Open access peer-reviewed chapter

MicroRNAs and Pancreatic ß Cell Functional Modulation

Written By

Shahzad Irfan, Farhat Jabeen and Haseeb Anwar

Submitted: 22 May 2022 Reviewed: 27 May 2022 Published: 04 July 2022

DOI: 10.5772/intechopen.105588

From the Edited Volume

Recent Advances in Noncoding RNAs

Edited by Lütfi Tutar

Chapter metrics overview

105 Chapter Downloads

View Full Metrics

Abstract

Recent reports of diabetes susceptibility loci located on the non-coding regions of the genome highlight the importance of epigenetic control in health and disease. Specifically, microRNAs have shown to have an important regulatory role in pancreatic ß cell physiology. Human studies implicated that ß cell mass and function are regulated by microRNAs in health and disease. Further, the microRNAs are also implicated in ensuing diabetic complications. Delineating the peculiar role of microRNAs in ß cell physiology and pathophysiology will fill the missing gaps in our current knowledge and help to devise better treatment regimens for diabetes. This chapter will discuss multiple effects of different microRNAs on the ß cell physiology in the context of maintenance and function in Type 2 diabetes mellitus.

Keywords

  • microRNA
  • beta-cell
  • insulin
  • diabetes mellitus

1. Introduction

MicroRNAs (miRNAs/miRs) are pertinent genetic regulators of embryonic development and differentiation, postnatal growth, and metabolism [1]. Initially discovered in C. Elegans, miRNAs are now considered a universal phenomenon for the gene expression regulation from multicellular organisms to mammals [2, 3, 4, 5, 6]. The mammalian genome has been reported to encode more than 500 known microRNA genes. microRNA binds with the mRNA (3′ untranslated region 3’UTR) and result either in the translational repression of that specific gene or in the total degradation of the mRNA [6, 7, 8, 9]. Apart from 3’UTR, miRNA has been reported to interact with 5’UTR and gene promoters and enhancers [10]. miRNAs are critical for normal animal development. As expression pattern of numerous genes important for the embryonic development and subsequent cellular differentiation have been shown to be regulated by microRNAs [11].

The pathological importance of miRNAs is highlighted by the substantial reports of the association of aberrant expression of miRNAs in many human diseases [12, 13]. As miRNAs are also present in the extracellular fluids and act as a signaling molecule for cell communication, their diagnostic importance as a biomarker in different lethal diseases has also been reported [14, 15] Biogenesis of miRNAs and the mechanisms through which they regulate gene expression and their tissue distribution has been previously discussed in detail [1, 10, 16, 17, 18]. The transcriptional activity of miRNA sequences is solely dependent upon RNA polymerase II/III activity [19, 20]. The transcribed miRNA can act through canonical as well as non-canonical pathways [20, 21]. A complementary match either exact or partial between microRNA and the 3’-UTR (conserved) site of the mRNA will result in either the subsequent degradation of the mRNA or moderate reduction in the mRNA levels resulting in translation repression [22, 23]. The degradation of mRNA will result in complete blockage of protein expression. Whereas specific translation repression results in a reduction in the protein formation from the target genes. This fine-tuning of gene expression and subsequent protein formation by microRNAs is important for normal physiological responses. A single miRNA can target hundreds of different genes and mRNAs thus exhibiting multiple levels of complementarity. It is estimated that more than 60% of protein-coding genes are microRNA targets.

In the canonical pathway of microRNA generation (explained in more detail in [24]) events start with the transcription of microRNA sequence by the RNA polymerase II or III, the microRNA transcript is spliced, capped and polyadenylated in the nucleus and is called pri-microRNA. The nuclear pri-microRNA is cleaved by a nuclear protein complex called microprocessor complex and includes Drosha and DGRC8 proteins. After this cleavage event, the pri-microRNA is exported to the cytoplasm by the help of a protein called exportin 5 (Exp 5). Cytoplasmic pre-microRNA is further cleaved by another protein called Dicer and converting the pri-microRNA into a microRNA duplex of around 22 nucleotides. These duplex molecules are loaded onto another protein called Argonaute (AGO). Duplex microRNAs are unwinded and the passenger strand is removed while the guide microRNA strand remains with the AGO proteins and forms the mature effector complex called RNA-induced silencing complex (RISC). RISC complex is guided by the microRNA complementarity to bind with mRNA targets and 3’end de-adenylation occurs resulting in the mRNA degradation of the translation arrest on mRNA. Interestingly deletion of the components of this canonical pathway like DICER or AGO proteins leads to β cells functional impairment in terms of glucose-induced insulin secretion and dedifferentiation into a progenitor-like state [25, 26, 27, 28]. The current chapter will focus on the physiology and pathophysiology of the β cells focusing on the insulin signaling responsible for glucose homeostasis and the influence and interplay of different microRNAs onβ cells functionality in Type 2 diabetes (T2DM).

Advertisement

2. β cells, insulin and macronutrient metabolism

Insulin is the primary endogenous protein responsible for the physiological regulation of metabolism [29]. The regulated secretion of insulin from the pancreas helps to maintain glucose homeostasis in distinct physiological conditions [30, 31]. Endocrine cells that secret specific hormones for the maintenance of glucose levels are present in specialized closed zones in the pancreas called Islets of Langerhans [31, 32, 33]. Islets of Langerhans contain numerous cell types with an appropriate amount of vascular and nervous innervation [31]. Four different types of endocrine cells populated Islets of Langerhans namely: alpha (α) cells, beta (ß) cells, delta (δ) cells, and pancreatic polypeptide (PP) cells. ß cells secrete insulin and initiate postprandial glucose metabolism and control the rising blood glucose levels and α cells secrete of glucagon which enhances blood glucose levels during fasting [34, 35]. Insulin is also responsible for lipid and protein metabolism [36, 37, 38]. The release of insulin from ß cells is initiated by glucose uptake, a phenomenon known as glucose-stimulated insulin secretion (GSIS) [39, 40, 41]. At physiological postprandial levels, glucose initiates insulin gene transcription by recruiting specific transcription factors (PDX-1, MafA, and NeuroD) and post-transcriptionally improves the insulin mRNA stability thus acting as a major physiologic regulator of insulin [42, 43, 44]. Glucose-specific channels/gates are present on the ß cell membrane which are commonly known as glucose transporters (GLUT) [30, 41]. Many types of glucose transporters are present throughout the body [45]. Specifically, GLUT2 is abundant in the pancreas and liver whereas GLUT4 is abundant in skeletal and cardiac muscles as well as adipocytes [45, 46].

Insulin initiates carbohydrate metabolism through phosphorylation of glucose and subsequent formation of glucose-6-phosphate [47, 48]. Insulin phosphorylates glucose through the hexokinase enzyme in muscles and glucokinase (GCK) in ß cells and hepatocytes [49, 50, 51, 52]. Insulin binds to the cell surface insulin receptors commonly present in different tissues [53]. The insulin-insulin receptor binding activates cytoplasmic adaptor proteins called insulin receptor substrates (e.g., IRS1, IRS2) [54]. IRS proteins activate phosphoinositide2-kinase enzyme (PI3K) and convert tyrosine phosphorylation signal into the lipid kinase. Activated PI3K recruits ATP molecules which initiates the AKT activation (serine and threonine kinase) [55]. Activation of AKT by insulin proved landmark and explained the conversion of tyrosine phosphorylation into serine/threonine phosphorylation signal. Insulin-induced AKT activation also helped to describe the insulin-induced regulation of key steps in insulin signaling (a) glucose uptake by glucose transporter, (b) phosphorylation and subsequent inactivation of glycogen synthase kinase 3 (GSK3), (c) activation of the mechanistic target of rapamycin (mTOR) and subsequent synthesis of protein and fats, (d) Transcriptional control of gene expression by forkhead family box O (FOXO) transcription factor proteins. Insulin increases glucose uptake via GLUT4 in muscle and adipose tissue thus enhancing the postprandial rate of glucose transport, insulin also inactivates GSK3 and stimulates glycogen synthesis in these tissues [56]. Insulin prevents hepatic glucose synthesis by inhibiting glycogenolysis and gluconeogenesis. Insulin via AKT inhibits FoxO1 which results in the decreased expression of phosphoenolpyruvate carboxykinase (PEPCK) and glucose −6-phosphatase (G6PC) genes [57, 58, 59]. Insulin regulates lipid and protein metabolism also through AKT. Insulin via AKT induced an increase in mTOR and a decrease in FoxO1 and sterol regulatory element-binding protein (SREBP) inhibits adipocyte lipolysis and helps to lower the plasma fatty acid levels by promoting lipogenesis and enhancing the hepatic formation of very-low-density lipoprotein (VLDL) [60, 61, 62, 63, 64, 65]. Insulin increases the protein synthesis in skeletal muscles and the liver by enhancing the amino acid transport inside the cells, accelerating mRNA translation, and reducing protein degradation and urea formation via AKT activated mTOR [37, 66, 67, 68, 69, 70, 71].

Advertisement

3. Glucose homeostasis: α and β interplay

β cells exhibit a remarkable degree of plasticity in terms of insulin production. The α and ß cells work in conjunction to stabilize glucose levels in feeding and fasting conditions through the periodic release of insulin and glucagon respectively thus achieving glucose homeostasis [30, 49, 72]. Rising plasma glucose levels after feeding require immediate systemic activation of glucose metabolism. Delayed or decreased activation of glucose metabolism results in abnormally high glucose levels known as hyperglycemia. Higher-than-normal concentrations of glucose incite glucotoxic reactions inside the cells. ß cells are triggered by rising postprandial glucose levels to synthesize and secrete insulin. The resulting increasing insulin levels activate enzymes that help to metabolize glucose. Glucokinase and hexokinase initiate glucose phosphorylation primarily in hepatocytes and muscle cells respectively. The postprandial rise in the insulin declines over time because of the decline in the glucose level, a feature of GSIS [39]. Interestingly insulin negatively impacts glucagon secretion [73, 74, 75, 76]. Plasma glucose levels decline under fasting conditions. As glucose is the primary cellular source to generate ATP, a minimum threshold of plasma glucose levels must be maintained to avoid hypoglycemia. Glucagon maintains plasma glucose levels, to avoid hypoglycemia during fasting, by initiating hepatic gluconeogenesis/glucogenolysis [34, 77, 78, 79, 80, 81].

Advertisement

4. Diabetes mellitus

Diabetes mellitus is metabolic dysfunction that results in a significant decline in the cellular ability to metabolize glucose either because of the lack of insulin or insulin inactivity (insulin resistance) or both [82, 83]. Diabetes mellitus is estimated to affect 700 million people worldwide by the year 2040 [84]. The inability of the cells to metabolize glucose cause an abnormal increase in the cellular and plasma levels of glucose (hyperglycemia). Hyperglycemia causes cellular damage via multiple mechanisms including (a) the increased influx of glucose through the polyol pathway, (b) increased formation of advanced glycation end products (AGEs), and (c) subsequent increased expression of AGE receptors and their ligand [85, 86]. Hyperglycemia initiates the reactive oxygen species (ROS) formation via activated glycation reactions and mitochondrial electron transport chain where ROS acts as the main trigger for the initiation of the polyol pathway, formation of AGEs, and subsequent increased AGE receptor expression [87].

Diabetes mellitus has been categorized in two primary forms: Type 1 Diabetes Mellitus (T1DM) and Type 2 Diabetes Mellitus (T2DM). T1DM has been characterized by a mutation in the insulin gene or immune cell-mediated destruction of ß cell resulting in either the synthesis of abnormal insulin protein that fails to activate insulin receptors or a complete lack of endogenous insulin secretion [88]. T1DM patients are usually diagnosed early in their life. The only possible medical treatment referred to these patients is the multiple daily doses of synthetic insulin. T2DM on the other hand is much more complicated and requires a thorough diagnostic approach [82, 89, 90, 91]. T2DM is considered one of the most common metabolic disorders globally. Major risk factors for T2DM include a sedentary lifestyle, lack of exercise, excessive use of a high-carb and high-fat diet, overweight, and obesity [92]. Poor lifestyle and dietary habits have been attributed to the global incidence of Type 2 diabetes in the last 2 decades. Obesity, visceral fat deposition and increased body mass index (BMI) play a central role in the pathophysiology of Type 2 diabetic patients [82].

Advertisement

5. Pathophysiology of Type 2 diabetes mellitus (T2DM)

The development of T2DM is mainly attributed to the progressive decline in insulin secretion over time from ß cells and the progressive inability of insulin-responsive tissues (muscles, fat, and liver) to respond to insulin over time, resulting in hyperinsulinemia and subsequent insulin resistance [93, 94, 95, 96]. The inability of the insulin hormone to activate insulin receptors at the cellular level has been attributed to be the major cause of hyperinsulinemia and insulin resistance [97, 98]. Ex vivoanalysis of pancreatic islets from T2DM donors in terms of GSIS exhibit disproportional insulin levels where increased glucose concentration fails to elicit appropriate ß cells response in terms of insulin [99, 100, 101]. Insulin binding to insulin receptors at the plasma membrane activates a signaling cascade that initiates glucose metabolism inside the cells. Insulin-bound insulin receptors or activated insulin receptors go through internalization at the plasma membrane, a phenomenon known as insulin receptor endocytosis [29, 102]. Following the activation, the endocytosis of the insulin receptor is the primary physiological mechanism through which the duration and intensity of insulin signaling are controlled. Hyperinsulinemia accelerates insulin receptor endocytosis and affects the presence of adequate functional insulin receptors at the plasma membrane resulting in insulin resistance [103]. Apart from accelerated insulin receptor endocytosis, insulin-stimulated insulin receptor kinase activity is also decreased in diabetic patients [104].

Compromised insulin signaling in T2DM fails to activate glucose metabolic enzymes like glucokinase and hexokinase resulting in hyperglycemia. The expression levels of GCK and GLUT2 were found to be lower in human T2D islets as compared to healthy control [105]. As high plasma glucose levels initiate glucose-stimulated insulin secretory (GSIS) response from ß cells resulting in the rise of plasma insulin levels. The rising insulin levels should be normalized over time because of the renal insulin clearance mechanism. But compromised renal insulin clearance rate in diabetic subjects results in abnormally high plasma levels of insulin (hyperinsulinemia) [106, 107]. Hyperinsulinemia and hyperglycemia in theory cannot trigger alpha cells to secrete glucagon. But it has been observed that T2DM patients with insulin resistance, hyperinsulinemia, and hyperglycemia also have abnormally high plasma levels of glucagon [108]. Hinting toward the disturbance in the α and β cell interplay through the inability of the insulin to block glucagon gene transcription [109].

T2D has also been attributed to altering the gene expression of key proteins participating in the processes of insulin secretion and function. Specifically, the insulin receptor gene (IR) and genes involved in Ca2+ influx (SUR1, TMEM37), and mitochondrial metabolism (GPD2, PCK1, FXYD2) [110, 111, 112, 113]. The compromised mitochondrial activity in the T2D β cells through reduced ATP/ADP ratio fails to close the KATP channels and subsequently impacts Ca2+ influx and exocytosis of insulin. ATP also enhances insulin granular priming prior to the exocytosis and thus regulates insulin quantity by initiating the modification of proinsulin into insulin inside the insulin granule. Substantial β cell loss has been considered the hallmark of T2DM which even starts during the prediabetic stage [114]. Up to ~50% loss in the β cell mass has been reported in T2DM patients [115, 116, 117, 118]. Apoptosis and dedifferentiation have been attributed as the main reasons for this substantial decline in the β mass. Hyperinsulinemia has been attributed to the activation of caspases, formation of H2O2 increased expression of nitric oxide synthase (iNOS) in the β cells resulting in their substantial loss [119, 120, 121]. Hyperglycemia-induced production of AGE products and the subsequent activation of AGE receptors further promote the release of cytochrome and caspase activation [122]. Hyperglycemia along with increased levels of free fatty acids (glucolipotoxicity) has also been shown to induce β apoptosis and reduction in their overall mass in T2DM [123, 124]. Over accumulation of lipids and over-activation of lipid signaling pathways inβ cells are of pathological significance as it contributes to β cell loss and further lead to the significant decline in insulin secretion and the onset of T2DM. Glucolipotoxicity induces a cascade of events that starts from mitochondrial dysfunction leading to oxidative stress (ROS) which further leads to ER stress which results in improper protein unfolding response and results in the loss of GSIS and inflammation and activated autophagy. These above-mentioned cascades of events represent altered cell signaling pathways, lipogenic and pro-apoptotic genes, and proteins, increased expression of cytokines, and accumulation of lipids molecules like di and triacylglycerols, cholesterol, and cholesterol esters.

Apart from the above-mentioned mechanisms, β cell dedifferentiation has recently been shown to be another important contributing factor in the progressive decline in the β mass in T2DM [116, 125, 126]. Dedifferentiation is defined as the ability of a cell to revert to its progenitor-like stage (developmental stage). Interestingly it has been observed that dedifferentiation in β cells not only resulted in progenitor-like cells but an increase in other islets cells type has been observed. Implying the possibility that dedifferentiated β cells have further converted into α and δ cells, which interestingly involves the removal of epigenetic brake on glucagon and somatostatin genes for their expression in the functional β cells. This also explains the abnormally high levels of glucagon observed in T2D patients with hyperglycemia, where glucose fails to down-regulate the glucagon gene expression. The transition or dedifferentiation of β cells is characterized by the downregulation of β specific genes and upregulation of pluripotent genes [127, 128, 129, 130, 131]. β cell identity genes like MafA, Nkx6.1, and FoxO1 were shown to be downregulated. Whereas pluripotent genes like Ngn3, Oct4, Nanog, and L-Myc were upregulated. In fact, it has been shown in T2DM patients that glucagon (α) and somatostatin (δ) positive cells have inactivated (cytoplasmic) FOXO1 and NKX6.1 proteins supporting the notion that trans-differentiation of β cells into other endocrine islet cell types [125].

Advertisement

6. Epigenetic regulation of β cell-specific gene expression profile

Epigenetic mechanisms resulting in the alteration of gene expression have been attributed to the regulation or dysregulation of proper β cell function in health and disease [132, 133, 134, 135, 136, 137]. DNA methylation, chromatin modifications, and post-translational modifications of histones have been implicated as the main epigenetic mechanisms. Alteration in the expression pattern of Non-coding RNA sequences like microRNA (miRNA/miR) and long non-coding RNA (lncRNA) has also been shown to regulate the β cell function as well as its cellular identity [138, 139, 140, 141, 142, 143, 144, 145, 146, 147, 148, 149, 150, 151].

Advertisement

7. MicroRNAs and β cell regulation

MicroRNA (miRNA/miR) sequences are typically small around ~22 nucleotide non-protein-coding transcripts. Activation of their transcriptional activity results in the formation of miRNAs. Formation or presence of miRNAs has been shown to regulate the expression of a network of genes involved in β cell development, function, and pathogenesis [152, 153, 154, 155, 156, 157, 158]. microRNAs have been shown to play a vital role in the compensation process of β cells during pregnancy, obesity and T2D [158, 159, 160, 161]. MicroRNAs mediate mRNA silencing and posttranscriptional regulation of gene expression in different physiological states of β cells. These miRNAs induced fine tunned changes in β cell gene expression are considered vital in health and disease. As the inability of microRNAs to fine-tune the gene expression levels lead to β cells decompensation which results in abnormal insulin secretion and ultimately the development of diabetes. β cell decompensation is referred to as two main pathological conditions: (1) impaired (reduced) β cell function in terms of glucose sensitivity and insulin secretion (Reduced GSIS), and (2) reduced β cell number/mass due to dedifferentiation or apoptosis. Impaired β cell function involves molecular defects in insulin biosynthesis, glucose uptake, and exocytosis of insulin granules. β cell maintenance is regulated by microRNAs through the downregulation of non-β cell specific genes and upregulation of the β cells specific genes thus avoiding the β cell dedifferentiation process which is impaired in T2DM patients [162, 163]. As discussed earlier, different molecular components are involved in the secretion of insulin in response to the rising blood glucose levels. Failure of either one of these molecular components results in the failure of β cells to respond to the systemic demands of insulin for macronutrient metabolism. MicroRNAs are involved in the fine-tuned adjustments of β cells and regulate multiple aspects of glucose homeostasis via insulin secretion.

Advertisement

8. MicroRNAs and insulin synthesis and signaling

The insulin gene transcription results in the generation of mRNA which is translated in preproinsulin and cleaved into proinsulin in the endoplasmic reticulum [164, 165, 166, 167]. Cleavage of proinsulin into insulin and C-peptide requires activation of convertase enzymes (PC1/3, PC1) through the change in the granular pH into acidic pH. Insulin gene (ins1) expression has been found to be compromised after the deletion or inactivation of Dicer1 gene [27, 28]. Reduced insulin gene promoter activity and decreased insulin content was observed in isolated islets and cultured β cells after the experimental knock-down of a set of microRNAs including miRNA-24, miRNA-26, miRNA-148 and miRNA-181 [168]. A positive correlation in the insulin mRNA levels and negative correlation in terms of GSIS response and the expression profile of miRNA-127-3p and miRNA-184 was observed in the islets isolated from healthy subjects. Interestingly these correlations were found to be absent in the isolated islets of glucose-intolerant (diabetic) donors [169]. In vitro expansion of β cells from adult pancreatic islets were found to be dedifferentiated after the proliferation and levels of miRNA-375 (an islet-specific microRNA) were found to be significantly reduced [170] Experimental overexpression of miRNA-375 in dedifferentiated β cells lead to the activation of the β cell specific gene expression profile. MafA, a key β cell specific transcription factor regulating insulin gene expression, has been shown to be directly targeted by the miRNA-204 [171]. Thioredoxin-interacting protein (TXNIP), a redox regulator intracellular protein, has been found to be upregulated in diabetes and its deficiency protect β cells from diabetes-associated apoptosis. TXNIP induces the expression of miRNA-204 by inhibiting STAT3 activity. miRNA-204 thus blocks insulin production in diabetes by downregulating MafA. microRNA-181c-5p has been shown to induce the expression of INS1, PDX1, NKX6.1, and MafA in human induced pluripotent stem cells (hiPSCs) [172]. microRNA-124a has been shown to be highly expressed in islets of T2DM patients [173]. microRNA-124a has shown to be a negative regulator of insulin by targeting 3’UTR regions of mRNAs of Akt, Sirt1, NeuroD, and Foxa2 proteins. All these proteins are vital in the insulin gene transcriptional activity as well as the insulin signaling via insulin receptors. microRNA-7 has been shown to influence insulin signaling by down-regulating insulin receptor substrate protein genes (IRS1, IRS2) in gestational diabetes in humans [174]. Glucagon-like peptide 1 receptors (GLP1R) are present in the β cells and activation of these receptors by GLP-1 induces GSIS response from β cells [175, 176]. GLP1 receptor expression has shown to be downregulated by microRNA-204 in rat ins-1 cell line and primary mouse and human islets cells by targeting the 3’UTR region of GLP1R mRNA [177]. This is particularly important as GLP-1 receptor agonists are primary drugs used by T2DM patients to increase insulin levels.

Advertisement

9. MicroRNAs and insulin secretion

Several microRNAs have been shown to impact glucose uptake and metabolism by the β cells and the exocytotic process of insulin secretion. As GSIS regulates the key components of the process of β cell insulin release. Islets isolated from T2DM patients have shown decreased expression of proteins involved in KATP and Ca2+ channel regulation, metabolism, and exocytosis of insulin resulting in hyperglycemia [178]. Islets from T2DM patients show increased levels of microRNA-130a, microRNA-130b, and microRNA-152 and were found to be glucose-induced [179]. microRNA-130a, microRNA-130b, and microRNA-152 target Gck gene and downregulates ATP synthesis by reducing glucose metabolism in β cells thus affecting insulin exocytosis by not only blocking the closure of KATP channels but also affecting the insulin granule priming. microRNA-29a, microRNA-29b and microRNA-124 have been shown to suppress the gene expression of a β cell disallowed mitochondrial protein called monocarboxylate transporter (MCT-1) [180]. MCT-1 inhibition allows the β cells to only consume glucose for metabolism. Thus microRNA-29a, microRNA-29b, and microRNA-124 help in preserving the cell-specific identity. Human and mouse islets incubated with high glucose (hyperglycemic conditions) exhibit an increased expression of microRNA-29 in β cells [181]. microRNA-29 has been shown to positively impact the expression of a transcription factor called Onecut2. This transcription factor upregulates the granuphilin protein gene expression. Granuphilin protein blocks the exocytosis of secretory granules in endocrine cells. Apart from Onecut2-granuphilin mediated blocking of exocytosis, microRNA-29 also influences exocytotic protein genes like Syntaxin1A and SNAP25 (members of SNARE complex: a family of proteins involved in membrane fusion during exocytosis) [182].

Advertisement

10. MicroRNAs and lipid accumulation in β cells

The characteristic hallmark of T2D is the dysregulation of lipid metabolism and subsequent obesity. Excess amounts of fats present in the bloodstream of T2D patients have been attributed to β cell mass reduction and decreased insulin secretion. MicroRNA sequences have been shown to be manipulated by the exposure of β cells to circulating fats in T2D patients [183]. β cell-specific genes which encode proteins that are involved in lipid metabolism and subsequent cholesterol homeostasis have been found to be under miRNA control [184]. miRNA-33 has been reported to control cholesterol homeostasis through modulating the expression of sterol regulatory binding protein (SREBP) genes [185] miRNA-34a has been shown to induce β cell lipotoxicity through multiple mechanisms after the in vitro and in vivo exposure to saturated fatty acids [186, 187, 188, 189]. Increased β cell influx of fatty acids through di/triacylglycerol and esterified cholesterol pathways including targeting sirtuin1 (SIRT1). Importantly SIRT1 (NAD+-dependent deacetylase) regulates the expression of multiple genes encoding proteins like tumor suppressor protein p53 and DNA repair factor ku70 and transcription factor protein including nuclear factor κB (NF-κB) and FOXO family. Exposure of β cells to saturated fatty acids alters the expression profile of important miRNAs which might contribute to β cell lipotoxicity. These miRNAs include miRNA-146 [184], miRNA-182-5p [190], miRNA-297b-5p [191], and miRNA375 [192]. miRNA-146 target genes which are involved with the inflammatory response (Toll-like receptors) [193]. It has been demonstrated that in cultured human islets, the miRNA-146 level increased after the exposure to pro-inflammatory cytokines and decreased exposure to high glucose concentrations but interestingly miRNA-146 levels were unchanged after the palmitate exposure [194]. miRNA-146 along with miRNA 182-5p have been shown to protect against high-fat diet-induced steatohepatitis in a mice model by decreasing in the expression of IL-1 receptor-associated kinase (IRAK1) and tumor necrosis factor (TNF) receptor-associated factor 6 (TRAF6) which results in the reduced cytoplasmic lipid accumulation and inflammation [195]. miRNA-182-5p was found to be increased in INS-1 cells incubated with palmitate and significantly decreased the cell viability and increased palmitate-induced apoptosis. When INS-1 cells were supplemented with miRNA-182-5p inhibitors a significant increase in the cellular viability and palmitate-induced apoptosis was observed [190]. C57BL/6 mouse and β-TC6 cell line exposed to stearic acid and palmitic acid lowers the expression of miRNA-297b-5p. Upregulation of miRNA-297b-5p reduced the stearic acid-induced apoptosis but decreased insulin secretion in β-TC6 cells by inhibiting large-tumor-suppressor kinase 2 (LATS-2) [191] miRNA-375 has been shown to block high-fat diet-induced insulin resistance and obesity in mice by promoting hepatic expression of insulin-responsive genes [196].

11. Conclusion

The important role of epigenetics in the functional regulation of β cells in health and disease has been supported by the increasing number of high levels studies on human and rodent models. Excellent research efforts in the last decades have established that microRNAs help to define the β cell identity and dysregulation of β cell microRNA expression acts as a key event in the pathogenesis of T2DM. As the mechanisms involved in the biogenesis of microRNAs are well understood, it is plausible to assume that microRNAs represent attractive therapeutic targets. Selective inhibition of Dicer molecules by using specific siRNAs or pharmacological tools and selective modulation of a single microRNA. Specific microRNAs involved in diabetes represent a new class of specialized biomarkers for the early detection of diabetes. As microRNAs are biochemically quite stable in the extracellular fluid like blood, they represent a very good biomarker for diagnostics. Further in-depth studies are needed to assess the predictive value of serum levels of microRNAs in relation to the disease progression. Novel therapeutic approaches need to be devised for the prevention and treatment of metabolic disorders.

References

  1. 1. O’Brien J, Hayder H, Zayed Y, Peng C. Overview of microRNA biogenesis, mechanisms of actions, and circulation. Frontiers in Endocrinology. 2018;9(Aug):402
  2. 2. Lee RC, Feinbaum RL, Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75(5):843-854
  3. 3. Friedländer MR, Lizano E, Houben AJS, Bezdan D, Báñez-Coronel M, Kudla G, et al. Evidence for the biogenesis of more than 1,000 novel human microRNAs. Genome Biology. 2014;15(4):R57
  4. 4. Li SC, Chan WC, Hu LY, Lai CH, Hsu CN, Lin W, et al. Identification of homologous microRNAs in 56 animal genomes. Genomics. 2010;96(1):1-9
  5. 5. de Rie D, Abugessaisa I, Alam T, Arner E, Arner P, Ashoor H, et al. An integrated expression atlas of miRNAs and their promoters in human and mouse. Nature Biotechnology. 2017;35(9):872-878
  6. 6. Wightman B, Ha I, Ruvkun G. Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell. 1993;75(5):855-862
  7. 7. Bartel DP. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell. 2004;116(2):281-297
  8. 8. Pasquinelli AE, Hunter S, Bracht J. MicroRNAs: A developing story. Current Opinion in Genetics & Development. 2005;15(2):200-205
  9. 9. Giraldez AJ, Mishima Y, Rihel J, Grocock RJ, van Dongen S, Inoue K, et al. Zebrafish MiR-430 promotes deadenylation and clearance of maternal mRNAs. Science. 2006;312(5770):75-79
  10. 10. Vasudevan S. Posttranscriptional upregulation by MicroRNAs. Wiley Interdisciplinary Reviews: RNA. 2012;3(3):311-330. DOI: 10.1002/wrna.121
  11. 11. Fu G, Brkić J, Hayder H, Peng C. MicroRNAs in human placental development and pregnancy complications. International Journal of Molecular Sciences. 2013;14:5519-5544
  12. 12. Paul P, Chakraborty A, Sarkar D, Langthasa M, Rahman M, Bari M, et al. Interplay between miRNAs and human diseases. Journal of Cellular Physiology. 2018;233(3):2007-2018. DOI: 10.1002/jcp.25854
  13. 13. Tüfekci KU, Öner MG, Meuwissen RLJ, Genç Ş. The role of MicroRNAs in human diseases. Methods in Molecular Biology. 2014;1107:33-50. DOI: 10.1007/978-1-62703-748-8_3
  14. 14. Huang W. MicroRNAs: Biomarkers, diagnostics, and therapeutics. Methods in Molecular Biology. 2017;1617:57-67. DOI: 10.1007/978-1-4939-7046-9_4
  15. 15. Wang J, Chen J, Sen S. MicroRNA as biomarkers and diagnostics. Journal of Cellular Physiology. 2016;231(1):25-30. DOI: 10.1002/jcp.25056
  16. 16. Hartig SM, Hamilton MP, Bader DA, McGuire SE. The miRNA Interactome in metabolic homeostasis. Trends in Endocrinology and Metabolism. 2015;26(12):733-745
  17. 17. Jeon T, il, Osborne TF. miRNA and cholesterol homeostasis. Biochimica et Biophysica Acta—Molecular and cell biology of. Lipids. 2016;1861(12):2041-2046
  18. 18. Brandao BB, Lino M, Kahn CR. Extracellular miRNAs as mediators of obesity-associated disease. Journal of Physiology. 2022;600(5):1155-1169
  19. 19. Borchert GM, Lanier W, Davidson BL. RNA polymerase III transcribes human microRNAs. Nature Structural & Molecular Biology. 2006;13(12):1097-1101
  20. 20. Lee Y, Kim M, Han J, Yeom KH, Lee S, Baek SH, et al. MicroRNA genes are transcribed by RNA polymerase II. The EMBO Journal. 2004;23(20):4051-4060
  21. 21. Abdelfattah AM, Park C, Choi MY. Update on non-canonical microRNAs. Biomolecular Concepts. 2014;5(4):275-287
  22. 22. Grimson A, Farh KKH, Johnston WK, Garrett-Engele P, Lim LP, Bartel DP. MicroRNA targeting specificity in mammals: Determinants beyond seed pairing. Molecular Cell. 2007;27(1):91-105
  23. 23. Moore MJ, Scheel TKH, Luna JM, Park CY, Fak JJ, Nishiuchi E, et al. miRNA-target chimeras reveal miRNA 3′-end pairing as a major determinant of Argonaute target specificity. Nature Communications. 2015;6
  24. 24. Ha M, Kim VN. Regulation of microRNA biogenesis. Nature Reviews Molecular Cell Biology. 2014;15(8):509-524
  25. 25. Mandelbaum AD, Melkman-Zehavi T, Oren R, Kredo-Russo S, Nir T, Dor Y, et al. Dysregulation of Dicer1 in beta cells impairs islet architecture and glucose metabolism. Experimental Diabetes Research. 2012;2012
  26. 26. Tattikota SG, Rathjen T, McAnulty SJ, Wessels HH, Akerman I, van de Bunt M, et al. Argonaute2 mediates compensatory expansion of the pancreatic β cell. Cell Metabolism. 2013;19(1):122-134
  27. 27. Martinez-Sanchez A, Nguyen-Tu MS, Rutter GA. DICER inactivation identifies pancreatic β-cell “disallowed” genes targeted by MicroRNAs. Molecular Endocrinology (Baltimore, Md). 2015;29(7):1067-1079
  28. 28. Kalis M, Bolmeson C, Esguerra JLS, Gupta S, Edlund A, Tormo-Badia N, et al. Beta-cell specific deletion of Dicer1 leads to defective insulin secretion and diabetes mellitus. PLoS One. 2011;6(12):e29166-e29166
  29. 29. Petersen MC, Shulman GI. Mechanisms of insulin action and insulin resistance. Physiological Reviews. 2018;98(4):2133-2223
  30. 30. Thorens B. GLUT2, glucose sensing and glucose homeostasis. Diabetologia. 2015;58:221-232
  31. 31. Atkinson MA, Campbell-Thompson M, Kusmartseva I, Kaestner KH. Organisation of the human pancreas in health and in diabetes. Diabetologia. 2020;63:1966-1973
  32. 32. Tengholm A, Gylfe E. cAMP signalling in insulin and glucagon secretion. Diabetes, Obesity and Metabolism. 2017;19:42-53
  33. 33. Rutter GA, Hodson DJ. Minireview: Intraislet regulation of insulin secretion in humans. Molecular Endocrinology. 2013;27:1984-1995
  34. 34. Yu Q , Shuai H, Ahooghalandari P, Gylfe E, Tengholm A. Glucose controls glucagon secretion by directly modulating cAMP in alpha cells. Diabetologia. 2019;62(7):1212-1224
  35. 35. Kim MK, Shin HM, Jung HS, Lee EJ, Kim TK, Kim TN, et al. Comparison of pancreatic beta cells and alpha cells under hyperglycemia: Inverse coupling in pAkt-FoxO1. Diabetes Research and Clinical Practice. 2017;131:1-11
  36. 36. Lee S, Dong HH. FoxO integration of insulin signaling with glucose and lipid metabolism. Journal of Endocrinology. 2017;233:R67-R79
  37. 37. Tessari P. Changes in protein, carbohydrate, and fat metabolism with aging: Possible role of insulin. Nutrition Reviews. 2000;58(1):11-19
  38. 38. Tessari P. Role of insulin in age-related changes in macronutrient metabolism. European Journal of Clinical Nutrition. 2000;54(Suppl 3):S126-S130
  39. 39. Komatsu M, Takei M, Ishii H, Sato Y. Glucose-stimulated insulin secretion: A newer perspective. Journal of Diabetes Investigation. 2013;4(6):511-516
  40. 40. Straub SG, Sharp GWG. Glucose-stimulated signaling pathways in biphasic insulin secretion. Diabetes/Metabolism Research and Reviews. 2002;18(6):451-463
  41. 41. Wang J, Gu W, Chen C. Knocking down insulin receptor in pancreatic beta cell lines with lentiviral-small hairpin RNA reduces glucose-stimulated insulin secretion via decreasing the gene expression of insulin, GLUT2 and Pdx1. International Journal of Molecular Sciences. 2018;19(4):985
  42. 42. Poitout V, Hagman D, Stein R, Artner I, Robertson RP, Harmon JS. Regulation of the insulin gene by glucose and fatty acids. The Journal of Nutrition. 2006;136(4):873-876
  43. 43. Tillmar L, Carlsson C, Welsh N. Control of insulin mRNA stability in rat pancreatic islets. Regulatory role of a 3′-untranslated region pyrimidine-rich sequence. The Journal of Biological Chemistry. 2002;277(2):1099-1106
  44. 44. Tillmar L, Welsh N. Glucose-induced binding of the polypyrimidine tract-binding protein (PTB) to the 3′-untranslated region of the insulin mRNA (ins-PRS) is inhibited by rapamycin. Molecular and Cellular Biochemistry. 2004;260(1):85-90
  45. 45. Thorens B, Mueckler M. Glucose transporters in the 21st century. American Journal of Physiology. Endocrinology and Metabolism. 2010;298:141-145
  46. 46. Mueckler M, Thorens B. The SLC2 (GLUT) family of membrane transporters. Molecular Aspects of Medicine. 2013;34(2-3):121-138
  47. 47. Posner BI. Insulin signalling: The inside story. Canadian Journal of Diabetes. 2017;41(1):108
  48. 48. Newsholme EA, Dimitriadis G. Integration of biochemical and physiologic effects of insulin on glucose metabolism. Experimental and Clinical Endocrinology and Diabetes. 2001;109(Suppl 2):S122-S134. DOI: 10.1055/s-2001-18575
  49. 49. Matschinsky FM, Wilson DF. The central role of glucokinase in glucose homeostasis: A perspective 50 years after demonstrating the presence of the enzyme in islets of Langerhans. Frontiers in Physiology. 2019;10:148
  50. 50. Massa ML, Gagliardino JJ, Francini F. Liver glucokinase: An overview on the regulatorymechanisms of its activity. IUBMB Life. 2011;63:1-6
  51. 51. Iynedjian PB. Molecular physiology of mammalian glucokinase. Cellular and Molecular Life Sciences. 2009;66:27-42
  52. 52. Matschinsky FM. Regulation of pancreatic-cell glucokinase from basics to therapeutics. Diabetes. 2002;51(Suppl 3):S394-S404
  53. 53. Haeusler RA, TE MG, Accili D. Metabolic Signalling: Biochemical and cellular properties of insulin receptor signalling. Nature Reviews Molecular Cell Biology. 2018;19:31-44
  54. 54. White MF. IRS proteins and the common path to diabetes. American Journal of Physiology—Endocrinology and Metabolism. 2002;283(3):E413-E422
  55. 55. Kohn AD, Kovacina KS, Roth RA. Insulin stimulates the kinase activity of RAC-PK, a pleckstrin homology domain containing ser/thr kinase. The EMBO Journal. 1995;14(17):4288-4295. DOI: 10.1002/j.1460-2075.1995.tb00103.x
  56. 56. Zorzano A, Palacín M, Gumà A. Mechanisms regulating GLUT4 glucose transporter expression and glucose transport in skeletal muscle. Acta Physiologica Scandinavica. 2005;183(1):43-58
  57. 57. Wang L, Liu Q , Kitamoto T, Hou J, Qin J, Accili D. Identification of insulin-responsive transcription factors that regulate glucose production by hepatocytes. Diabetes. 2019;68(6):1156-1167
  58. 58. den Boer MAM, Voshol PJ, Kuipers F, Romijn JA, Havekes LM. Hepatic glucose production is more sensitive to insulin-mediated inhibition than hepatic VLDL-triglyceride production. American Journal of Physiology. Endocrinology and Metabolism. 2006;291:1360-1364
  59. 59. Hatting M, Tavares CDJ, Sharabi K, Rines AK, Puigserver P. Insulin regulation of gluconeogenesis. Annals of the New York Academy of Sciences. 2018;1411:21-35
  60. 60. Kamagate A, Dong HH. FoxO1 integrates insulin signaling to VLDL production NIH public access. Cell Cycle. 2008;7(20):3162-3170
  61. 61. Kamagate A, Qu S, Perdomo G, Su D, Dae HK, Slusher S, et al. FoxO1 mediates insulin-dependent regulation of hepatic VLDL production in mice. Journal of Clinical Investigation. 2008;118(6):2347-2364
  62. 62. Kim DH, Zhang T, Lee S, Calabuig-Navarro V, Yamauchi J, Piccirillo A, et al. FoxO6 integrates insulin signaling with MTP for regulating VLDL production in the liver. Endocrinology. 2014;155(4):1255-1267
  63. 63. Scherer T, OHare J, Diggs-Andrews K, Schweiger M, Cheng B, Lindtner C, et al. Brain insulin controls adipose tissue lipolysis and lipogenesis. Cell Metabolism. 2011;13(2):183-194
  64. 64. Iwen KA, Scherer T, Heni M, Sayk F, Wellnitz T, Machleidt F, et al. Intranasal insulin suppresses systemic but not subcutaneous lipolysis in healthy humans. Journal of Clinical Endocrinology and Metabolism. 2014;99(2)
  65. 65. Saltiel AR. Insulin signaling in the control of glucose and lipid homeostasis. Handbook of Experimental Pharmacology. 2016;233:51-71
  66. 66. Charlton M, Nair KS. Protein metabolism in insulin-dependent diabetes mellitus. The Journal of Nutrition. 1998;128(2):323S-327S
  67. 67. Fukagawa NK, Minaker KL, Rowe JW, Goodman MN, Matthews DE, Bier DM, et al. Insulin-mediated reduction of whole body protein breakdown. Dose-response effects on leucine metabolism in postabsorptive men. Journal of Clinical Investigation. 1985;76(6):2306-2311
  68. 68. Newgard CB, An J, Bain JR, Muehlbauer MJ, Stevens RD, Lien LF, et al. A branched-chain amino acid-related metabolic signature that differentiates obese and lean humans and contributes to insulin resistance. Cell Metabolism. 2009;9(4):311-326
  69. 69. Tessari P, Nosadini R, Trevisan R, de Kreutzenberg SV, Inchiostro S, Duner E, et al. Defective suppression by insulin of leucine-carbon appearance and oxidation in type 1, insulin-dependent diabetes mellitus. Evidence for insulin resistance involving glucose and amino acid metabolism. Journal of Clinical Investigation. 1986;77(6):1797-1804
  70. 70. Yoon MS. The emerging role of branched-chain amino acids in insulin resistance and metabolism. Nutrients. 2016;8(7)
  71. 71. Asghari G, Farhadnejad H, Teymoori F, Mirmiran P, Tohidi M, Azizi F. High dietary intake of branched-chain amino acids is associated with an increased risk of insulin resistance in adults. Journal of Diabetes. 2018;10(5):357-364
  72. 72. Holst JJ, Holland W, Gromada J, Lee Y, Unger RH, Yan H, et al. Insulin and glucagon: Partners for life. Endocrinology. 2017;158(4):696-701
  73. 73. Vergari E, Knudsen JG, Ramracheya R, Salehi A, Zhang Q , Adam J, et al. Insulin inhibits glucagon release by SGLT2-induced stimulation of somatostatin secretion. Nature. Communications. 2019;10(1):139
  74. 74. Wan CKN, Giacca A, Matsuhisa M, El-Bahrani B, Lam L, Rodgers C, et al. Increased responses of glucagon and glucose production to hypoglycemia with intraperitoneal versus subcutaneous insulin treatment. Metabolism: Clinical and Experimental. 2000;49(8):984-989
  75. 75. Ito K, Maruyama H, Hirose H, Kido K, Koyama K, Kataoka K, et al. Exogenous insulin dose-dependently suppresses glucopenia-induced glucagon secretion from perfused rat pancreas. Metabolism. 1995;44(3):358-362
  76. 76. Göbl C, Morettini M, Salvatori B, Alsalim W, Kahleova H, Ahrén B, et al. Temporal patterns of glucagon and its relationships with glucose and insulin following ingestion of different classes of macronutrients. Nutrients. 2011;14(2):376
  77. 77. Walker JN, Ramracheya R, Zhang Q , Johnson PRV, Braun M, Rorsman P. Regulation of glucagon secretion by glucose: Paracrine, intrinsic or both? Diabetes, Obesity and Metabolism. 2011;13(Suppl 1):95-105. DOI: 10.1111/j.1463-1326.2011.01450.x
  78. 78. Onyango AN. Mechanisms of the regulation and dysregulation of glucagon secretion. Oxidative Medicine and Cellular Longevity. 2020;2020:3089139
  79. 79. Barthel A, Schmoll D. Novel Concepts in Insulin Regulation of Hepatic Gluconeogenesis. 2003. Available from: http://www.ajpendo.org
  80. 80. Cooperberg BA, Cryer PE. Insulin reciprocally regulates glucagon secretion in humans. Diabetes. 2010;59(11):2936-2940
  81. 81. Bansal P, Wang Q. Insulin as a physiological modulator of glucagon secretion. American Journal of Physiology—Endocrinology and Metabolism. 2008;295(4)
  82. 82. Galicia-Garcia U, Benito-Vicente A, Jebari S, Larrea-Sebal A, Siddiqi H, Uribe KB, et al. Pathophysiology of type 2 diabetes mellitus. International Journal of Molecular Sciences. MDPI AG. 2020;21:1-34
  83. 83. Harrison LC. The dark side of insulin: A primary autoantigen and instrument of self-destruction in type 1 diabetes. Molecular Metabolism. 2021;52
  84. 84. Maiese K. New insights for oxidative stress and diabetes mellitus. Oxidative Medicine and Cellular Longevity. 2015;2015:875961
  85. 85. Brownlee M. The pathobiology of diabetic complications a unifying mechanism. Diabetes. 2005;54(6):1615-1625
  86. 86. Giacco F, Brownlee M. Oxidative stress and diabetic complications. Circulation Research. 2010;107:1058-1070
  87. 87. Pi J, Zhang Q , Fu J, Woods CG, Hou Y, Corkey BE, et al. ROS signaling, oxidative stress and Nrf2 in pancreatic beta-cell function. Toxicology and Applied Pharmacology. 2010;244:77-83
  88. 88. Harrison LC. The dark side of insulin: A primary autoantigen and instrument of self-destruction in type 1 diabetes. Molecular Metabolism. 2021;52:101288
  89. 89. Hanley SC, Austin E, Assouline-Thomas B, Kapeluto J, Blaichman J, Moosavi M, et al. β-Cell mass dynamics and islet cell plasticity in human type 2 diabetes. Endocrinology. 2010;151(4):1462-1472
  90. 90. Satin LS, Butler PC, Ha J, Sherman AS. Pulsatile insulin secretion, impaired glucose tolerance and type 2 diabetes. Molecular Aspects of Medicine. 2015;42:61-77
  91. 91. Moin ASM, Dhawan S, Cory M, Butler PC, Rizza RA, Butler AE. Increased frequency of hormone negative and polyhormonal endocrine cells in lean individuals with type 2 diabetes. Journal of Clinical Endocrinology and Metabolism. 2016;101(10):3628-3636
  92. 92. Roden M, Shulman GI. The integrative biology of type 2 diabetes. Nature. 2019;576(7785):51-60
  93. 93. Zick Y. Insulin resistance: A phosphorylation-based uncoupling of insulin signaling. Trends in Cell Biology. 2001;11(11):437-441
  94. 94. Chatterjee S, Khunti K, Davies MJ. Type 2 diabetes. The Lancet. 2017;389(10085):2239-2251
  95. 95. Lewis GF, Carpentier A, Vranic M, Giacca A. Resistance to insulin’s acute direct hepatic effect in suppressing steady-state glucose production in individuals with type 2 diabetes. Diabetes. 1999;48(3):570-576
  96. 96. Weyer C, Bogardus C, Mott DM, Pratley RE. The natural history of insulin secretory dysfunction and insulin resistance in the pathophysiology of type diabetes mellitus. Journal of Clinical Investigation. 1999;104(6):787-794
  97. 97. RA Haeusler TMDA. Biochemical and cellular properties of insulin receptor signalling. Nature Reviews. Molecular Cell Biology. 2018;19:31-44
  98. 98. Boucher J, Kleinridders A, Ronald KC. Insulin receptor signaling in normal and insulin-resistant states. Cold Spring Harbor Perspectives in Biology. 2014;6(1)
  99. 99. Solimena M, Schulte AM, Marselli L, Ehehalt F, Richter D, Kleeberg M, et al. Systems biology of the IMIDIA biobank from organ donors and pancreatectomised patients defines a novel transcriptomic signature of islets from individuals with type 2 diabetes. Diabetologia. 2018;61(3):641-657
  100. 100. Lyon J, Manning Fox JE, Spigelman AF, Kim R, Smith N, O’Gorman D, et al. Research-focused isolation of human islets from donors with and without diabetes at the Alberta diabetes institute IsletCore. Endocrinology. 2016;157(2):560-569
  101. 101. Deng S, Vatamaniuk M, Huang X, Doliba N, Lian MM, Frank A, et al. Structural and functional abnormalities in the islets isolated from type 2 diabetic subjects. Diabetes. 2004;53(3):624-632
  102. 102. Kaksonen M, Roux A. Mechanisms of clathrin-mediated endocytosis. Nature Reviews. Molecular Cell Biology. 2018;19(5):313-326
  103. 103. Hall C, Yu H, Choi E. Insulin receptor endocytosis in the pathophysiology of insulin resistance. Experimental & Molecular Medicine. 2020;52(6):911-920
  104. 104. Caro JF, Ittoop O, Pories WJ, Meelheim D, Flickinger EG, Thomas F, et al. Studies on the mechanism of insulin resistance in the liver from humans with noninsulin-dependent diabetes. Insulin action and binding in isolated hepatocytes, insulin receptor structure, and kinase activity. The Journal of Clinical Investigation. 1986;78(1):249-258
  105. 105. Gunton JE, Kulkarni RN, Yim SH, Okada T, Hawthorne WJ, Tseng YH, et al. Loss of ARNT/HIF1beta mediates altered gene expression and pancreatic-islet dysfunction in human type 2 diabetes. Cell. 2005;122(3):337-349
  106. 106. Thomas DD, Corkey BE, Istfan NW, Apovian CM. Hyperinsulinemia: An early indicator of metabolic dysfunction. Journal of the Endocrine Society. 2019;3(9):1727
  107. 107. Kurauti MA, Ferreira SM, Soares GM, Vettorazzi JF, Carneiro EM, Boschero AC, et al. Hyperinsulinemia is associated with increasing insulin secretion but not with decreasing insulin clearance in an age-related metabolic dysfunction mice model. Journal of Cellular Physiology. 2019;234(6):9802-9809
  108. 108. D’alessio D. The role of dysregulated glucagon secretion in type 2 diabetes. Diabetes, Obesity and Metabolism. 2011;13(Suppl 1):126-132
  109. 109. Philippe J, Knepel W, Waeber G. Insulin regulation of the glucagon gene is mediated by an insulin-responsive DNA element. Proceedings of the National Academy of Sciences of the United States of America. 1991;88(16):7224
  110. 110. Ottosson-Laakso E, Krus U, Storm P, Prasad RB, Oskolkov N, Ahlqvist E, et al. Glucose-induced changes in gene expression in human pancreatic islets: Causes or consequences of chronic hyperglycemia. Diabetes. 2017;66(12):3013-3028
  111. 111. Segerstolpe Å, Palasantza A, Eliasson P, Andersson EM, Andréasson AC, Sun X, et al. Single-cell transcriptome profiling of human pancreatic islets in health and type 2 diabetes. Cell Metabolism. 2016;24(4):593-607
  112. 112. Marselli L, Piron A, Suleiman M, Colli ML, Yi X, Khamis A, et al. Persistent or transient human β cell dysfunction induced by metabolic stress: Specific signatures and shared gene expression with type 2 diabetes. Cell Reports. 2020;33(9):108466
  113. 113. Marselli L, Thorne J, Dahiya S, Sgroi DC, Sharma A, Bonner-Weir S, et al. Gene expression profiles of Beta-cell enriched tissue obtained by laser capture microdissection from subjects with type 2 diabetes. PLoS One. 2010;5(7):e11499
  114. 114. Weir GC, Bonner-Weir S. Five of stages of evolving β-cell dysfunction during progression to diabetes. Diabetes. 2004;53(Suppl 3):S16-S21
  115. 115. Khin PP, Lee JH, Jun HS. A brief review of the mechanisms of β-cell dedifferentiation in type 2 diabetes. Nutrients. 2021;13(5):1593
  116. 116. Bensellam M, Jonas JC, Laybutt DR. Mechanisms of β;-cell dedifferentiation in diabetes: Recent findings and future research directions. Journal of Endocrinology. 2018;236(2):R109-R143
  117. 117. Sun T, Han X. Death versus dedifferentiation: The molecular bases of beta cell mass reduction in type 2 diabetes. Seminars in Cell & Developmental Biology. 2020;103:76-82
  118. 118. Marrif HI, Al-Sunousi SI. Pancreatic β cell mass death. Frontiers in Pharmacology. 2016;7:83
  119. 119. Rachdaoui N, Polo-Parada L, Ismail-Beigi F. Prolonged exposure to insulin inactivates Akt and Erk 1/2 and increases pancreatic islet and INS1E β-cell apoptosis. Journal of the Endocrine Society. 2018;3(1):69-90
  120. 120. Bucris E, Beck A, Boura-Halfon S, Isaac R, Vinik Y, Rosenzweig T, et al. Prolonged insulin treatment sensitizes apoptosis pathways in pancreatic β cells. The Journal of Endocrinology. 2016;230(3):291-307
  121. 121. Sampson SR, Bucris E, Horovitz-Fried M, Parnas A, Kahana S, Abitbol G, et al. Insulin increases H2O2-induced pancreatic beta cell death. Apoptosis. 2010;15(10):1165-1176
  122. 122. Zhu Y, Shu T, Lin Y, Wang H, Yang J, Shi Y, et al. Inhibition of the receptor for advanced glycation endproducts (RAGE) protects pancreatic β-cells. Biochemical and Biophysical Research Communications. 2011;404(1):159-165
  123. 123. Imai Y, Cousins RS, Liu S, Phelps BM, Promes JA. Connecting pancreatic islet lipid metabolism with insulin secretion and the development of type 2 diabetes. Annals of the New York Academy of Sciences. 2020;1461(1):53-72
  124. 124. Lytrivi M, Castell AL, Poitout V, Cnop M. Recent insights into mechanisms of β-cell lipo- and glucolipotoxicity in type 2 diabetes. Journal of Molecular Biology. 2020;432(5):1514-1534
  125. 125. Hunter CS, Stein RW. Evidence for loss in identity, de-differentiation, and trans-differentiation of islet β-cells in type 2 diabetes. Frontiers in Genetics. 2017;8:35
  126. 126. Cinti F, Bouchi R, Kim-Muller JY, Ohmura Y, Sandoval PR, Masini M, et al. Evidence of β-cell dedifferentiation in human type 2 diabetes. The Journal of Clinical Endocrinology and Metabolism. 2016;101(3):1044-1054
  127. 127. Brereton MF, Iberl M, Shimomura K, Zhang Q , Adriaenssens AE, Proks P, et al. Reversible changes in pancreatic islet structure and function produced by elevated blood glucose. Nature Communications. 2014;5
  128. 128. Wang Z, York NW, Nichols CG, Remedi MS. Pancreatic β cell dedifferentiation in diabetes and redifferentiation following insulin therapy. Cell Metabolism. 2014;19(5):872-882
  129. 129. Neelankal John A, Ram R, Jiang FX. RNA-Seq analysis of islets to characterise the dedifferentiation in type 2 diabetes model mice db/db. Endocrine Pathology. 2018;29(3):207-221
  130. 130. Talchai C, Xuan S, Lin H, v., Sussel L, Accili D. Pancreatic β cell dedifferentiation as a mechanism of diabetic β cell failure. Cell. 2012;150(6):1223-1234
  131. 131. Spijker HS, Ravelli RBG, Mommaas-Kienhuis AM, van Apeldoorn AA, Engelse MA, Zaldumbide A, et al. Conversion of mature human β-cells into glucagon-producing α-cells. Diabetes. 2013;62(7):2471-2480
  132. 132. Schuit F. Epigenetic programming of glucose-regulated insulin release. The Journal of Clinical Investigation. 2015;125(7):2565-2568
  133. 133. Dhawan S, Tschen SI, Zeng C, Guo T, Hebrok M, Matveyenko A, et al. DNA methylation directs functional maturation of pancreatic β cells. Journal of Clinical Investigation. 2015;125(7):2851-2860
  134. 134. Spaeth JM, Walker EM, Stein R. Impact of Pdx1-associated chromatin modifiers on islet β-cells. Diabetes, Obesity & Metabolism. 2016;18(Suppl 1):123-127
  135. 135. Dayeh T, Ling C. Does epigenetic dysregulation of pancreatic islets contribute to impaired insulin secretion and type 2 diabetes? Biochemistry and Cell Biology. 2015;93(5):511-521
  136. 136. Campbell SA, Hoffman BG. Chromatin regulators in pancreas development and diabetes. Trends in Endocrinology and Metabolism. 2016;27(3):142-152
  137. 137. Astro V, Adamo A. Epigenetic control of endocrine pancreas differentiation in vitro: Current knowledge and future perspectives. Frontiers in Cell and Developmental Biology. 2018;6:141
  138. 138. Akerman I, Tu Z, Beucher A, Rolando DMY, Sauty-Colace C, Benazra M, et al. Human pancreatic β cell lncRNAs control cell-specific regulatory networks. Cell Metabolism. 2017;25(2):400-411
  139. 139. Brodnicki TC. A role for lncRNAs in regulating inflammatory and autoimmune responses underlying type 1 diabetes. Advances in Experimental Medicine and Biology. 2022;1363:97-118
  140. 140. Ruan Y, Lin N, Ma Q , Chen R, Zhang Z, Wen W, et al. Circulating LncRNAs analysis in patients with type 2 diabetes reveals novel genes influencing glucose metabolism and islet β-cell function. Cellular Physiology and Biochemistry. 2018;46(1):335-350
  141. 141. Motterle A, Sanchez-Parra C, Regazzi R. Role of long non-coding RNAs in the determination of β-cell identity. Diabetes, Obesity and Metabolism. 2016;18:41-50
  142. 142. Wilson ME, Pullen TJ. The role of long non-coding RNAs in the regulation of pancreatic beta cell identity. Biochemical Society Transactions. 2021;49(5):2153-2161
  143. 143. López-Noriega L, Rutter GA. Long non-coding RNAs as key modulators of pancreatic β-cell mass and function. Front Endocrinol (Lausanne). 2021;11:610213
  144. 144. Motterle A, Gattesco S, Peyot ML, Esguerra JLS, Gomez-Ruiz A, Laybutt DR, et al. Identification of islet-enriched long non-coding RNAs contributing to β-cell failure in type 2 diabetes. Molecular Metabolism. 2017;6(11):1407-1418
  145. 145. Latreille M, Hausser J, Stützer I, Zhang Q , Hastoy B, Gargani S, et al. MicroRNA-7a regulates pancreatic β cell function. The Journal of Clinical Investigation. 2014;124(6):2722-2735
  146. 146. Dumortier O, Fabris G, Pisani DF, Casamento V, Gautier N, Hinault C, et al. microRNA-375 regulates glucose metabolism-related signaling for insulin secretion. Journal of Endocrinology. 2020;244(1):189-200
  147. 147. Wan S, Zhang J, Chen X, Lang J, Li L, Chen F, et al. MicroRNA-17-92 regulates beta-cell restoration after streptozotocin treatment. Frontiers in Endocrinology. 2020;11:9
  148. 148. Rodríguez-Comas J, Moreno-Asso A, Moreno-Vedia J, Martín M, Castaño C, Marzà-Florensa A, et al. Stress-induced microrna-708 impairs b-cell function and growth. Diabetes. 2017;66(12):3029-3040
  149. 149. Fu X, Dong B, Tian Y, Lefebvre P, Meng Z, Wang X, et al. MicroRNA-26a regulates insulin sensitivity and metabolism of glucose and lipids. Journal of Clinical Investigation. 2015;125(6):2497-2509
  150. 150. Baroukh N, Ravier MA, Loder MK, Hill EV, Bounacer A, Scharfmann R, et al. MicroRNA-124a regulates foxa2 expression and intracellular signaling in pancreatic β-cell lines. Journal of Biological Chemistry. 2007;282(27):19575-19588
  151. 151. Xu Y, Huang Y, Guo Y, Xiong Y, Zhu S, Xu L, et al. MicroRNA-690 regulates induced pluripotent stem cells (iPSCs) differentiation into insulin-producing cells by targeting Sox9. Stem Cell Research and Therapy. 2019;10(1):59
  152. 152. Filios SR, Shalev A. β-Cell MicroRNAs: Small but powerful. Diabetes. 2015;64(11):3631-3644
  153. 153. Fernández-Hernando C, Ramírez CM, Goedeke L, Suárez Y. MicroRNAs in metabolic disease. Arteriosclerosis, Thrombosis, and Vascular Biology. 2013;33(2):178-185
  154. 154. Esguerra JLS, Nagao M, Ofori JK, Wendt A, Eliasson L. MicroRNAs in islet hormone secretion. Diabetes, Obesity and Metabolism. 2018;20:11-19
  155. 155. Kim KH, Hartig SM. Contributions of microRNAs to peripheral insulin sensitivity. Endocrinology. 2022;163(2):bqab250
  156. 156. Özcan S. Minireview: MicroRNA function in pancreatic β cells. Molecular Endocrinology. 2014;28(12):1922-1933
  157. 157. Kaviani M, Azarpira N, Karimi MH, Al-Abdullah I. The role of microRNAs in islet β-cell development. Cell Biology International. 2016;40(12):1248-1255
  158. 158. LaPierre MP, Stoffel M. MicroRNAs as stress regulators in pancreatic beta cells and diabetes. Molecular Metabolism. 2017;6(9):1010-1023
  159. 159. Jacovetti C, Abderrahmani A, Parnaud G, Jonas JC, Peyot ML, Cornu M, et al. MicroRNAs contribute to compensatory β cell expansion during pregnancy and obesity. The Journal of Clinical Investigation. 2012;122(10):3541-3551
  160. 160. Eliasson L, Esguerra JLS. Role of non-coding RNAs in pancreatic beta-cell development and physiology. Acta Physiologica (Oxford, England). 2014;211(2):273-284
  161. 161. Esguerra JLS, Mollet IG, Salunkhe VA, Wendt A, Eliasson L. Regulation of pancreatic Beta cell stimulus-secretion coupling by microRNAs. Genes (Basel). 2014;5(4):1018-1031
  162. 162. Cerf ME. Beta cell dynamics: Beta cell replenishment, beta cell compensation and diabetes. Endocrine. 2013;44(2):303-311
  163. 163. Geach T. Diabetes: Preventing β-cell apoptosis and T2DM with microRNAs—A role for MIR-200? Nature Reviews Endocrinology. 2015;11(8):444
  164. 164. Liu M, Wright J, Guo H, Xiong Y, Arvan P. Proinsulin entry and transit through the endoplasmic reticulum in pancreatic beta cells. Vitamins and Hormones. 2014;95:35-62
  165. 165. Liu M, Weiss MA, Arunagiri A, Yong J, Rege N, Sun J, et al. Biosynthesis, structure, and folding of the insulin precursor protein. Diabetes, Obesity and Metabolism. 2018;20:28-50
  166. 166. Steiner DF. The proinsulin C-peptide—A multirole model. Experimental Diabesity Research. 2004;5(1):7-14
  167. 167. Steiner DF. On the discovery of precursor processing. Methods in Molecular Biology. 2011;768:3-11
  168. 168. Melkman-Zehavi T, Oren R, Kredo-Russo S, Shapira T, Mandelbaum AD, Rivkin N, et al. miRNAs control insulin content in pancreatic β-cells via downregulation of transcriptional repressors. The EMBO Journal. 2011;30(5):835-845
  169. 169. Bolmeson C, Esguerra JLS, Salehi A, Speidel D, Eliasson L, Cilio CM. Differences in islet-enriched miRNAs in healthy and glucose intolerant human subjects. Biochemical and Biophysical Research Communications. 2010;404(1):16-22
  170. 170. Nathan G, Kredo-Russo S, Geiger T, Lenz A, Kaspi H, Hornstein E, et al. MiR-375 promotes redifferentiation of adult human β cells expanded in vitro. PLoS One. 2015;10(4):e0122108-e0122108
  171. 171. Xu G, Chen J, Jing G, Shalev A. Thioredoxin-interacting protein regulates insulin transcription through microRNA-204. Nature Medicine. 2013;19(9):1141-1146
  172. 172. Li N, Jiang D, He Q , He F, Li Y, Deng C, et al. microRNA-181c-5p promotes the formation of insulin-producing cells from human induced pluripotent stem cells by targeting smad7 and TGIF2. Cell Death & Disease. 2020;11(6):1-12
  173. 173. MicroRNA miR-124a, a negative regulator of insulin secretion, is hyperexpressed in human pancreatic islets of type 2 diabetic patients. RNA & DISEASE. 2015
  174. 174. Bhushan R, Rani A, Gupta D, Akhtar A, Dubey PK. MicroRNA-7 regulates insulin signaling pathway by targeting IRS1, IRS2, and RAF1 genes in gestational diabetes mellitus. MicroRNA (Shariqah, United Arab Emirates). 2022;11
  175. 175. Drucker DJ. Mechanisms of action and therapeutic application of glucagon-like peptide-1. Cell Metabolism. 2018;27(4):740-756
  176. 176. Drucker DJ, Nauck MA. The incretin system: Glucagon-like peptide-1 receptor agonists and dipeptidyl peptidase-4 inhibitors in type 2 diabetes. Lancet. 2006;368(9548):1696-1705
  177. 177. Jo SH, Chen J, Xu G, Grayson TB, Thielen LA, Shalev A. miR-204 controls glucagon-like peptide 1 receptor expression and agonist function. Diabetes. 2018;67(2):256-264
  178. 178. Rorsman P, Ashcroft FM. Pancreatic β-cell electrical activity and insulin secretion: Of mice and men. Physiological Reviews. 2018;98(1):117-214
  179. 179. Ofori JK, Salunkhe VA, Bagge A, Vishnu N, Nagao M, Mulder H, et al. Elevated miR-130a/miR130b/miR-152 expression reduces intracellular ATP levels in the pancreatic beta cell. Scientific Reports. 2017;7(1):1-15
  180. 180. Pullen TJ, da Silva XG, Kelsey G, Rutter GA. miR-29a and miR-29b contribute to pancreatic β-cell-specific silencing of monocarboxylate transporter 1 (Mct1). Molecular and Cellular Biology. 2011;31(15):3182-3194
  181. 181. Bagge A, Clausen TR, Larsen S, Ladefoged M, Rosenstierne MW, Larsen L, et al. MicroRNA-29a is up-regulated in beta-cells by glucose and decreases glucose-stimulated insulin secretion. Biochemical and Biophysical Research Communications. 2012;426(2):266-272
  182. 182. Bagge A, Dahmcke CM, Dalgaard LT. Syntaxin-1a is a direct target of miR-29a in insulin-producing β-cells. Hormone and Metabolic Research. 2013;45(6):463-466
  183. 183. Tarlton JMR, Patterson S, Graham A. MicroRNA sequences modulated by beta cell lipid metabolism: Implications for type 2 diabetes mellitus. Biology (Basel). 2021;10(6):534
  184. 184. Lovis P, Roggli E, Laybutt DR, Gattesco S, Yang JY, Widmann C, et al. Alterations in microRNA expression contribute to fatty acid-induced pancreatic beta-cell dysfunction. Diabetes. 2008;57(10):2728-2736
  185. 185. Najafi-Shoushtari SH, Kristo F, Li Y, Shioda T, Cohen DE, Gerszten RE, et al. MicroRNA-33 and the SREBP host genes cooperate to control cholesterol homeostasis. Science. 2010;328(5985):1566-1569
  186. 186. Han YB, Wang MN, Li Q , Guo L, Yang YM, Li PJ, et al. MicroRNA-34a contributes to the protective effects of glucagon-like peptide-1 against lipotoxicity in INS-1 cells. Chinese Medical Journal. 2012;125(23):4202-4208
  187. 187. Lu H, Hao L, Li S, Lin S, Lv L, Chen Y, et al. Elevated circulating stearic acid leads to a major lipotoxic effect on mouse pancreatic beta cells in hyperlipidaemia via a miR-34a-5p-mediated PERK/p53-dependent pathway. Diabetologia. 2016;59(6):1247-1257
  188. 188. Kong X, Liu CX, Wang GD, Yang H, Yao XM, Hua Q , et al. LncRNA LEGLTBC functions as a ceRNA to antagonize the effects of miR-34a on the downregulation of SIRT1 in glucolipotoxicity-induced INS-1 Beta cell oxidative stress and apoptosis. Oxidative Medicine and Cellular Longevity. 2019;2019:4010764
  189. 189. Lin X, Guan H, Huang Z, Liu J, Li H, Wei G, et al. Downregulation of Bcl-2 expression by miR-34a mediates palmitate-induced Min6 cells apoptosis. Journal Diabetes Research. 2014;2014:258695
  190. 190. Liu Y, Dong J, Ren BO. MicroRNA-182-5p contributes to the protective effects of thrombospondin 1 against lipotoxicity in INS-1 cells. Experimental and Therapeutic Medicine. 2018;16(6):5272-5279
  191. 191. Guo R, Yu Y, Zhang Y, Li Y, Chu X, Lu H, et al. Overexpression of miR-297b-5p protects against stearic acid-induced pancreatic β-cell apoptosis by targeting LATS2. American Journal of Physiology. Endocrinology and Metabolism. 2020;318(3):E430-E439
  192. 192. Li Y, Xu X, Liang Y, Liu S, Xiao H, Li F, et al. miR-375 enhances palmitate-induced lipoapoptosis in insulin-secreting NIT-1 cells by repressing myotrophin (V1) protein expression. International Journal of Clinical and Experimental Pathology. 2010;3(3):254
  193. 193. Paterson MR, Kriegel AJ. MiR-146a/b: A family with shared seeds and different roots. Physiological Genomics. 2017;49(4):243-252
  194. 194. Fred RG, Bang-Berthelsen CH, Mandrup-Poulsen T, Grunnet LG, Welsh N. High glucose suppresses human islet insulin biosynthesis by inducing miR-133a leading to decreased polypyrimidine tract binding protein-expression. PLoS One. 2010;5(5)
  195. 195. Jiang W, Liu J, Dai Y, Zhou N, Ji C, Li X. MiR-146b attenuates high-fat diet-induced non-alcoholic steatohepatitis in mice. Journal of Gastroenterology and Hepatology. 2015;30(5):933-943
  196. 196. Kumar A, Ren Y, Sundaram K, Mu J, Sriwastva MK, Dryden GW, et al. miR-375 prevents high-fat diet-induced insulin resistance and obesity by targeting the aryl hydrocarbon receptor and bacterial tryptophanase (tnaA) gene. Theranostics. 2021;11(9):4061-4077

Written By

Shahzad Irfan, Farhat Jabeen and Haseeb Anwar

Submitted: 22 May 2022 Reviewed: 27 May 2022 Published: 04 July 2022