNA Cargos. J Nutr 2018;148:1886-94. [Crossref] [PubMed]
  34. Betker JL, Angle BM, Graner MW, et al. The Potential of Exosomes From Cow Milk for Oral Delivery. J Pharm Sci 2019;108:1496-505. [Crossref] [PubMed]
  35. Sadri M, Shu J, Kachman SD, et al. Milk exosomes and miRNA cross the placenta and promote embryo survival in mice. Reproduction 2020;160:501-9. [Crossref] [PubMed]
  36. Sanwlani R, Fonseka P, Mathivanan S. Are Dietary Extracellular Vesicles Bioavailable and Functional in Consuming Organisms? Subcell Biochem 2021;97:509-21. [Crossref] [PubMed]
  37. Vaswani K, Koh YQ, Almughlliq FB, et al. A method for the isolation and enrichment of purified bovine milk exosomes. Reprod Biol 2017;17:341-8. [Crossref] [PubMed]
  38. Koh YQ, Almughlliq FB, Vaswani K, et al. Exosome enrichment by ultracentrifugation and size exclusion chromatography. Front Biosci (Landmark Ed) 2018;23:865-74. [Crossref] [PubMed]
  39. Wijenayake S, Eisha S, Tawhidi Z, et al. Comparison of methods for pre-processing, exosome isolation, and RNA extraction in unpasteurized bovine and human milk. PLoS One 2021;16:e0257633. [Crossref] [PubMed]
  40. Arntz OJ, Pieters BC, Oliveira MC, et al. Oral administration of bovine milk derived extracellular vesicles attenuates arthritis in two mouse models. Mol Nutr Food Res 2015;59:1701-12. [Crossref] [PubMed]
  41. Pieters BCH, Arntz OJ, Aarts J, et al. Bovine Milk-Derived Extracellular Vesicles Inhibit Catabolic and Inflammatory Processes in Cartilage from Osteoarthritis Patients. Mol Nutr Food Res 2022;66:e2100764. [Crossref] [PubMed]
  42. Oliveira MC, Pieters BCH, Guimarães PB, et al. Bovine Milk Extracellular Vesicles Are Osteoprotective by Increasing Osteocyte Numbers and Targeting RANKL/OPG System in Experimental Models of Bone Loss. Front Bioeng Biotechnol 2020;8:891. [Crossref] [PubMed]
  43. Go G, Jeon J, Lee G, et al. Bovine milk extracellular vesicles induce the proliferation and differentiation of osteoblasts and promote osteogenesis in rats. J Food Biochem 2021;45:e13705. [Crossref] [PubMed]
  44. Xie MY, Chen T, Xi QY, et al. Porcine milk exosome miRNAs protect intestinal epithelial cells against deoxynivalenol-induced damage. Biochem Pharmacol 2020;175:113898. [Crossref] [PubMed]
  45. Xie MY, Hou LJ, Sun JJ, et al. Porcine Milk Exosome MiRNAs Attenuate LPS-Induced Apoptosis through Inhibiting TLR4/NF-κB and p53 Pathways in Intestinal Epithelial Cells. J Agric Food Chem 2019;67:9477-91. [Crossref] [PubMed]
  46. Li B, Hock A, Wu RY, et al. Bovine milk-derived exosomes enhance goblet cell activity and prevent the development of experimental necrotizing enterocolitis. PLoS One 2019;14:e0211431. [Crossref] [PubMed]
  47. Stremmel W, Weiskirchen R, Melnik BC. Milk Exosomes Prevent Intestinal Inflammation in a Genetic Mouse Model of Ulcerative Colitis: A Pilot Experiment. Inflamm Intest Dis 2020;5:117-23. [Crossref] [PubMed]
  48. Benmoussa A, Diallo I, Salem M, et al. Concentrates of two subsets of extracellular vesicles from cow's milk modulate symptoms and inflammation in experimental colitis. Sci Rep 2019;9:14661. [Crossref] [PubMed]
  49. Reif S, Elbaum-Shiff Y, Koroukhov N, et al. Cow and Human Milk-Derived Exosomes Ameliorate Colitis in DSS Murine Model. Nutrients 2020;12:2589. [Crossref] [PubMed]
  50. Zhang C, Lu X, Hu J, et al. Bovine Milk Exosomes Alleviate Cardiac Fibrosis via Enhancing Angiogenesis In Vivo and In Vitro. J Cardiovasc Transl Res 2022;15:560-70. [Crossref] [PubMed]
  51. Melnik BC, Weiskirchen R, Stremmel W, et al. Narrative review of bovine milk exosomes as a potential therapeutic option for hepatic fibrosis. Dig Med Res 2022;5:14. [Crossref]
  52. Kim H, Kim DE, Han G, et al. Harnessing the Natural Healing Power of Colostrum: Bovine Milk-Derived Extracellular Vesicles from Colostrum Facilitating the Transition from Inflammation to Tissue Regeneration for Accelerating Cutaneous Wound Healing. Adv Healthc Mater 2022;11:e2102027. [Crossref] [PubMed]
  53. Ahn G, Kim YH, Ahn JY. Multifaceted effects of milk-exosomes (Mi-Exo) as a modulator of scar-free wound healing. Nanoscale Adv 2021;3:528. [Crossref] [PubMed]
  54. Yan C, Chen J, Wang C, et al. Milk exosomes-mediated miR-31-5p delivery accelerates diabetic wound healing through promoting angiogenesis. Drug Deliv 2022;29:214-28. [Crossref] [PubMed]
  55. Feng X, Chen X, Zheng X, et al. Latest Trend of Milk Derived Exosomes: Cargos, Functions, and Applications. Front Nutr 2021;8:747294. [Crossref] [PubMed]
  56. Munagala R, Aqil F, Jeyabalan J, et al. Bovine milk-derived exosomes for drug delivery. Cancer Lett 2016;371:48-61. [Crossref] [PubMed]
  57. Sedykh S, Kuleshova A, Nevinsky G. Milk Exosomes: Perspective Agents for Anticancer Drug Delivery. Int J Mol Sci 2020;21:6646. [Crossref] [PubMed]
  58. Adriano B, Cotto NM, Chauhan N, et al. Milk exosomes: Nature's abundant nanoplatform for theranostic applications. Bioact Mater 2021;6:2479-90. [Crossref] [PubMed]
  59. Li S, Tang Y, Dou Y. The Potential of Milk-Derived Exosomes for Drug Delivery. Curr Drug Deliv 2021;18:688-99. [Crossref] [PubMed]
  60. Zhong J, Xia B, Shan S, et al. High-quality milk exosomes as oral drug delivery system. Biomaterials 2021;277:121126. [Crossref] [PubMed]
  61. Del Pozo-Acebo L, Hazas MLL, Tomé-Carneiro J, et al. Bovine Milk-Derived Exosomes as a Drug Delivery Vehicle for miRNA-Based Therapy. Int J Mol Sci 2021;22:1105. [Crossref] [PubMed]
  62. Bryder L. From breast to bottle: a history of modern infant feeding. Endeavour 2009;33:54-9. [Crossref] [PubMed]
  63. Melnik BC, John SM, Schmitz G. Milk is not just food but most likely a genetic transfection system activating mTORC1 signaling for postnatal growth. Nutr J 2013;12:103. [Crossref] [PubMed]
  64. Melnik BC, John SM, Schmitz G. Milk: an exosomal microRNA transmitter promoting thymic regulatory T cell maturation preventing the development of atopy? J Transl Med 2014;12:43. [Crossref] [PubMed]
  65. Zempleni J, Aguilar-Lozano A, Sadri M, et al. Biological Activities of Extracellular Vesicles and Their Cargos from Bovine and Human Milk in Humans and Implications for Infants. J Nutr 2017;147:3-10. [Crossref] [PubMed]
  66. van Esch BCAM, Porbahaie M, Abbring S, et al. The Impact of Milk and Its Components on Epigenetic Programming of Immune Function in Early Life and Beyond: Implications for Allergy and Asthma. Front Immunol 2020;11:2141. [Crossref] [PubMed]
  67. Röszer T. Mother-to-Child Signaling through Breast Milk Biomolecules. Biomolecules 2021;11:1743. [Crossref] [PubMed]
  68. Röszer T. Co-Evolution of Breast Milk Lipid Signaling and Thermogenic Adipose Tissue. Biomolecules 2021;11:1705. [Crossref] [PubMed]
  69. Melnik BC, Stremmel W, Weiskirchen R, et al. Exosome-Derived MicroRNAs of Human Milk and Their Effects on Infant Health and Development. Biomolecules 2021;11:851. [Crossref] [PubMed]
  70. Jiang X, You L, Zhang Z, et al. Biological Properties of Milk-Derived Extracellular Vesicles and Their Physiological Functions in Infant. Front Cell Dev Biol 2021;9:693534. [Crossref] [PubMed]
  71. Camacho-Morales A, Caba M, García-Juárez M, et al. Breastfeeding Contributes to Physiological Immune Programming in the Newborn. Front Pediatr 2021;9:744104. [Crossref] [PubMed]
  72. Ma J, Wang C, Long K, et al. Exosomal microRNAs in giant panda (Ailuropoda melanoleuca) breast milk: potential maternal regulators for the development of newborn cubs. Sci Rep 2017;7:3507. [Crossref] [PubMed]
  73. Ozkan H, Tuzun F, Taheri S, et al. Epigenetic Programming Through Breast Milk and Its Impact on Milk-Siblings Mating. Front Genet 2020;11:569232. [Crossref] [PubMed]
  74. Sherwood WB, Kothalawala DM, Kadalayil L, et al. Epigenome-Wide Association Study Reveals Duration of Breastfeeding Is Associated with Epigenetic Differences in Children. Int J Environ Res Public Health 2020;17:3569. [Crossref] [PubMed]
  75. Briollais L, Rustand D, Allard C, et al. DNA methylation mediates the association between breastfeeding and early-life growth trajectories. Clin Epigenetics 2021;13:231. [Crossref] [PubMed]
  76. Rubio M, Bustamante M, Hernandez-Ferrer C, et al. Circulating miRNAs, isomiRs and small RNA clusters in human plasma and breast milk. PLoS One 2018;13:e0193527. [Crossref] [PubMed]
  77. Reif S, Elbaum Shiff Y, Golan-Gerstl R. Milk-derived exosomes (MDEs) have a different biological effect on normal fetal colon epithelial cells compared to colon tumor cells in a miRNA-dependent manner. J Transl Med 2019;17:325. [Crossref] [PubMed]
  78. Samuel M, Fonseka P, Sanwlani R, et al. Oral administration of bovine milk-derived extracellular vesicles induces senescence in the primary tumor but accelerates cancer metastasis. Nat Commun 2021;12:3950. [Crossref] [PubMed]
  79. Mecocci S, Pietrucci D, Milanesi M, et al. Transcriptomic Characterization of Cow, Donkey and Goat Milk Extracellular Vesicles Reveals Their Anti-Inflammatory and Immunomodulatory Potential. Int J Mol Sci 2021;22:12759. [Crossref] [PubMed]
  80. Grossen P, Portmann M, Koller E, et al. Evaluation of bovine milk extracellular vesicles for the delivery of locked nucleic acid antisense oligonucleotides. Eur J Pharm Biopharm 2021;158:198-210. [Crossref] [PubMed]
  81. Myrzabekova M, Labeit S, Niyazova R, et al. Identification of Bovine miRNAs with the Potential to Affect Human Gene Expression. Front Genet 2021;12:705350. [Crossref] [PubMed]
  82. Melnik BC, Weiskirchen R, Stremmel W, et al. Letter to the editor regarding "Dietary bovine milk miRNAs transported in extracellular vesicles are partially stable during GI digestion, are bioavailable and reach target tissues but need a minimum dose to impact on gene expression". Eur J Nutr 2022;61:1695-6. [Crossref] [PubMed]
  83. López de Las Hazas MC, Del Pozo-Acebo L, Dávalos A. Response to: Letter to the editor regarding "Dietary bovine milk miRNAs transported in extracellular vesicles are partially stable during GI digestion, are bioavailable and reach target tissues but need a minimum dose to impact on gene expression". Eur J Nutr 2022;61:1697-8. [Crossref] [PubMed]
  84. Melnik BC, Schmitz G. Milk's Role as an Epigenetic Regulator in Health and Disease. Diseases 2017;5:12. [Crossref] [PubMed]
  85. Melnik BC, Schmitz G. MicroRNAs: Milk's epigenetic regulators. Best Pract Res Clin Endocrinol Metab 2017;31:427-42. [Crossref] [PubMed]
  86. Guo MM, Zhang K, Zhang JH. Human Breast Milk-Derived Exosomal miR-148a-3p Protects Against Necrotizing Enterocolitis by Regulating p53 and Sirtuin 1. Inflammation 2022;45:1254-68. [Crossref] [PubMed]
  87. Melnik BC. Milk: an epigenetic amplifier of FTO-mediated transcription? Implications for Western diseases. J Transl Med 2015;13:385. [Crossref] [PubMed]
  88. Melnik BC. Milk disrupts p53 and DNMT1, the guardians of the genome: implications for acne vulgaris and prostate cancer. Nutr Metab (Lond) 2017;14:55. [Crossref] [PubMed]
  89. Curry A. Archaeology: The milk revolution. Nature 2013;500:20-2. [Crossref] [PubMed]
  90. Melnik BC, Schmitz G. Pasteurized non-fermented cow's milk but not fermented milk is a promoter of mTORC1-driven aging and increased mortality. Ageing Res Rev 2021;67:101270. [Crossref] [PubMed]
  91. Yu S, Zhao Z, Sun L, et al. Fermentation Results in Quantitative Changes in Milk-Derived Exosomes and Different Effects on Cell Growth and Survival. J Agric Food Chem 2017;65:1220-8. [Crossref] [PubMed]
  92. Kleinjan M, van Herwijnen MJ, Libregts SF, et al. Regular Industrial Processing of Bovine Milk Impacts the Integrity and Molecular Composition of Extracellular Vesicles. J Nutr 2021;151:1416-25. [Crossref] [PubMed]
  93. Zhang Y, Xu Q, Hou J, et al. Loss of bioactive microRNAs in cow's milk by ultra-high-temperature treatment but not by pasteurization treatment. J Sci Food Agric 2022;102:2676-85. [Crossref] [PubMed]
  94. Heppe DH, van Dam RM, Willemsen SP, et al. Maternal milk consumption, fetal growth, and the risks of neonatal complications: the Generation R Study. Am J Clin Nutr 2011;94:501-9. [Crossref] [PubMed]
  95. Olsen SF, Halldorsson TI, Willett WC, et al. Milk consumption during pregnancy is associated with increased infant size at birth: prospective cohort study. Am J Clin Nutr 2007;86:1104-10. [Crossref] [PubMed]
  96. Brantsæter AL, Olafsdottir AS, Forsum E, et al. Does milk and dairy consumption during pregnancy influence fetal growth and infant birthweight? A systematic literature review. Food Nutr Res 2012; [Crossref] [PubMed]
  97. Melnik BC, John SM, Schmitz G. Milk consumption during pregnancy increases birth weight, a risk factor for the development of diseases of civilization. J Transl Med 2015;13:13. [Crossref] [PubMed]
  98. Achón M, Úbeda N, García-González Á, et al. Effects of Milk and Dairy Product Consumption on Pregnancy and Lactation Outcomes: A Systematic Review. Adv Nutr 2019;10:S74-87. [Crossref] [PubMed]
  99. Roland MC, Friis CM, Voldner N, et al. Fetal growth versus birthweight: the role of placenta versus other determinants. PLoS One 2012;7:e39324. [Crossref] [PubMed]
  100. Chen X, Gao C, Li H, et al. Identification and characterization of microRNAs in raw milk during different periods of lactation, commercial fluid, and powdered milk products. Cell Res 2010;20:1128-37. [Crossref] [PubMed]
  101. Smyczynska U, Bartlomiejczyk MA, Stanczak MM, et al. Impact of processing method on donated human breast milk microRNA content. PLoS One 2020;15:e0236126. [Crossref] [PubMed]
  102. Balaguer N, Moreno I, Herrero M, et al. MicroRNA-30d deficiency during preconception affects endometrial receptivity by decreasing implantation rates and impairing fetal growth. Am J Obstet Gynecol 2019;221:46.e1-46.e16. [Crossref] [PubMed]
  103. Jiang H, Wu W, Zhang M, et al. Aberrant upregulation of miR-21 in placental tissues of macrosomia. J Perinatol 2014;34:658-63. [Crossref] [PubMed]
  104. Zhang JT, Cai QY, Ji SS, et al. Decreased miR-143 and increased miR-21 placental expression levels are associated with macrosomia. Mol Med Rep 2016;13:3273-80. [Crossref] [PubMed]
  105. Melnik BC. Milk--A Nutrient System of Mammalian Evolution Promoting mTORC1-Dependent Translation. Int J Mol Sci 2015;16:17048-87. [Crossref] [PubMed]
  106. Liu Z, Xie Y, Guo J, et al. Comparison of porcine milk microRNA expression in milk exosomes versus whole swine milk and prediction of target genes. Arch Anim Breed 2022;65:37-46. [Crossref] [PubMed]
  107. Wen HY, Abbasi S, Kellems RE, et al. mTOR: a placental growth signaling sensor. Placenta 2005;26 Suppl A:S63-9.
  108. Roos S, Lagerlöf O, Wennergren M, et al. Regulation of amino acid transporters by glucose and growth factors in cultured primary human trophoblast cells is mediated by mTOR signaling. Am J Physiol Cell Physiol 2009;297:C723-31. [Crossref] [PubMed]
  109. Jansson T, Aye IL, Goberdhan DC. The emerging role of mTORC1 signaling in placental nutrient-sensing. Placenta 2012;33:e23-9. [Crossref] [PubMed]
  110. Rosario FJ, Kanai Y, Powell TL, et al. Mammalian target of rapamycin signalling modulates amino acid uptake by regulating transporter cell surface abundance in primary human trophoblast cells. J Physiol 2013;591:609-25. [Crossref] [PubMed]
  111. Xu J, Lu C, Wang J, et al. Regulation of Human Trophoblast GLUT3 Glucose Transporter by Mammalian Target of Rapamycin Signaling. Int J Mol Sci 2015;16:13815-28. [Crossref] [PubMed]
  112. Rosario FJ, Powell TL, Gupta MB, et al. mTORC1 Transcriptional Regulation of Ribosome Subunits, Protein Synthesis, and Molecular Transport in Primary Human Trophoblast Cells. Front Cell Dev Biol 2020;8:583801. [Crossref] [PubMed]
  113. Chen T, Xie MY, Sun JJ, et al. Porcine milk-derived exosomes promote proliferation of intestinal epithelial cells. Sci Rep 2016;6:33862. [Crossref] [PubMed]
  114. Hock A, Miyake H, Li B, et al. Breast milk-derived exosomes promote intestinal epithelial cell growth. J Pediatr Surg 2017;52:755-9. [Crossref] [PubMed]
  115. Kahn S, Liao Y, Du X, et al. Exosomal MicroRNAs in Milk from Mothers Delivering Preterm Infants Survive in Vitro Digestion and Are Taken Up by Human Intestinal Cells. Mol Nutr Food Res 2018;62:e1701050. [Crossref] [PubMed]
  116. Zhou Y, Yu Z, Wang X, et al. Exosomal circRNAs contribute to intestinal development via the VEGF signalling pathway in human term and preterm colostrum. Aging (Albany NY) 2021;13:11218-33. [Crossref] [PubMed]
  117. Gao HN, Ren FZ, Wen PC, et al. Yak milk-derived exosomal microRNAs regulate intestinal epithelial cells on proliferation in hypoxic environment. J Dairy Sci 2021;104:1291-303. [Crossref] [PubMed]
  118. Gao R, Zhang R, Qian T, et al. A comparison of exosomes derived from different periods breast milk on protecting against intestinal organoid injury. Pediatr Surg Int 2019;35:1363-8. [Crossref] [PubMed]
  119. Miyake H, Lee C, Chusilp S, et al. Human breast milk exosomes attenuate intestinal damage. Pediatr Surg Int 2020;36:155-63. [Crossref] [PubMed]
  120. Hou L, Tong X, Lin S, et al. MiR-221/222 Ameliorates Deoxynivalenol-Induced Apoptosis and Proliferation Inhibition in Intestinal Epithelial Cells by Targeting PTEN. Front Cell Dev Biol 2021;9:652939. [Crossref] [PubMed]
  121. Maghraby MK, Li B, Chi L, et al. Extracellular vesicles isolated from milk can improve gut barrier dysfunction induced by malnutrition. Sci Rep 2021;11:7635. [Crossref] [PubMed]
  122. He S, Liu G, Zhu X. Human breast milk-derived exosomes may help maintain intestinal epithelial barrier integrity. Pediatr Res 2021;90:366-72. [Crossref] [PubMed]
  123. Aarts J, Boleij A, Pieters BCH, et al. Flood Control: How Milk-Derived Extracellular Vesicles Can Help to Improve the Intestinal Barrier Function and Break the Gut-Joint Axis in Rheumatoid Arthritis. Front Immunol 2021;12:703277. [Crossref] [PubMed]
  124. Wu D, Kittana H, Shu J, et al. Dietary Depletion of Milk Exosomes and Their MicroRNA Cargos Elicits a Depletion of miR-200a-3p and Elevated Intestinal Inflammation and Chemokine (C-X-C Motif) Ligand 9 Expression in Mdr1a-/- Mice. Curr Dev Nutr 2019;3:nzz122. [Crossref] [PubMed]
  125. Gao HN, Hu H, Wen PC, et al. Yak milk-derived exosomes alleviate lipopolysaccharide-induced intestinal inflammation by inhibiting PI3K/AKT/C3 pathway activation. J Dairy Sci 2021;104:8411-24. [Crossref] [PubMed]
  126. Wang L, Shi Z, Wang X, et al. Protective effects of bovine milk exosomes against oxidative stress in IEC-6 cells. Eur J Nutr 2021;60:317-27. [Crossref] [PubMed]
  127. Wang L, Wang X, Shi Z, et al. Bovine milk exosomes attenuate the alteration of purine metabolism and energy status in IEC-6 cells induced by hydrogen peroxide. Food Chem 2021;350:129142. [Crossref] [PubMed]
  128. Le Doare K, Holder B, Bassett A, et al. Mother's Milk: A Purposeful Contribution to the Development of the Infant Microbiota and Immunity. Front Immunol 2018;9:361. [Crossref] [PubMed]
  129. Carr LE, Virmani MD, Rosa F, et al. Role of Human Milk Bioactives on Infants' Gut and Immune Health. Front Immunol 2021;12:604080. [Crossref] [PubMed]
  130. Askenase PW. Exosomes provide unappreciated carrier effects that assist transfers of their miRNAs to targeted cells; I. They are 'The Elephant in the Room'. RNA Biol 2021;18:2038-53. [Crossref] [PubMed]
  131. Tong L, Hao H, Zhang X, et al. Oral administration of bovine milk‐derived mice. Mol Nutr Food Res 2020;64:e1901251. [Crossref] [PubMed]
  132. Zhou F, Paz HA, Sadri M, et al. Dietary bovine milk exosomes elicit changes in bacterial communities in C57BL/6 mice. Am J Physiol Gastrointest Liver Physiol 2019;317:G618-24. [Crossref] [PubMed]
  133. Jiang K, Yang J, Yang C, et al. miR-148a suppresses inflammation in lipopolysaccharide-induced endometritis. J Cell Mol Med 2020;24:405-17. [Crossref] [PubMed]
  134. Liu X, Zhan Z, Xu L, et al. MicroRNA-148/152 impair innate response and antigen presentation of TLR-triggered dendritic cells by targeting CaMKIIα. J Immunol 2010;185:7244-51. [Crossref] [PubMed]
  135. Zheng D, Li Z, Wei X, et al. Role of miR-148a in Mitigating Hepatic Ischemia-Reperfusion Injury by Repressing the TLR4 Signaling Pathway via Targeting CaMKIIα in Vivo and in Vitro. Cell Physiol Biochem 2018;49:2060-72. [Crossref] [PubMed]
  136. Zhu Y, Gu L, Li Y, et al. miR-148a inhibits colitis and colitis-associated tumorigenesis in mice. Cell Death Differ 2017;24:2199-209. [Crossref] [PubMed]
  137. Patel V, Carrion K, Hollands A, et al. The stretch responsive microRNA miR-148a-3p is a novel repressor of IKBKB, NF-κB signaling, and inflammatory gene expression in human aortic valve cells. FASEB J 2015;29:1859-68. [Crossref] [PubMed]
  138. Yin M, Lu J, Guo Z, et al. Reduced SULT2B1b expression alleviates ox-LDL-induced inflammation by upregulating miR-148-3P via inhibiting the IKKβ/NF-κB pathway in macrophages. Aging (Albany NY) 2021;13:3428-42. [Crossref] [PubMed]
  139. Raso A, Dirkx E, Philippen LE, et al. Therapeutic Delivery of miR-148a Suppresses Ventricular Dilation in Heart Failure. Mol Ther 2019;27:584-99. [Crossref] [PubMed]
  140. Chen W, Wang X, Yan X, et al. The emerging role of exosomes in the pathogenesis, prognosis and treatment of necrotizing enterocolitis. Am J Transl Res 2020;12:7020-33. [PubMed]
  141. Pisano C, Galley J, Elbahrawy M, et al. Human Breast Milk-Derived Extracellular Vesicles in the Protection Against Experimental Necrotizing Enterocolitis. J Pediatr Surg 2020;55:54-8. [Crossref] [PubMed]
  142. Toker A, Engelbert D, Garg G, et al. Active demethylation of the Foxp3 locus leads to the generation of stable regulatory T cells within the thymus. J Immunol 2013;190:3180-8. [Crossref] [PubMed]
  143. Polansky JK, Kretschmer K, Freyer J, et al. DNA methylation controls Foxp3 gene expression. Eur J Immunol 2008;38:1654-63. [Crossref] [PubMed]
  144. Lal G, Bromberg JS. Epigenetic mechanisms of regulation of Foxp3 expression. Blood 2009;114:3727-35. [Crossref] [PubMed]
  145. Braga M, Quecchia C, Cavallucci E, et al. T regulatory cells in allergy. Int J Immunopathol Pharmacol 2011;24:55S-64S. [PubMed]
  146. Eggenhuizen PJ, Ng BH, Ooi JD. Treg Enhancing Therapies to Treat Autoimmune Diseases. Int J Mol Sci 2020;21:7015. [Crossref] [PubMed]
  147. Zhou F, Ebea P, Mutai E, et al. Small Extracellular Vesicles in Milk Cross the Blood-Brain Barrier in Murine Cerebral Cortex Endothelial Cells and Promote Dendritic Complexity in the Hippocampus and Brain Function in C57BL/6J Mice. Front Nutr 2022;9:838543. [Crossref] [PubMed]
  148. Mutai E, Zhou F, Zempleni J. Depletion of dietary bovine milk exosomes impairs sensorimotor gating and spatial learning in C57BL/6 mice. FASEB J 2018;31:S1.
  149. Jakowec MW, Donaldson DM, Barba J, et al. Postnatal expression of alpha-synuclein protein in the rodent substantia nigra and striatum. Dev Neurosci 2001;23:91-9. [Crossref] [PubMed]
  150. Varkey J, Isas JM, Mizuno N, et al. Membrane curvature induction and tubulation are common features of synucleins and apolipoproteins. J Biol Chem 2010;285:32486-93. [Crossref] [PubMed]
  151. Matsumoto L, Takuma H, Tamaoka A, et al. CpG demethylation enhances alpha-synuclein expression and affects the pathogenesis of Parkinson's disease. PLoS One 2010;5:e15522. [Crossref] [PubMed]
  152. Monks DT, Palanisamy A. Oxytocin: at birth and beyond. A systematic review of the long-term effects of peripartum oxytocin. Anaesthesia 2021;76:1526-37. [Crossref] [PubMed]
  153. Gutman-Ido E, Reif S, Musseri M, et al. Oxytocin Regulates the Expression of Selected Colostrum-derived microRNAs. J Pediatr Gastroenterol Nutr 2022;74:e8-e15. [Crossref] [PubMed]
  154. Shah KB, Chernausek SD, Garman LD, et al. Human Milk Exosomal MicroRNA: Associations with Maternal Overweight/Obesity and Infant Body Composition at 1 Month of Life. Nutrients 2021;13:1091. [Crossref] [PubMed]
  155. Shah KB, Fields DA, Pezant NP, et al. Gestational Diabetes Mellitus Is Associated with Altered Abundance of Exosomal MicroRNAs in Human Milk. Clin Ther 2022;44:172-185.e1. Erratum in: Clin Ther 2022 doi: 10.1016/j.clinthera.2022.04.011. [Crossref] [PubMed]
  156. Hu F, Wang M, Xiao T, et al. miR-30 promotes thermogenesis and the development of beige fat by targeting RIP140. Diabetes 2015;64:2056-68. [Crossref] [PubMed]
  157. Kiskinis E, Hallberg M, Christian M, et al. RIP140 directs histone and DNA methylation to silence Ucp1 expression in white adipocytes. EMBO J 2007;26:4831-40. [Crossref] [PubMed]
  158. Kiskinis E, Chatzeli L, Curry E, et al. RIP140 represses the "brown-in-white" adipocyte program including a futile cycle of triacylglycerol breakdown and synthesis. Mol Endocrinol 2014;28:344-56. [Crossref] [PubMed]
  159. Garzon R, Liu S, Fabbri M, et al. MicroRNA-29b induces global DNA hypomethylation and tumor suppressor gene reexpression in acute myeloid leukemia by targeting directly DNMT3A and 3B and indirectly DNMT1. Blood 2009;113:6411-8. [Crossref] [PubMed]
  160. Zhang Z, Cao Y, Zhai Y, et al. MicroRNA-29b regulates DNA methylation by targeting Dnmt3a/3b and Tet1/2/3 in porcine early embryo development. Dev Growth Differ 2018;60:197-204. [Crossref] [PubMed]
  161. Izumi H, Tsuda M, Sato Y, et al. Bovine milk exosomes contain microRNA and mRNA and are taken up by human macrophages. J Dairy Sci 2015;98:2920-33. [Crossref] [PubMed]
  162. TargetScan Human Release 8.0. Human NRIP1 ENST00000400199.1. Available online: http://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000400199.1&taxid=9606&showcnc=0&shownc=0&shownc_nc=&showncf1=&showncf2=&subset=1
  163. Lane DP. Cancer. p53, guardian of the genome. Nature 1992;358:15-6. [Crossref] [PubMed]
  164. Marcel V, Nguyen Van Long F, Diaz JJ. 40 Years of Research Put p53 in Translation. Cancers (Basel) 2018;10:152. [Crossref] [PubMed]
  165. Hoh J, Jin S, Parrado T, et al. The p53MH algorithm and its application in detecting p53-responsive genes. Proc Natl Acad Sci U S A 2002;99:8467-72. [Crossref] [PubMed]
  166. Pappas K, Xu J, Zairis S, et al. p53 Maintains Baseline Expression of Multiple Tumor Suppressor Genes. Mol Cancer Res 2017;15:1051-62. [Crossref] [PubMed]
  167. Feng Z. p53 regulation of the IGF-1/AKT/mTOR pathways and the endosomal compartment. Cold Spring Harb Perspect Biol 2010;2:a001057. [Crossref] [PubMed]
  168. Krstic J, Reinisch I, Schupp M, et al. p53 Functions in Adipose Tissue Metabolism and Homeostasis. Int J Mol Sci 2018;19:2622. [Crossref] [PubMed]
  169. White E. Autophagy and p53. Cold Spring Harb Perspect Med 2016;6:a026120. [Crossref] [PubMed]
  170. Lima S, Takabe K, Newton J, et al. TP53 is required for BECN1- and ATG5-dependent cell death induced by sphingosine kinase 1 inhibition. Autophagy 2018;14:942-57. [Crossref] [PubMed]
  171. Kruiswijk F, Labuschagne CF, Vousden KH. p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat Rev Mol Cell Biol 2015;16:393-405. [Crossref] [PubMed]
  172. Laptenko O, Prives C. p53: master of life, death, and the epigenome. Genes Dev 2017;31:955-6. [Crossref] [PubMed]
  173. Hoffman WH, Biade S, Zilfou JT, et al. Transcriptional repression of the anti-apoptotic survivin gene by wild type p53. J Biol Chem 2002;277:3247-57. [Crossref] [PubMed]
  174. Le MT, Shyh-Chang N, Khaw SL, et al. Conserved regulation of p53 network dosage by microRNA-125b occurs through evolving miRNA-target gene pairs. PLoS Genet 2011;7:e1002242. [Crossref] [PubMed]
  175. Baddela VS, Nayan V, Rani P, et al. Physicochemical Biomolecular Insights into Buffalo Milk-Derived Nanovesicles. Appl Biochem Biotechnol 2016;178:544-57. [Crossref] [PubMed]
  176. Kumar M, Lu Z, Takwi AA, et al. Negative regulation of the tumor suppressor p53 gene by microRNAs. Oncogene 2011;30:843-53. [Crossref] [PubMed]
  177. Ogawara Y, Kishishita S, Obata T, et al. Akt enhances Mdm2-mediated ubiquitination and degradation of p53. J Biol Chem 2002;277:21843-50. [Crossref] [PubMed]
  178. Haupt Y, Maya R, Kazaz A, et al. Mdm2 promotes the rapid degradation of p53. Nature 1997;387:296-9. [Crossref] [PubMed]
  179. Haupt S, Mejía-Hernández JO, Vijayakumaran R, et al. The long and the short of it: the MDM4 tail so far. J Mol Cell Biol 2019;11:231-44. [Crossref] [PubMed]
  180. Li E, Zhang Y. DNA methylation in mammals. Cold Spring Harb Perspect Biol 2014;6:a019133. [Crossref] [PubMed]
  181. Goyal D, Limesand SW, Goyal R. Epigenetic responses and the developmental origins of health and disease. J Endocrinol 2019;242:T105-19. [Crossref] [PubMed]
  182. Estève PO, Chin HG, Pradhan S. Human maintenance DNA (cytosine-5)-methyltransferase and p53 modulate expression of p53-repressed promoters. Proc Natl Acad Sci U S A 2005;102:1000-5. [Crossref] [PubMed]
  183. Della Ragione F, Vacca M, Fioriniello S, et al. MECP2, a multi-talented modulator of chromatin architecture. Brief Funct Genomics 2016;15:420-31. [Crossref] [PubMed]
  184. Ballestar E, Pile LA, Wassarman DA, et al. A Drosophila MBD family member is a transcriptional corepressor associated with specific genes. Eur J Biochem 2001;268:5397-406. [Crossref] [PubMed]
  185. Fujita N, Watanabe S, Ichimura T, et al. Methyl-CpG binding domain 1 (MBD1) interacts with the Suv39h1-HP1 heterochromatic complex for DNA methylation-based transcriptional repression. J Biol Chem 2003;278:24132-8. [Crossref] [PubMed]
  186. Schmidt A, Zhang H, Cardoso MC. MeCP2 and Chromatin Compartmentalization. Cells 2020;9:878. [Crossref] [PubMed]
  187. Kimura H, Shiota K. Methyl-CpG-binding protein, MeCP2, is a target molecule for maintenance DNA methyltransferase, Dnmt1. J Biol Chem 2003;278:4806-12. [Crossref] [PubMed]
  188. Lee W, Kim J, Yun JM, et al. MeCP2 regulates gene expression through recognition of H3K27me3. Nat Commun 2020;11:3140. [Crossref] [PubMed]
  189. Di Croce L, Helin K. Transcriptional regulation by Polycomb group proteins. Nat Struct Mol Biol 2013;20:1147-55. [Crossref] [PubMed]
  190. Simon JA, Kingston RE. Occupying chromatin: Polycomb mechanisms for getting to genomic targets, stopping transcriptional traffic, and staying put. Mol Cell 2013;49:808-24. [Crossref] [PubMed]
  191. Zha L, Li F, Wu R, et al. The Histone Demethylase UTX Promotes Brown Adipocyte Thermogenic Program Via Coordinated Regulation of H3K27 Demethylation and Acetylation. J Biol Chem 2015;290:25151-63. [Crossref] [PubMed]
  192. Li F, Jing J, Movahed M, et al. Epigenetic interaction between UTX and DNMT1 regulates diet-induced myogenic remodeling in brown fat. Nat Commun 2021;12:6838. [Crossref] [PubMed]
  193. Margueron R, Reinberg D. The Polycomb complex PRC2 and its mark in life. Nature 2011;469:343-9. [Crossref] [PubMed]
  194. Yuan W, Wu T, Fu H, et al. Dense chromatin activates Polycomb repressive complex 2 to regulate H3 lysine 27 methylation. Science 2012;337:971-5. [Crossref] [PubMed]
  195. Yu JR, Lee CH, Oksuz O, et al. PRC2 is high maintenance. Genes Dev 2019;33:903-35. [Crossref] [PubMed]
  196. Guajardo L, Aguilar R, Bustos FJ, et al. Downregulation of the Polycomb-Associated Methyltransferase Ezh2 during Maturation of Hippocampal Neurons Is Mediated by MicroRNAs Let-7 and miR-124. Int J Mol Sci 2020;21:8472. [Crossref] [PubMed]
  197. Kong D, Heath E, Chen W, et al. Loss of let-7 up-regulates EZH2 in prostate cancer consistent with the acquisition of cancer stem cell signatures that are attenuated by BR-DIM. PLoS One 2012;7:e33729. [Crossref] [PubMed]
  198. Viré E, Brenner C, Deplus R, et al. The Polycomb group protein EZH2 directly controls DNA methylation. Nature 2006;439:871-4. [Crossref] [PubMed]
  199. Oshima M, Hasegawa N, Mochizuki-Kashio M, et al. Ezh2 regulates the Lin28/let-7 pathway to restrict activation of fetal gene signature in adult hematopoietic stem cells. Exp Hematol 2016;44:282-96.e3. [Crossref] [PubMed]
  200. TagetScanHuman. Release 8.0. Human EED ENST00000327320.4. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000327320.4&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  201. Song PP, Hu Y, Liu CM, et al. Embryonic ectoderm development protein is regulated by microRNAs in human neural tube defects. Am J Obstet Gynecol 2011;204:544.e9-17. [Crossref] [PubMed]
  202. TargetScanHuman release 8.0. SUZ12 ENST00000322652.5. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000322652.5&taxid=9606&showcnc=0&shownc=0&shownc_nc=&showncf1=&showncf2=&subset=1
  203. Pasini D, Cloos PA, Walfridsson J, et al. JARID2 regulates binding of the Polycomb repressive complex 2 to target genes in ES cells. Nature 2010;464:306-10. [Crossref] [PubMed]
  204. Sanulli S, Justin N, Teissandier A, et al. Jarid2 Methylation via the PRC2 Complex Regulates H3K27me3 Deposition during Cell Differentiation. Mol Cell 2015;57:769-83. [Crossref] [PubMed]
  205. TargetScanHuman. Releease 8.0. Human JARID2 ENST00000341776.2. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000341776.2&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  206. Seok HY, Chen J, Kataoka M, et al. Loss of MicroRNA-155 protects the heart from pathological cardiac hypertrophy. Circ Res 2014;114:1585-95. [Crossref] [PubMed]
  207. Nakagawa R, Leyland R, Meyer-Hermann M, et al. MicroRNA-155 controls affinity-based selection by protecting c-MYC+ B cells from apoptosis. J Clin Invest 2016;126:377-88. [Crossref] [PubMed]
  208. Piunti A, Shilatifard A. The roles of Polycomb repressive complexes in mammalian development and cancer. Nat Rev Mol Cell Biol 2021;22:326-45. [Crossref] [PubMed]
  209. Geng Z, Gao Z. Mammalian PRC1 Complexes: Compositional Complexity and Diverse Molecular Mechanisms. Int J Mol Sci 2020;21:8594. [Crossref] [PubMed]
  210. TragetScanHuman. Release 8.0 Human CBX2 ENST00000310942.4. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000310942.4&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  211. Han Q, Li C, Cao Y, et al. CBX2 is a functional target of miRNA let-7a and acts as a tumor promoter in osteosarcoma. Cancer Med 2019;8:3981-91. [Crossref] [PubMed]
  212. Sato T, Kataoka K, Ito Y, et al. Lin28a/let-7 pathway modulates the Hox code via Polycomb regulation during axial patterning in vertebrates. Elife 2020;9:53608. [Crossref] [PubMed]
  213. Nautiyal J. Transcriptional coregulator RIP140: an essential regulator of physiology. J Mol Endocrinol 2017;58:R147-58. [Crossref] [PubMed]
  214. Wei LN, Hu X, Chandra D, et al. Receptor-interacting protein 140 directly recruits histone deacetylases for gene silencing. J Biol Chem 2000;275:40782-7. [Crossref] [PubMed]
  215. Nautiyal J, Christian M, Parker MG. Distinct functions for RIP140 in development, inflammation, and metabolism. Trends Endocrinol Metab 2013;24:451-9. [Crossref] [PubMed]
  216. Leiferman A, Shu J, Upadhyaya B, et al. Storage of Extracellular Vesicles in Human Milk, and MicroRNA Profiles in Human Milk Exosomes and Infant Formulas. J Pediatr Gastroenterol Nutr 2019;69:235-8. [Crossref] [PubMed]
  217. Chiba T, Kooka A, Kowatari K, et al. Expression profiles of hsa-miR-148a-3p and hsa-miR-125b-5p in human breast milk and infant formulae. Int Breastfeed J 2022;17:1. [Crossref] [PubMed]
  218. Schuit F. Epigenetic programming of glucose-regulated insulin release. J Clin Invest 2015;125:2565-8. [Crossref] [PubMed]
  219. Dhawan S, Tschen SI, Zeng C, et al. DNA methylation directs functional maturation of pancreatic β cells. J Clin Invest 2015;125:2851-60. [Crossref] [PubMed]
  220. Ni Q, Gu Y, Xie Y, et al. Raptor regulates functional maturation of murine beta cells. Nat Commun 2017;8:15755. [Crossref] [PubMed]
  221. Melnik BC. Milk exosomal miRNAs: potential drivers of AMPK-to-mTORC1 switching in β-cell de-differentiation of type 2 diabetes mellitus. Nutr Metab (Lond) 2019;16:85. [Crossref] [PubMed]
  222. Sun JF, Zhang D, Gao CJ, et al. Exosome-Mediated MiR-155 Transfer Contributes to Hepatocellular Carcinoma Cell Proliferation by Targeting PTEN. Med Sci Monit Basic Res 2019;25:218-28. [Crossref] [PubMed]
  223. López-Leal R, Díaz-Viraqué F, Catalán RJ, et al. Schwann cell reprogramming into repair cells increases miRNA-21 expression in exosomes promoting axonal growth. J Cell Sci 2020;133:jcs239004. [Crossref] [PubMed]
  224. Jalava SE, Urbanucci A, Latonen L, et al. Androgen-regulated miR-32 targets BTG2 and is overexpressed in castration-resistant prostate cancer. Oncogene 2012;31:4460-71. [Crossref] [PubMed]
  225. TargetScanHuman release 8.0. PRKAA1. Available online: http://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000397128.2&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  226. TargetScanHuman release 8.0. PRKAG2. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000287878.4&taxid=9606&showcnc=0&shownc=0&shownc_nc=&showncf1=&showncf2=&subset=1
  227. Jaafar R, Tran S, Shah AN, et al. mTORC1 to AMPK switching underlies β-cell metabolic plasticity during maturation and diabetes. J Clin Invest 2019;129:4124-37. [Crossref] [PubMed]
  228. El Ouaamari A, Baroukh N, Martens GA, et al. miR-375 targets 3'-phosphoinositide-dependent protein kinase-1 and regulates glucose-induced biological responses in pancreatic beta-cells. Diabetes 2008;57:2708-17. [Crossref] [PubMed]
  229. Eliasson L. The small RNA miR-375 - a pancreatic islet abundant miRNA with multiple roles in endocrine beta cell function. Mol Cell Endocrinol 2017;456:95-101. [Crossref] [PubMed]
  230. Poy MN, Hausser J, Trajkovski M, et al. miR-375 maintains normal pancreatic alpha- and beta-cell mass. Proc Natl Acad Sci U S A 2009;106:5813-8. [Crossref] [PubMed]
  231. Latreille M, Herrmanns K, Renwick N, et al. miR-375 gene dosage in pancreatic β-cells: implications for regulation of β-cell mass and biomarker development. J Mol Med (Berl) 2015;93:1159-69. [Crossref] [PubMed]
  232. Gao H, Luo Z, Jin Z, et al. Adipose Tissue Macrophages Modulate Obesity-Associated β Cell Adaptations through Secreted miRNA-Containing Extracellular Vesicles. Cells 2021;10:2451. [Crossref] [PubMed]
  233. Zhu M, Wei Y, Geißler C, et al. Hyperlipidemia-Induced MicroRNA-155-5p Improves β-Cell Function by Targeting Mafb. Diabetes 2017;66:3072-84. [Crossref] [PubMed]
  234. Cheng P, Chen C, He HB, et al. miR-148a regulates osteoclastogenesis by targeting V-maf musculoaponeurotic fibrosarcoma oncogene homolog B. J Bone Miner Res 2013;28:1180-90. [Crossref] [PubMed]
  235. TargetScanHuman release 8.0. Human MAFB ENST00000373313.2. Available online: http://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000373313.2&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  236. Yeo GS. The role of the FTO (Fat Mass and Obesity Related) locus in regulating body size and composition. Mol Cell Endocrinol 2014;397:34-41. [Crossref] [PubMed]
  237. Deng X, Su R, Stanford S, et al. Critical Enzymatic Functions of FTO in Obesity and Cancer. Front Endocrinol (Lausanne) 2018;9:396. [Crossref] [PubMed]
  238. Qu F, Tsegay PS, Liu Y. N6-Methyladenosine, DNA Repair, and Genome Stability. Front Mol Biosci 2021;8:645823. [Crossref] [PubMed]
  239. Chen A, Chen X, Cheng S, et al. FTO promotes SREBP1c maturation and enhances CIDEC transcription during lipid accumulation in HepG2 cells. Biochim Biophys Acta Mol Cell Biol Lipids 2018;1863:538-48. [Crossref] [PubMed]
  240. Yang Z, Yu GL, Zhu X, et al. Critical roles of FTO-mediated mRNA m6A demethylation in regulating adipogenesis and lipid metabolism: Implications in lipid metabolic disorders. Genes Dis 2022;9:51-61. [Crossref] [PubMed]
  241. Zhang M, Zhang Y, Ma J, et al. The Demethylase Activity of FTO (Fat Mass and Obesity Associated Protein) Is Required for Preadipocyte Differentiation. PLoS One 2015;10:e0133788. [Crossref] [PubMed]
  242. Kajimura S, Seale P, Kubota K, et al. Initiation of myoblast to brown fat switch by a PRDM16-C/EBP-beta transcriptional complex. Nature 2009;460:1154-8. [Crossref] [PubMed]
  243. Zhao X, Yang Y, Sun BF, et al. FTO-dependent demethylation of N6-methyladenosine regulates mRNA splicing and is required for adipogenesis. Cell Res 2014;24:1403-19. [Crossref] [PubMed]
  244. Lee EK, Lee MJ, Abdelmohsen K, et al. miR-130 suppresses adipogenesis by inhibiting peroxisome proliferator-activated receptor gamma expression. Mol Cell Biol 2011;31:626-38. [Crossref] [PubMed]
  245. Gray NE, Lam LN, Yang K, et al. Angiopoietin-like 4 (Angptl4) protein is a physiological mediator of intracellular lipolysis in murine adipocytes. J Biol Chem 2012;287:8444-56. [Crossref] [PubMed]
  246. Wang CY, Shie SS, Wen MS, et al. Loss of FTO in adipose tissue decreases Angptl4 translation and alters triglyceride metabolism. Sci Signal 2015;8:ra127. [Crossref] [PubMed]
  247. de Araújo TM, Velloso LA. Hypothalamic IRX3: A New Player in the Development of Obesity. Trends Endocrinol Metab 2020;31:368-77. [Crossref] [PubMed]
  248. Stratigopoulos G, LeDuc CA, Cremona ML, et al. Cut-like homeobox 1 (CUX1) regulates expression of the fat mass and obesity-associated and retinitis pigmentosa GTPase regulator-interacting protein-1-like (RPGRIP1L) genes and coordinates leptin receptor signaling. J Biol Chem 2011;286:2155-70. [Crossref] [PubMed]
  249. Cheshmeh S, Nachvak SM, Rezvani N, et al. Effects of Breastfeeding and Formula Feeding on the Expression Level of FTO, CPT1A and PPAR-α Genes in Healthy Infants. Diabetes Metab Syndr Obes 2020;13:2227-37. [Crossref] [PubMed]
  250. Tews D, Fischer-Posovszky P, Fromme T, et al. FTO deficiency induces UCP-1 expression and mitochondrial uncoupling in adipocytes. Endocrinology 2013;154:3141-51. [Crossref] [PubMed]
  251. Na RS. Expressional analysis of immune-related miRNAs in breast milk. Genet Mol Res 2015;14:11371-6. [Crossref] [PubMed]
  252. TargetSacnHuman Release 8.0. Human FTO. Available online: http://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000471389.1&taxid=9606&showcn c=0&shownc=0&shownc_nc=&showncf1=&showncf2=&subset=1
  253. Zhang X, Wang Y, Zhao H, et al. Extracellular vesicle-encapsulated miR-22-3p from bone marrow mesenchymal stem cell promotes osteogenic differentiation via FTO inhibition. Stem Cell Res Ther 2020;11:227. [Crossref] [PubMed]
  254. Mussa BM, Taneera J, Mohammed AK, et al. Potential role of hypothalamic microRNAs in regulation of FOS and FTO expression in response to hypoglycemia. J Physiol Sci 2019;69:981-91. [Crossref] [PubMed]
  255. Sun L, Gao M, Qian Q, et al. Triclosan-induced abnormal expression of miR-30b regulates fto-mediated m6A methylation level to cause lipid metabolism disorder in zebrafish. Sci Total Environ 2021;770:145285. [Crossref] [PubMed]
  256. Alvarez-Pitti J, Ros-Forés MA, Bayo-Pérez A, et al. Blood cell transcript levels in 5-year-old children as potential markers of breastfeeding effects in those small for gestational age at birth. J Transl Med 2019;17:145. [Crossref] [PubMed]
  257. Shen F, Huang W, Huang JT, et al. Decreased N(6)-methyladenosine in peripheral blood RNA from diabetic patients is associated with FTO expression rather than ALKBH5. J Clin Endocrinol Metab 2015;100:E148-54. [Crossref] [PubMed]
  258. Yang Y, Shen F, Huang W, et al. Glucose Is Involved in the Dynamic Regulation of m6A in Patients With Type 2 Diabetes. J Clin Endocrinol Metab 2019;104:665-73. [Crossref] [PubMed]
  259. Bornaque F, Delannoy CP, Courty E, et al. Glucose Regulates m6A Methylation of RNA in Pancreatic Islets. Cells 2022;11:291. [Crossref] [PubMed]
  260. De Jesus DF, Zhang Z, Kahraman S, et al. m6A mRNA Methylation Regulates Human β-Cell Biology in Physiological States and in Type 2 Diabetes. Nat Metab 2019;1:765-74. [Crossref] [PubMed]
  261. Wang Y, Sun J, Lin Z, et al. m6A mRNA Methylation Controls Functional Maturation in Neonatal Murine β-Cells. Diabetes 2020;69:1708-22. [Crossref] [PubMed]
  262. He D, Fu M, Miao S, et al. FTO gene variant and risk of hypertension: a meta-analysis of 57,464 hypertensive cases and 41,256 controls. Metabolism 2014;63:633-9. [Crossref] [PubMed]
  263. Chen J, Du B. Novel positioning from obesity to cancer: FTO, an m6A RNA demethylase, regulates tumour progression. J Cancer Res Clin Oncol 2019;145:19-29. [Crossref] [PubMed]
  264. Lan N, Lu Y, Zhang Y, et al. FTO - A Common Genetic Basis for Obesity and Cancer. Front Genet 2020;11:559138. [Crossref] [PubMed]
  265. Zheng QK, Ma C, Ullah I, et al. Roles of N6-Methyladenosine Demethylase FTO in Malignant Tumors Progression. Onco Targets Ther 2021;14:4837-46. [Crossref] [PubMed]
  266. Wang J, Fu Q, Yang J, et al. RNA N6-methyladenosine demethylase FTO promotes osteoporosis through demethylating Runx2 mRNA and inhibiting osteogenic differentiation. Aging (Albany NY) 2021;13:21134-41. [Crossref] [PubMed]
  267. Li HB, Tong J, Zhu S, et al. m6A mRNA methylation controls T cell homeostasis by targeting the IL-7/STAT5/SOCS pathways. Nature 2017;548:338-42. [Crossref] [PubMed]
  268. Shulman Z, Stern-Ginossar N. The RNA modification N6-methyladenosine as a novel regulator of the immune system. Nat Immunol 2020;21:501-12. [Crossref] [PubMed]
  269. Wang Y, Li L, Li J, et al. The Emerging Role of m6A Modification in Regulating the Immune System and Autoimmune Diseases. Front Cell Dev Biol 2021;9:755691. [Crossref] [PubMed]
  270. Melnik BC. The potential mechanistic link between allergy and obesity development and infant formula feeding. Allergy Asthma Clin Immunol 2014;10:37. [Crossref] [PubMed]
  271. Tong J, Cao G, Zhang T, et al. m6A mRNA methylation sustains Treg suppressive functions. Cell Res 2018;28:253-6. [Crossref] [PubMed]
  272. Wood H, Acharjee A, Pearce H, et al. Breastfeeding promotes early neonatal regulatory T-cell expansion and immune tolerance of non-inherited maternal antigens. Allergy 2021;76:2447-60. [Crossref] [PubMed]
  273. Boor KJ, Wiedmann M, Murphy S, et al. A 100-Year Review: Microbiology and safety of milk handling. J Dairy Sci 2017;100:9933-51. [Crossref] [PubMed]
  274. Fusco V, Chieffi D, Fanelli F, et al. Microbial quality and safety of milk and milk products in the 21st century. Compr Rev Food Sci Food Saf 2020;19:2013-49. [Crossref] [PubMed]
  275. Baier SR, Nguyen C, Xie F, et al. MicroRNAs are absorbed in biologically meaningful amounts from nutritionally relevant doses of cow milk and affect gene expression in peripheral blood mononuclear cells, HEK-293 kidney cell cultures, and mouse livers. J Nutr 2014;144:1495-500. [Crossref] [PubMed]
  276. Howard KM, Jati Kusuma R, Baier SR, et al. Loss of miRNAs during processing and storage of cow's (Bos taurus) milk. J Agric Food Chem 2015;63:588-92. [Crossref] [PubMed]
  277. Melnik BC, Schmitz G. Exosomes of pasteurized milk: potential pathogens of Western diseases. J Transl Med 2019;17:3. [Crossref] [PubMed]
  278. Fromm B, Tosar JP, Lu Y, et al. Human and Cow Have Identical miR-21-5p and miR-30a-5p Sequences, Which Are Likely Unsuited to Study Dietary Uptake from Cow Milk. J Nutr 2018;148:1506-7. [Crossref] [PubMed]
  279. Fraser GE, Jaceldo-Siegl K, Orlich M, et al. Dairy, soy, and risk of breast cancer: those confounded milks. Int J Epidemiol 2020;49:1526-37. [Crossref] [PubMed]
  280. Kaluza J, Komatsu S, Lauriola M, et al. Long-term consumption of non-fermented and fermented dairy products and risk of breast cancer by estrogen receptor status - Population-based prospective cohort study. Clin Nutr 2021;40:1966-73. [Crossref] [PubMed]
  281. Harrison S, Lennon R, Holly J, et al. Does milk intake promote prostate cancer initiation or progression via effects on insulin-like growth factors (IGFs)? A systematic review and meta-analysis. Cancer Causes Control 2017;28:497-528. [Crossref] [PubMed]
  282. Sargsyan A, Dubasi HB. Milk Consumption and Prostate Cancer: A Systematic Review. World J Mens Health 2021;39:419-28. [Crossref] [PubMed]
  283. Lee JH, Kim JA, Kwon MH, et al. In situ single step detection of exosome microRNA using molecular beacon. Biomaterials 2015;54:116-25. [Crossref] [PubMed]
  284. Rodríguez-Martínez A, de Miguel-Pérez D, Ortega FG, et al. Exosomal miRNA profile as complementary tool in the diagnostic and prediction of treatment response in localized breast cancer under neoadjuvant chemotherapy. Breast Cancer Res 2019;21:21. [Crossref] [PubMed]
  285. Wang Z, Zong S, Liu Y, et al. Simultaneous detection of multiple exosomal microRNAs for exosome screening based on rolling circle amplification. Nanotechnology 2021;32:085504. [Crossref] [PubMed]
  286. Abdulhussain MM, Hasan NA, Hussain AG. Interrelation of the Circulating and Tissue MicroRNA-21 with Tissue PDCD4 Expression and the Invasiveness of Iraqi Female Breast Tumors. Indian J Clin Biochem 2019;34:26-38. [PubMed]
  287. Foj L, Ferrer F, Serra M, et al. Exosomal and Non-Exosomal Urinary miRNAs in Prostate Cancer Detection and Prognosis. Prostate 2017;77:573-83. [Crossref] [PubMed]
  288. Endzeliņš E, Berger A, Melne V, et al. Detection of circulating miRNAs: comparative analysis of extracellular vesicle-incorporated miRNAs and cell-free miRNAs in whole plasma of prostate cancer patients. BMC Cancer 2017;17:730. [Crossref] [PubMed]
  289. Sánchez CA, Andahur EI, Valenzuela R, et al. Exosomes from bulk and stem cells from human prostate cancer have a differential microRNA content that contributes cooperatively over local and pre-metastatic niche. Oncotarget 2016;7:3993-4008. [Crossref] [PubMed]
  290. Shin S, Park YH, Jung SH, et al. Urinary exosome microRNA signatures as a noninvasive prognostic biomarker for prostate cancer. NPJ Genom Med 2021;6:45. [Crossref] [PubMed]
  291. Davis NM, Sokolosky M, Stadelman K, et al. Deregulation of the EGFR/PI3K/PTEN/Akt/mTORC1 pathway in breast cancer: possibilities for therapeutic intervention. Oncotarget 2014;5:4603-50. [Crossref] [PubMed]
  292. Hare SH, Harvey AJ. mTOR function and therapeutic targeting in breast cancer. Am J Cancer Res 2017;7:383-404. [PubMed]
  293. Butt G, Shahwar D, Qureshi MZ, et al. Role of mTORC1 and mTORC2 in Breast Cancer: Therapeutic Targeting of mTOR and Its Partners to Overcome Metastasis and Drug Resistance. Adv Exp Med Biol 2019;1152:283-92. [Crossref] [PubMed]
  294. Hsieh AC, Liu Y, Edlind MP, et al. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 2012;485:55-61. [Crossref] [PubMed]
  295. Zabala-Letona A, Arruabarrena-Aristorena A, Martín-Martín N, et al. mTORC1-dependent AMD1 regulation sustains polyamine metabolism in prostate cancer. Nature 2017;547:109-13. [Crossref] [PubMed]
  296. Melnik BC, John SM, Weiskirchen R, et al. The endocrine and epigenetic impact of persistent cow milk consumption on prostate carcinogenesis. J Transl Genet Genom 2022;6:1-45. [Crossref]
  297. Aubrey BJ, Strasser A, Kelly GL. Tumor-Suppressor Functions of the TP53 Pathway. Cold Spring Harb Perspect Med 2016;6:a026062. [Crossref] [PubMed]
  298. Blagih J, Buck MD, Vousden KH. p53, cancer and the immune response. J Cell Sci 2020;133:jcs237453. [Crossref] [PubMed]
  299. Shi D, Jiang P. A Different Facet of p53 Function: Regulation of Immunity and Inflammation During Tumor Development. Front Cell Dev Biol 2021;9:762651. [Crossref] [PubMed]
  300. Dybos SA, Flatberg A, Halgunset J, et al. Increased levels of serum miR-148a-3p are associated with prostate cancer. APMIS 2018;126:722-31. [Crossref] [PubMed]
  301. Murata T, Takayama K, Katayama S, et al. miR-148a is an androgen-responsive microRNA that promotes LNCaP prostate cell growth by repressing its target CAND1 expression. Prostate Cancer Prostatic Dis 2010;13:356-61. [Crossref] [PubMed]
  302. Lee E, Wang J, Yumoto K, et al. DNMT1 Regulates Epithelial-Mesenchymal Transition and Cancer Stem Cells, Which Promotes Prostate Cancer Metastasis. Neoplasia 2016;18:553-66. [Crossref] [PubMed]
  303. Saber SH, Ali HEA, Gaballa R, et al. Exosomes are the Driving Force in Preparing the Soil for the Metastatic Seeds: Lessons from the Prostate Cancer. Cells 2020;9:564. [Crossref] [PubMed]
  304. Qin W, Tsukasaki Y, Dasgupta S, et al. Exosomes in Human Breast Milk Promote EMT. Clin Cancer Res 2016;22:4517-24. [Crossref] [PubMed]
  305. Liu F, Kong X, Lv L, et al. TGF-β1 acts through miR-155 to down-regulate TP53INP1 in promoting epithelial-mesenchymal transition and cancer stem cell phenotypes. Cancer Lett 2015;359:288-98. [Crossref] [PubMed]
  306. Santos JC, Lima NDS, Sarian LO, et al. Exosome-mediated breast cancer chemoresistance via miR-155 transfer. Sci Rep 2018;8:829. [Crossref] [PubMed]
  307. Dong X, Bai X, Ni J, et al. Exosomes and breast cancer drug resistance. Cell Death Dis 2020;11:987. [Crossref] [PubMed]
  308. Pieters BC, Arntz OJ, Bennink MB, et al. Commercial cow milk contains physically stable extracellular vesicles expressing immunoregulatory TGF-β. PLoS One 2015;10:e0121123. [Crossref] [PubMed]
  309. Wu M, Tan X, Liu P, et al. Role of exosomal microRNA-125b-5p in conferring the metastatic phenotype among pancreatic cancer cells with different potential of metastasis. Life Sci 2020;255:117857. [Crossref] [PubMed]
  310. Zhou X, Yan T, Huang C, et al. Melanoma cell-secreted exosomal miR-155-5p induce proangiogenic switch of cancer-associated fibroblasts via SOCS1/JAK2/STAT3 signaling pathway. J Exp Clin Cancer Res 2018;37:242. [Crossref] [PubMed]
  311. Melnik BC, John SM, Carrera-Bastos P, et al. MicroRNA-21-Enriched Exosomes as Epigenetic Regulators in Melanomagenesis and Melanoma Progression: The Impact of Western Lifestyle Factors. Cancers (Basel) 2020;12:2111. [Crossref] [PubMed]
  312. Saenz-de-Juano MD, Silvestrelli G, Bauersachs S, et al. Determining extracellular vesicles properties and miRNA cargo variability in bovine milk from healthy cows and cows undergoing subclinical mastitis. BMC Genomics 2022;23:189. [Crossref] [PubMed]
  313. Gao Y, Lin L, Li T, et al. The role of miRNA-223 in cancer: Function, diagnosis and therapy. Gene 2017;616:1-7. [Crossref] [PubMed]
  314. Shi C, Zhang M, Tong M, et al. miR-148a is Associated with Obesity and Modulates Adipocyte Differentiation of Mesenchymal Stem Cells through Wnt Signaling. Sci Rep 2015;5:9930. [Crossref] [PubMed]
  315. Shi C, Pang L, Ji C, et al. Obesity-associated miR148a is regulated by cytokines and adipokines via a transcriptional mechanism. Mol Med Rep 2016;14:5707-12. [Crossref] [PubMed]
  316. Cho YM, Kim TM, Hun Kim D, et al. miR-148a is a downstream effector of X-box-binding protein 1 that silences Wnt10b during adipogenesis of 3T3-L1 cells. Exp Mol Med 2016;48:e226. [Crossref] [PubMed]
  317. Londoño Gentile T, Lu C, Lodato PM, et al. DNMT1 is regulated by ATP-citrate lyase and maintains methylation patterns during adipocyte differentiation. Mol Cell Biol 2013;33:3864-78. [Crossref] [PubMed]
  318. Yang X, Wu R, Shan W, et al. DNA Methylation Biphasically Regulates 3T3-L1 Preadipocyte Differentiation. Mol Endocrinol 2016;30:677-87. [Crossref] [PubMed]
  319. Braud M, Magee DA, Park SD, et al. Genome-Wide microRNA Binding Site Variation between Extinct Wild Aurochs and Modern Cattle Identifies Candidate microRNA-Regulated Domestication Genes. Front Genet 2017;8:3. [Crossref] [PubMed]
  320. Do DN, Dudemaine PL, Li R, et al. Co-Expression Network and Pathway Analyses Reveal Important Modules of miRNAs Regulating Milk Yield and Component Traits. Int J Mol Sci 2017;18:1560. [Crossref] [PubMed]
  321. Schwenk RW, Vogel H, Schürmann A. Genetic and epigenetic control of metabolic health. Mol Metab 2013;2:337-47. [Crossref] [PubMed]
  322. Monda KL, Chen GK, Taylor KC, et al. A meta-analysis identifies new loci associated with body mass index in individuals of African ancestry. Nat Genet 2013;45:690-6. [Crossref] [PubMed]
  323. Voisin S, Almén MS, Zheleznyakova GY, et al. Many obesity-associated SNPs strongly associate with DNA methylation changes at proximal promoters and enhancers. Genome Med 2015;7:103. [Crossref] [PubMed]
  324. Goedeke L, Rotllan N, Canfrán-Duque A, et al. MicroRNA-148a regulates LDL receptor and ABCA1 expression to control circulating lipoprotein levels. Nat Med 2015;21:1280-9. [Crossref] [PubMed]
  325. Rotllan N, Price N, Pati P, et al. microRNAs in lipoprotein metabolism and cardiometabolic disorders. Atherosclerosis 2016;246:352-60. [Crossref] [PubMed]
  326. Goldstein JL, Brown MS. The LDL receptor. Arterioscler Thromb Vasc Biol 2009;29:431-8. [Crossref] [PubMed]
  327. Langmann T, Klucken J, Reil M, et al. Molecular cloning of the human ATP-binding cassette transporter 1 (hABC1): evidence for sterol-dependent regulation in macrophages. Biochem Biophys Res Commun 1999;257:29-33. [Crossref] [PubMed]
  328. Schmitz G, Kaminski WE, Orsó E. ABC transporters in cellular lipid trafficking. Curr Opin Lipidol 2000;11:493-501. [Crossref] [PubMed]
  329. Gao J, Li X, Wang Y, et al. Adipocyte-derived extracellular vesicles modulate appetite and weight through mTOR signalling in the hypothalamus. Acta Physiol (Oxf) 2020;228:e13339. [Crossref] [PubMed]
  330. Zanchi D, Depoorter A, Egloff L, et al. The impact of gut hormones on the neural circuit of appetite and satiety: A systematic review. Neurosci Biobehav Rev 2017;80:457-75. [Crossref] [PubMed]
  331. Clerc P, Coll Constans MG, Lulka H, et al. Involvement of cholecystokinin 2 receptor in food intake regulation: hyperphagia and increased fat deposition in cholecystokinin 2 receptor-deficient mice. Endocrinology 2007;148:1039-49. [Crossref] [PubMed]
  332. Dourish CT, Rycroft W, Iversen SD. Postponement of satiety by blockade of brain cholecystokinin (CCK-B) receptors. Science 1989;245:1509-11. [Crossref] [PubMed]
  333. Chen H, Kent S, Morris MJ. Is the CCK2 receptor essential for normal regulation of body weight and adiposity? Eur J Neurosci 2006;24:1427-33. [Crossref] [PubMed]
  334. Yu B, Lv X, Su L, et al. MiR-148a Functions as a Tumor Suppressor by Targeting CCK-BR via Inactivating STAT3 and Akt in Human Gastric Cancer. PLoS One 2016;11:e0158961. [Crossref] [PubMed]
  335. TargetScanHuman release 8.0: CCKBR. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000334619.2&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  336. Kim YJ, Hwang SH, Cho HH, et al. MicroRNA 21 regulates the proliferation of human adipose tissue-derived mesenchymal stem cells and high-fat diet-induced obesity alters microRNA 21 expression in white adipose tissues. J Cell Physiol 2012;227:183-93. [Crossref] [PubMed]
  337. Mei Y, Bian C, Li J, et al. miR-21 modulates the ERK-MAPK signaling pathway by regulating SPRY2 expression during human mesenchymal stem cell differentiation. J Cell Biochem 2013;114:1374-84. [Crossref] [PubMed]
  338. Kang M, Yan LM, Zhang WY, et al. Role of microRNA-21 in regulating 3T3-L1 adipocyte differentiation and adiponectin expression. Mol Biol Rep 2013;40:5027-34. [Crossref] [PubMed]
  339. Guglielmi V, D'Adamo M, Menghini R, et al. MicroRNA 21 is up-regulated in adipose tissue of obese diabetic subjects. Nutr Healthy Aging 2017;4:141-5. [Crossref] [PubMed]
  340. Zhang XM, Wang LH, Su DJ, et al. MicroRNA-29b promotes the adipogenic differentiation of human adipose tissue-derived stromal cells. Obesity (Silver Spring) 2016;24:1097-105. [Crossref] [PubMed]
  341. Bian Y, Lei Y, Wang C, et al. Epigenetic Regulation of miR-29s Affects the Lactation Activity of Dairy Cow Mammary Epithelial Cells. J Cell Physiol 2015;230:2152-63. [Crossref] [PubMed]
  342. Chen Y, Siegel F, Kipschull S, et al. miR-155 regulates differentiation of brown and beige adipocytes via a bistable circuit. Nat Commun 2013;4:1769. [Crossref] [PubMed]
  343. Engin AB. MicroRNA and Adipogenesis. Adv Exp Med Biol 2017;960:489-509. [Crossref] [PubMed]
  344. Sluijs I, Forouhi NG, Beulens JW, et al. The amount and type of dairy product intake and incident type 2 diabetes: results from the EPIC-InterAct Study. Am J Clin Nutr 2012;96:382-90. [Crossref] [PubMed]
  345. Brouwer-Brolsma EM, Sluik D, Singh-Povel CM, et al. Dairy product consumption is associated with pre-diabetes and newly diagnosed type 2 diabetes in the Lifelines Cohort Study. Br J Nutr 2018;119:442-55. [Crossref] [PubMed]
  346. Chidester S, Livinski AA, Fish AF, et al. The Role of Extracellular Vesicles in β-Cell Function and Viability: A Scoping Review. Front Endocrinol (Lausanne) 2020;11:375. [Crossref] [PubMed]
  347. Guay C, Menoud V, Rome S, et al. Horizontal transfer of exosomal microRNAs transduce apoptotic signals between pancreatic beta-cells. Cell Commun Signal 2015;13:17. [Crossref] [PubMed]
  348. Guay C, Kruit JK, Rome S, et al. Lymphocyte-Derived Exosomal MicroRNAs Promote Pancreatic β Cell Death and May Contribute to Type 1 Diabetes Development. Cell Metab 2019;29:348-361.e6. [Crossref] [PubMed]
  349. Jalabert A, Vial G, Guay C, et al. Exosome-like vesicles released from lipid-induced insulin-resistant muscles modulate gene expression and proliferation of beta recipient cells in mice. Diabetologia 2016;59:1049-58. [Crossref] [PubMed]
  350. Gesmundo I, Pardini B, Gargantini E, et al. Adipocyte-derived extracellular vesicles regulate survival and function of pancreatic β cells. JCI Insight 2021;6:141962. [Crossref] [PubMed]
  351. Melnik BC. Synergistic Effects of Milk-Derived Exosomes and Galactose on α-Synuclein Pathology in Parkinson's Disease and Type 2 Diabetes Mellitus. Int J Mol Sci 2021;22:1059. [Crossref] [PubMed]
  352. Jacovetti C, Matkovich SJ, Rodriguez-Trejo A, et al. Postnatal β-cell maturation is associated with islet-specific microRNA changes induced by nutrient shifts at weaning. Nat Commun 2015;6:8084. [Crossref] [PubMed]
  353. Helman A, Cangelosi AL, Davis JC, et al. A Nutrient-Sensing Transition at Birth Triggers Glucose-Responsive Insulin Secretion. Cell Metab 2020;31:1004-1016.e5. [Crossref] [PubMed]
  354. Herzig S, Shaw RJ. AMPK: guardian of metabolism and mitochondrial homeostasis. Nat Rev Mol Cell Biol 2018;19:121-35. [Crossref] [PubMed]
  355. Entezari M, Hashemi D, Taheriazam A, et al. AMPK signaling in diabetes mellitus, insulin resistance and diabetic complications: A pre-clinical and clinical investigation. Biomed Pharmacother 2022;146:112563. [Crossref] [PubMed]
  356. Zhou G, Myers R, Li Y, et al. Role of AMP-activated protein kinase in mechanism of metformin action. J Clin Invest 2001;108:1167-74. [Crossref] [PubMed]
  357. Melnik BC, Schmitz G. Metformin: an inhibitor of mTORC1 signaling. J Endocrinol Diabetes Obes 2014;2:1029.
  358. Pulito C, Mori F, Sacconi A, et al. Metformin-induced ablation of microRNA 21-5p releases Sestrin-1 and CAB39L antitumoral activities. Cell Discov 2017;3:17022. [Crossref] [PubMed]
  359. Boudeau J, Baas AF, Deak M, et al. MO25alpha/beta interact with STRADalpha/beta enhancing their ability to bind, activate and localize LKB1 in the cytoplasm. EMBO J 2003;22:5102-14. [Crossref] [PubMed]
  360. Lee JH, Cho US, Karin M. Sestrin regulation of TORC1: Is Sestrin a leucine sensor? Sci Signal 2016;9:re5. [Crossref] [PubMed]
  361. Lee JH, Budanov AV, Karin M. Sestrins orchestrate cellular metabolism to attenuate aging. Cell Metab 2013;18:792-801. [Crossref] [PubMed]
  362. Munch EM, Harris RA, Mohammad M, et al. Transcriptome profiling of microRNA by Next-Gen deep sequencing reveals known and novel miRNA species in the lipid fraction of human breast milk. PLoS One 2013;8:e50564. [Crossref] [PubMed]
  363. Li R, Dudemaine PL, Zhao X, et al. Comparative Analysis of the miRNome of Bovine Milk Fat, Whey and Cells. PLoS One 2016;11:e0154129. [Crossref] [PubMed]
  364. TanLChenZTengMMicroRNA profiling reveals an abundant ssc-miRNA-148a-3p that promotes proliferation and differentiation during intramuscular adipogenesis in Chinese Guizhou Congjiang pigs breeds by targeting PPARGC1A.Research Square 2022. Available online: https://assets.researchsquare.com/files/rs-818015/v1/8127f26d-0c88-4ec9-8da4-7ed1ec056a19.pdf?c=1631888228
  365. Gwinn DM, Shackelford DB, Egan DF, et al. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol Cell 2008;30:214-26. [Crossref] [PubMed]
  366. Shaw RJ. LKB1 and AMP-activated protein kinase control of mTOR signalling and growth. Acta Physiol (Oxf) 2009;196:65-80. [Crossref] [PubMed]
  367. Ardestani A, Lupse B, Kido Y, et al. mTORC1 Signaling: A Double-Edged Sword in Diabetic β Cells. Cell Metab 2018;27:314-31. [Crossref] [PubMed]
  368. Artner I, Blanchi B, Raum JC, et al. MafB is required for islet beta cell maturation. Proc Natl Acad Sci U S A 2007;104:3853-8. [Crossref] [PubMed]
  369. Artner I, Hang Y, Mazur M, et al. MafA and MafB regulate genes critical to beta-cells in a unique temporal manner. Diabetes 2010;59:2530-9. [Crossref] [PubMed]
  370. Dai C, Kayton NS, Shostak A, et al. Stress-impaired transcription factor expression and insulin secretion in transplanted human islets. J Clin Invest 2016;126:1857-70. [Crossref] [PubMed]
  371. Guo S, Dai C, Guo M, et al. Inactivation of specific β cell transcription factors in type 2 diabetes. J Clin Invest 2013;123:3305-16. [Crossref] [PubMed]
  372. Conrad E, Dai C, Spaeth J, et al. The MAFB transcription factor impacts islet α-cell function in rodents and represents a unique signature of primate islet β-cells. Am J Physiol Endocrinol Metab 2016;310:E91-E102. [Crossref] [PubMed]
  373. Scoville DW, Cyphert HA, Liao L, et al. MLL3 and MLL4 Methyltransferases Bind to the MAFA and MAFB Transcription Factors to Regulate Islet β-Cell Function. Diabetes 2015;64:3772-83. [Crossref] [PubMed]
  374. Coppieters KT, Wiberg A, Amirian N, et al. Persistent glucose transporter expression on pancreatic beta cells from longstanding type 1 diabetic individuals. Diabetes Metab Res Rev 2011;27:746-54. [Crossref] [PubMed]
  375. Yang Y, Cai Z, Pan Z, et al. Rheb1 promotes glucose-stimulated insulin secretion in human and mouse β-cells by upregulating GLUT expression. Metabolism 2021;123:154863. [Crossref] [PubMed]
  376. Russell R, Carnese PP, Hennings TG, et al. Loss of the transcription factor MAFB limits β-cell derivation from human PSCs. Nat Commun 2020;11:2742. [Crossref] [PubMed]
  377. HumanTargetScan release 8.8. SLC2A1 Human SLC2A1 ENST00000426263.3. Available online: https://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000426263.3&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1)
  378. Ibrahim S, Johnson M, Stephens CH, et al. β-Cell pre-mir-21 induces dysfunction and loss of cellular identity by targeting transforming growth factor beta 2 (Tgfb2) and Smad family member 2 (Smad2) mRNAs. Mol Metab 2021;53:101289. [Crossref] [PubMed]
  379. Brown ML, Schneyer AL. Emerging roles for the TGFbeta family in pancreatic beta-cell homeostasis. Trends Endocrinol Metab 2010;21:441-8. [Crossref] [PubMed]
  380. Lee JH, Lee JH, Rane SG. TGF-β Signaling in Pancreatic Islet β Cell Development and Function. Endocrinology 2021;162:bqaa233. [Crossref] [PubMed]
  381. Backe MB, Novotny GW, Christensen DP, et al. Altering β-cell number through stable alteration of miR-21 and miR-34a expression. Islets 2014;6:e27754. [Crossref] [PubMed]
  382. Bai C, Li X, Gao Y, et al. Role of microRNA-21 in the formation of insulin-producing cells from pancreatic progenitor cells. Biochim Biophys Acta 2016;1859:280-93. [Crossref] [PubMed]
  383. Zhang W, Li Y. miR-148a downregulates the expression of transforming growth factor-β2 and SMAD2 in gastric cancer. Int J Oncol 2016;48:1877-85. [Crossref] [PubMed]
  384. Huang Z, Wen J, Yu J, et al. MicroRNA-148a-3p inhibits progression of hepatocelluar carcimoma by repressing SMAD2 expression in an Ago2 dependent manner. J Exp Clin Cancer Res 2020;39:150. [Crossref] [PubMed]
  385. Jiang F, Li Y, Mu J, et al. Glabridin inhibits cancer stem cell-like properties of human breast cancer cells: An epigenetic regulation of miR-148a/SMAd2 signaling. Mol Carcinog 2016;55:929-40. [Crossref] [PubMed]
  386. Wang J, Yao Y, Wang K, et al. MicroRNA-148a-3p alleviates high glucose-induced diabetic retinopathy by targeting TGFB2 and FGF2. Acta Diabetol 2020;57:1435-43. [Crossref] [PubMed]
  387. Grieco GE, Cataldo D, Ceccarelli E, et al. Serum Levels of miR-148a and miR-21-5p Are Increased in Type 1 Diabetic Patients and Correlated with Markers of Bone Strength and Metabolism. Noncoding RNA 2018;4:37. [Crossref] [PubMed]
  388. Totsuka Y, Tabuchi M, Kojima I, et al. Stimulation of insulin secretion by transforming growth factor-beta. Biochem Biophys Res Commun 1989;158:1060-5. [Crossref] [PubMed]
  389. Nomura M, Zhu HL, Wang L, et al. SMAD2 disruption in mouse pancreatic beta cells leads to islet hyperplasia and impaired insulin secretion due to the attenuation of ATP-sensitive K+ channel activity. Diabetologia 2014;57:157-66. [Crossref] [PubMed]
  390. Bennett K, James C, Hussain K. Pancreatic β-cell KATP channels: Hypoglycaemia and hyperglycaemia. Rev Endocr Metab Disord 2010;11:157-63. [Crossref] [PubMed]
  391. Gerst F, Kemter E, Lorza-Gil E, et al. The hepatokine fetuin-A disrupts functional maturation of pancreatic beta cells. Diabetologia 2021;64:1358-74. [Crossref] [PubMed]
  392. Dhawan S, Dirice E, Kulkarni RN, et al. Inhibition of TGF-β Signaling Promotes Human Pancreatic β-Cell Replication. Diabetes 2016;65:1208-18. [Crossref] [PubMed]
  393. Sjöholm A, Hellerström C. TGF-beta stimulates insulin secretion and blocks mitogenic response of pancreatic beta-cells to glucose. Am J Physiol 1991;260:C1046-51. [Crossref] [PubMed]
  394. Lin HM, Lee JH, Yadav H, et al. Transforming growth factor-beta/Smad3 signaling regulates insulin gene transcription and pancreatic islet beta-cell function. J Biol Chem 2009;284:12246-57. [Crossref] [PubMed]
  395. Gao Y, Zhang R, Dai S, et al. Role of TGF-β/Smad Pathway in the Transcription of Pancreas-Specific Genes During Beta Cell Differentiation. Front Cell Dev Biol 2019;7:351. [Crossref] [PubMed]
  396. Cheng CW, Villani V, Buono R, et al. Fasting-Mimicking Diet Promotes Ngn3-Driven β-Cell Regeneration to Reverse Diabetes. Cell 2017;168:775-788.e12. [Crossref] [PubMed]
  397. Bai C, Gao Y, Li X, et al. MicroRNAs can effectively induce formation of insulin-producing cells from mesenchymal stem cells. J Tissue Eng Regen Med 2017;11:3457-68. [Crossref] [PubMed]
  398. Terzuoli E, Nannelli G, Giachetti A, et al. Targeting endothelial-to-mesenchymal transition: the protective role of hydroxytyrosol sulfate metabolite. Eur J Nutr 2020;59:517-27. [Crossref] [PubMed]
  399. Bronevetsky Y, Burt TD, McCune JM. Lin28b Regulates Fetal Regulatory T Cell Differentiation through Modulation of TGF-β Signaling. J Immunol 2016;197:4344-50. [Crossref] [PubMed]
  400. Shen GY, Ren H, Shang Q, et al. Let-7f-5p regulates TGFBR1 in glucocorticoid-inhibited osteoblast differentiation and ameliorates glucocorticoid-induced bone loss. Int J Biol Sci 2019;15:2182-97. [Crossref] [PubMed]
  401. Kallen AN, Zhou XB, Xu J, et al. The imprinted H19 lncRNA antagonizes let-7 microRNAs. Mol Cell 2013;52:101-12. [Crossref] [PubMed]
  402. Sanchez-Parra C, Jacovetti C, Dumortier O, et al. Contribution of the Long Noncoding RNA H19 to β-Cell Mass Expansion in Neonatal and Adult Rodents. Diabetes 2018;67:2254-67. [Crossref] [PubMed]
  403. Zhang S. Role of microRNA let-7 in pancreatic beta cells. Dissertations, Master's Theses and Master's Reports – Open, Michigan Technical University, 2014. Available online: https://digitalcommons.mtu.edu/cgi/viewcontent.cgi?article=1803&context=etds
  404. Chen PY, Qin L, Barnes C, et al. FGF regulates TGF-β signaling and endothelial-to-mesenchymal transition via control of let-7 miRNA expression. Cell Rep 2012;2:1684-96. [Crossref] [PubMed]
  405. Shalev A, Pise-Masison CA, Radonovich M, et al. Oligonucleotide microarray analysis of intact human pancreatic islets: identification of glucose-responsive genes and a highly regulated TGFbeta signaling pathway. Endocrinology 2002;143:3695-8. [Crossref] [PubMed]
  406. Brunham LR, Kruit JK, Pape TD, et al. Beta-cell ABCA1 influences insulin secretion, glucose homeostasis and response to thiazolidinedione treatment. Nat Med 2007;13:340-7. [Crossref] [PubMed]
  407. Kruit JK, Wijesekara N, Westwell-Roper C, et al. Loss of both ABCA1 and ABCG1 results in increased disturbances in islet sterol homeostasis, inflammation, and impaired β-cell function. Diabetes 2012;61:659-64. [Crossref] [PubMed]
  408. Kruit JK, Wijesekara N, Fox JE, et al. Islet cholesterol accumulation due to loss of ABCA1 leads to impaired exocytosis of insulin granules. Diabetes 2011;60:3186-96. [Crossref] [PubMed]
  409. el-Deiry WS, Tokino T, Velculescu VE, et al. WAF1, a potential mediator of p53 tumor suppression. Cell 1993;75:817-25. [Crossref] [PubMed]
  410. Agarwal S, Bell CM, Taylor SM, et al. p53 Deletion or Hotspot Mutations Enhance mTORC1 Activity by Altering Lysosomal Dynamics of TSC2 and Rheb. Mol Cancer Res 2016;14:66-77. [Crossref] [PubMed]
  411. Jiang Y, Nishimura W, Devor-Henneman D, et al. Postnatal expansion of the pancreatic beta-cell mass is dependent on survivin. Diabetes 2008;57:2718-27. [Crossref] [PubMed]
  412. Wu X, Wang L, Schroer S, et al. Perinatal survivin is essential for the establishment of pancreatic beta cell mass in mice. Diabetologia 2009;52:2130-41. [Crossref] [PubMed]
  413. Liang H, Ward WF. PGC-1alpha: a key regulator of energy metabolism. Adv Physiol Educ 2006;30:145-51. [Crossref] [PubMed]
  414. Besseiche A, Riveline JP, Delavallée L, et al. Oxidative and energetic stresses mediate beta-cell dysfunction induced by PGC-1α. Diabetes Metab 2018;44:45-54. [Crossref] [PubMed]
  415. Nielsen JH, Haase TN, Jaksch C, et al. Impact of fetal and neonatal environment on beta cell function and development of diabetes. Acta Obstet Gynecol Scand 2014;93:1109-22. [Crossref] [PubMed]
  416. Valtat B, Riveline JP, Zhang P, et al. Fetal PGC-1α overexpression programs adult pancreatic β-cell dysfunction. Diabetes 2013;62:1206-16. [Crossref] [PubMed]
  417. Besseiche A, Riveline JP, Gautier JF, et al. Metabolic roles of PGC-1α and its implications for type 2 diabetes. Diabetes Metab 2015;41:347-57. [Crossref] [PubMed]
  418. Ling C, Del Guerra S, Lupi R, et al. Epigenetic regulation of PPARGC1A in human type 2 diabetic islets and effect on insulin secretion. Diabetologia 2008;51:615-22. [Crossref] [PubMed]
  419. Cao L, Kong LP, Yu ZB, et al. microRNA expression profiling of the developing mouse heart. Int J Mol Med 2012;30:1095-104. [Crossref] [PubMed]
  420. HumanTargetScan release 8.0. Human PPARGC1A ENST00000264867. Available online: http://www.targetscan.org/cgi-bin/targetscan/vert_80/view_gene.cgi?rs=ENST00000264867.2&taxid=9606&members=&showcnc=0&shownc=0&showncf1=&showncf2=&subset=1
  421. de Candia P, Spinetti G, Specchia C, et al. A unique plasma microRNA profile defines type 2 diabetes progression. PLoS One 2017;12:e0188980. [Crossref] [PubMed]
  422. Melkman-Zehavi T, Oren R, Kredo-Russo S, et al. miRNAs control insulin content in pancreatic β-cells via downregulation of transcriptional repressors. EMBO J 2011;30:835-45. [Crossref] [PubMed]
  423. Li X. MiR-375, a microRNA related to diabetes. Gene 2014;533:1-4. [Crossref] [PubMed]
  424. García-Jacobo RE, Uresti-Rivera EE, Portales-Pérez DP, et al. Circulating miR-146a, miR-34a and miR-375 in type 2 diabetes patients, pre-diabetic and normal-glycaemic individuals in relation to β-cell function, insulin resistance and metabolic parameters. Clin Exp Pharmacol Physiol 2019;46:1092-100. [Crossref] [PubMed]
  425. Iacomino G, Lauria F, Russo P, et al. The association of circulating miR-191 and miR-375 expression levels with markers of insulin resistance in overweight children: an exploratory analysis of the I.Family Study. Genes Nutr 2021;16:10. [Crossref] [PubMed]
  426. Poy MN, Eliasson L, Krutzfeldt J, et al. A pancreatic islet-specific microRNA regulates insulin secretion. Nature 2004;432:226-30. [Crossref] [PubMed]
  427. Xia HQ, Pan Y, Peng J, et al. Over-expression of miR375 reduces glucose-induced insulin secretion in Nit-1 cells. Mol Biol Rep 2011;38:3061-5. [Crossref] [PubMed]
  428. Wei R, Yang J, Liu GQ, et al. Dynamic expression of microRNAs during the differentiation of human embryonic stem cells into insulin-producing cells. Gene 2013;518:246-55. [Crossref] [PubMed]
  429. Li Y, Xu X, Liang Y, et al. miR-375 enhances palmitate-induced lipoapoptosis in insulin-secreting NIT-1 cells by repressing myotrophin (V1) protein expression. Int J Clin Exp Pathol 2010;3:254-64. [PubMed]
  430. Zhao H, Guan J, Lee HM, et al. Up-regulated pancreatic tissue microRNA-375 associates with human type 2 diabetes through beta-cell deficit and islet amyloid deposition. Pancreas 2010;39:843-6. [Crossref] [PubMed]
  431. Arnold N, Koppula PR, Gul R, et al. Regulation of cardiac expression of the diabetic marker microRNA miR-29. PLoS One 2014;9:e103284. [Crossref] [PubMed]
  432. Chakraborty C, Doss CG, Bandyopadhyay S, et al. Influence of miRNA in insulin signaling pathway and insulin resistance: micro-molecules with a major role in type-2 diabetes. Wiley Interdiscip Rev RNA 2014;5:697-712. [Crossref] [PubMed]
  433. Ślusarz A, Pulakat L. The two faces of miR-29. J Cardiovasc Med (Hagerstown) 2015;16:480-90. [Crossref] [PubMed]
  434. Wu B, Miller D. Involvement of MicroRNAs in Diabetes and Its Complications. Methods Mol Biol 2017;1617:225-39. [Crossref] [PubMed]
  435. Massart J, Sjögren RJO, Lundell LS, et al. Altered miR-29 Expression in Type 2 Diabetes Influences Glucose and Lipid Metabolism in Skeletal Muscle. Diabetes 2017;66:1807-18. [Crossref] [PubMed]
  436. McCormack SE, Shaham O, McCarthy MA, et al. Circulating branched-chain amino acid concentrations are associated with obesity and future insulin resistance in children and adolescents. Pediatr Obes 2013;8:52-61. [Crossref] [PubMed]
  437. Andersson-Hall U, Gustavsson C, Pedersen A, et al. Higher Concentrations of BCAAs and 3-HIB Are Associated with Insulin Resistance in the Transition from Gestational Diabetes to Type 2 Diabetes. J Diabetes Res 2018;2018:4207067. [Crossref] [PubMed]
  438. Yadao DR, MacKenzie S, Bergdahl A. Reducing branched-chain amino acid intake to reverse metabolic complications in obesity and type 2 diabetes. J Physiol 2018;596:3455-6. [Crossref] [PubMed]
  439. Cummings NE, Williams EM, Kasza I, et al. Restoration of metabolic health by decreased consumption of branched-chain amino acids. J Physiol 2018;596:623-45. [Crossref] [PubMed]
  440. Melnik BC, Schmitz G. The impact of persistent milk consumption in the pathogenesis of type 2 diabetes mellitus. Funct Foods Health Dis 2019;9:629-47. [Crossref]
  441. Mersey BD, Jin P, Danner DJ. Human microRNA (miR29b) expression controls the amount of branched chain alpha-ketoacid dehydrogenase complex in a cell. Hum Mol Genet 2005;14:3371-7. [Crossref] [PubMed]
  442. Brosnan JT, Brosnan ME. Branched-chain amino acids: enzyme and substrate regulation. J Nutr 2006;136:207S-11S. [Crossref] [PubMed]
  443. Neinast M, Murashige D, Arany Z. Branched Chain Amino Acids. Annu Rev Physiol 2019;81:139-64. [Crossref] [PubMed]
  444. Ozcan U, Ozcan L, Yilmaz E, et al. Loss of the tuberous sclerosis complex tumor suppressors triggers the unfolded protein response to regulate insulin signaling and apoptosis. Mol Cell 2008;29:541-51. [Crossref] [PubMed]
  445. Bartolomé A, Kimura-Koyanagi M, Asahara S, et al. Pancreatic β-cell failure mediated by mTORC1 hyperactivity and autophagic impairment. Diabetes 2014;63:2996-3008. [Crossref] [PubMed]
  446. Yuan T, Rafizadeh S, Gorrepati KD, et al. Reciprocal regulation of mTOR complexes in pancreatic islets from humans with type 2 diabetes. Diabetologia 2017;60:668-78. [Crossref] [PubMed]
  447. Li J, Zhang Y, Ye Y, et al. Pancreatic β cells control glucose homeostasis via the secretion of exosomal miR-29 family. J Extracell Vesicles 2021;10:e12055. [Crossref] [PubMed]
  448. Sun Y, Zhou Y, Shi Y, et al. Expression of miRNA-29 in Pancreatic β Cells Promotes Inflammation and Diabetes via TRAF3. Cell Rep 2021;34:108576. [Crossref] [PubMed]
  449. Chen H, Zhang SM, Hernán MA, et al. Diet and Parkinson’s disease: A potential role of dairy products in men. Ann Neurol 2002;52:793-801. [Crossref] [PubMed]
  450. Park M, Ross GW, Petrovitch H, et al. Consumption of milk and calcium in midlife and the future risk of Parkinson disease. Neurology 2005;64:1047-51. [Crossref] [PubMed]
  451. Sääksjärvi K, Knekt P, Lundqvist A, et al. A cohort study on diet and the risk of Parkinson’s disease: The role of food groups and diet quality. Br J Nutr 2013;109:329-37. [Crossref] [PubMed]
  452. Kyrozis A, Ghika A, Stathopoulos P, et al. Dietary and lifestyle variables in relation to incidence of Parkinson’s disease in Greece. Eur J Epidemiol 2013;28:67-77. [Crossref] [PubMed]
  453. Jiang W, Ju C, Jiang H, et al. Dairy foods intake and risk of Parkinson’s disease: A dose-response meta-analysis of prospective cohort studies. Eur J Epidemiol 2014;29:613-9. [Crossref] [PubMed]
  454. Hughes KC, Gao X, Kim IY, et al. Intake of dairy foods and risk of Parkinson disease. Neurology 2017;89:46-52. [Crossref] [PubMed]
  455. Olsson E, Byberg L, Höijer J, et al. Milk and fermented milk intake and Parkinson’s disease: Cohort study. Nutrients 2020;12:2763. [Crossref] [PubMed]
  456. Reich SG, Savitt JM. Parkinson’s disease. Med Clin N Am 2019;103:337-50. [Crossref] [PubMed]
  457. Lin KJ, Lin KL, Chen SD, et al. The overcrowded crossroads: Mitochondria, alpha-synuclein, and the endo-lysosomal system interaction in Parkinson’s disease. Int J Mol Sci 2019;20:5312. [Crossref] [PubMed]
  458. Armstrong MJ, Okun MS. Diagnosis and Treatment of Parkinson Disease: A Review. JAMA 2020;323:548-60. [Crossref] [PubMed]
  459. Holmqvist S, Chutna O, Bousset L, et al. Direct evidence of Parkinson pathology spread from the gastrointestinal tract to the brain in rats. Acta Neuropathol 2014;128:805-20. [Crossref] [PubMed]
  460. Klingelhoefer L, Reichmann H. Pathogenesis of Parkinson disease -The gut-brain axis and environmental factors. Nat Rev Neurol 2015;11:625-36. [Crossref] [PubMed]
  461. Liddle RA. Parkinson’s disease from the gut. Brain Res 2018;1693:201-6. [Crossref] [PubMed]
  462. Kim S, Kwon SH, Kam TI, et al. Transneuronal propagation of pathologic α-synuclein from the gut to the brain models Parkinson’s disease. Neuron 2019;103:627-41. [Crossref] [PubMed]
  463. Zhao Q, Liu H, Cheng J, et al. Neuroprotective effects of lithium on a chronic MPTP mouse model of Parkinson's disease via regulation of α synuclein methylation. Mol Med Rep 2019;19:4989-97. [Crossref] [PubMed]
  464. Horvath I, Wittung-Stafshede P. Cross-talk between amyloidogenic proteins in type-2 diabetes and Parkinson’s disease. Proc Natl Acad Sci USA 2016;113:12473-77. [Crossref] [PubMed]
  465. Mucibabic M, Steneberg P, Lidh E, et al. α-Synuclein promotes IAPP fibril formation in vitro and β-cell amyloid formation in vivo in mice. Sci Rep 2020;10:20438. [Crossref] [PubMed]
  466. Su C, Yang X, Lou J. Geniposide reduces α-synuclein by blocking microRNA-21/lysosome-associated membrane protein 2A interaction in Parkinson disease models. Brain Res 2016;1644:98-106. [Crossref] [PubMed]
  467. Alvarez-Erviti L, Seow Y, Schapira AH, et al. Influence of microRNA deregulation on chaperone-mediated autophagy and α-synuclein pathology in Parkinson’s disease. Cell Death Dis 2013;4:e545. [Crossref] [PubMed]
  468. Donda K, Torres BA, Maheshwari A. Non-coding RNAs in Neonatal Necrotizing Enterocolitis. Newborn (Clarksville) 2022;1:120-30. [Crossref] [PubMed]
  469. Mun D, Oh S, Kim Y. Perspectives on Bovine Milk-Derived Extracellular Vesicles for Therapeutic Applications in Gut Health. Food Sci Anim Resour 2022;42:197-209. [Crossref] [PubMed]
doi: 10.21037/exrna-22-5
Cite this article as: Melnik BC, Weiskirchen R, Schmitz G. Milk exosomal microRNAs: friend or foe?—a narrative review. ExRNA 2022;4:22.

Download Citation