Open access peer-reviewed chapter

Screening for Gestational Diabetes Mellitus: The Potential of MicroRNAs

Written By

Carmen Pheiffer, Stephanie Dias, Paul Rheeder and Sumaiya Adam

Submitted: 17 July 2018 Reviewed: 18 October 2018 Published: 19 November 2018

DOI: 10.5772/intechopen.82102

From the Edited Volume

Gestational Diabetes Mellitus - An Overview with Some Recent Advances

Edited by Amita Ray

Chapter metrics overview

1,298 Chapter Downloads

View Full Metrics

Abstract

Gestational diabetes mellitus (GDM) is associated with short- and long-term complications in both mothers and their offspring. Screening and early diagnosis of GDM is advocated as a strategy to prevent adverse pregnancy outcomes. However, there is currently no test that is amenable to routine screening, particularly in low-and middle-income countries. MicroRNAs (miRNAs) are small non-coding RNA molecules that regulate gene expression post-transcriptionally. In recent years, miRNAs have been the focus of increasing research due to their important role in regulating biological pathways and their aberrant expression during disease. The discovery of circulating miRNAs in maternal blood, and their altered expression during pregnancy-associated complications have increased interest into their potential as diagnostic biomarkers for GDM. In this review, we summarise studies that have investigated miRNAs in maternal blood thus providing an update of the current status of miRNAs as biomarkers for GDM. We also discuss the challenges of miRNA profiling, and highlight perspectives and recommendations for research.

Keywords

  • gestational diabetes mellitus
  • biomarkers
  • epigenetics
  • microRNAs
  • pregnancy

1. Introduction

Gestational diabetes mellitus (GDM) is defined as glucose intolerance that is first diagnosed during pregnancy with glucose homeostasis usually restored shortly after birth. The rate of GDM has constantly increased over the last 20 years [1], and in 2017 the International Diabetes Federation estimated that about 14% of women with live births had GDM [2]. Without appropriate glucose management, GDM is associated with short- and long-term complications in both mothers and their offspring [3, 4, 5, 6]. Treatment of GDM is effective in preventing these adverse outcomes [7, 8, 9, 10, 11], thus, universal screening and early detection of GDM is widely advocated as a strategy to promote timely treatment and improve pregnancy outcomes [3]. The oral glucose tolerance test (OGTT) conducted between the 24th and 28th week of pregnancy, is currently the gold standard for GDM diagnosis [12]. However, the test is time-consuming, expensive and unfeasible in most countries. The identification of simple and cost-effective biomarkers that do not require fasting and multiple blood draws would be more acceptable to pregnant women, and thereby facilitate screening for GDM. MicroRNAs (miRNAs) are small noncoding RNA molecules that regulate various metabolic pathways. They are implicated in the pathophysiology of various diseases and have attracted considerable interest as biomarkers of metabolic disease. Recently, several studies have explored their potential as biomarkers of GDM. The purpose of this review is to provide an update of the status of miRNAs as biomarkers for GDM. All studies that have profiled miRNAs in maternal blood during GDM to date are summarised. We also discuss the challenges of miRNA research, and highlight perspectives and recommendations for future research.

Advertisement

2. Overview of gestational diabetes

Hyperglycaemia during pregnancy creates an adverse intrauterine environment that predisposes both mother and offspring to perinatal complications and future metabolic disease [3, 4, 5, 6]. Maternal perinatal complications include caesarean section, preeclampsia and birth injuries. Women with pregnancies complicated by GDM also have an increased risk of developing disease in later life. In 2009, Bellamy et al. conducted a comprehensive review of the literature and found that women who have had GDM are at least seven-fold more likely to develop Type 2 diabetes (T2D) compared to women with normoglycaemic pregnancies [4]. Other studies showed that GDM is associated with the development of metabolic disease [13], cardiovascular disease [14] and breast cancer [15].

Foetal and neonatal complications associated with GDM include macrosomia, congenital malformations, perinatal death, hypertrophic cardiomyopathy, intrauterine growth restriction, preterm birth, respiratory distress syndrome, hypoglycaemia, hypocalcaemia, polycythaemia and hyperbilirubinemia [5]. In recent years, increasing evidence support the critical role of the intrauterine environment in programming the foetus and influencing long-term offspring health [16]. In the 1980s, David Barker and his colleagues proposed Barker’s hypothesis or the developmental origins of adult disease, which suggests that metabolic diseases have their origins in early development [17]. Subsequently, several other studies have reported that diabetes during pregnancy is associated with the development of obesity and diabetes in children [5].

The prevalence of GDM is rapidly increasing, spurred by the global obesity pandemic. Pregnant women who are overweight, obese or severely obese have a 2.14-, 3.56- and 8.56-fold risk of developing GDM compared to normal weight women [18]. The short- and long-term consequences of GDM are likely to have a major negative impact, particularly on low- and middle-income countries that already have limited financial and human resources, and are least able to respond to the challenge. Screening and treatment of GDM leads to improved pregnancy outcomes [7, 8, 9, 10, 11], thus universal screening for GDM is widely advocated as a strategy to prevent pregnancy complications. However, the OGTT, which is considered the gold standard for GDM diagnosis is not amenable to routine screening [3]. Currently, traditional risk-factor screening based on obesity, age older than ≥35 years, non-white ethnicity, and having a family history of diabetes [3] is mostly employed. Unfortunately, these risk factors have poor predictive value [19, 20]. A number of other laboratory tests such as glycated haemoglobin (HbA1c), insulin, adiponectin, glycosylated fibronectin and C-reactive protein have been explored, however, they too have several challenges and are not yet clinically applicable [3].

Advertisement

3. Characteristics of ideal biomarkers

Biomarkers are defined as “cellular, biochemical or molecular alterations that are measurable in biological media such as human tissues, cells or fluids” [21]. Screening, diagnostic and prognostic biomarkers offer several advantages and thus efforts to identify biomarkers of disease have intensified. They are clinically useful and can be used to detect or monitor disease progression, thus facilitating earlier diagnosis and disease management. Furthermore, biomarkers are able to monitor pharmacological responses and predict clinical outcome. As recently reviewed by Etheridge et al. [22], characteristics of the ideal biomarker include sensitivity, specificity, cost-effectiveness, reproducibility, robustness, accessibility, stability and ability to differentiate between pathologies. Recent advancements in molecular biology have led to the development of molecular biomarkers that are sensitive and specific, and are easily measured in biological fluids such as whole blood, plasma and serum [22].

Advertisement

4. MicroRNAs

MiRNAs are epigenetic mechanisms that reflect gene-environment interactions and are increasingly being implicated in the pathophysiology of metabolic diseases [23]. Since their discovery in Caenorhabditis elegans in 1993 [24], miRNAs have emerged as one of the most powerful epigenetic mechanisms regulating diverse biological processes including development, proliferation, differentiation and apoptosis [25]. They are short, single-stranded, highly conserved, non-coding RNA molecules of approximately 22 nucleotides in length that regulate gene expression through post-transcriptional mechanisms. MiRNAs bind to the 3′ untranslated region (UTR) of messenger RNA (mRNA) inducing degradation or translational repression of the mRNA transcript [26]. Using an elegant set of experiments, Guo et al. showed that destabilisation of target mRNAs rather than translational repression is the main mechanism whereby miRNAs reduce protein expression [26]. A single miRNA is able to regulate up to 200 target genes, implying that about 30% of the genome is regulated by miRNAs [27, 28] and confirming the important role of miRNAs as mediators of biological function. More than 2000 miRNAs are present in the human genome, and function in various biological processes [27, 28, 29].

Mature miRNAs are produced through a stepwise process. Briefly, primary miRNA transcripts (pri-miRNAs) are transcribed in the nucleus by RNA polymerase II (and possibly by RNA polymerase III), which are then cleaved by Drosha RNase III endonuclease to produce stem-loop precursor miRNAs (pre-miRNAs) that are approximately 70 nucleotides long. Ran-GTP and the export receptor, Exportin-5 transports pre-miRNAs to the cytoplasm, where Dicer, also a RNase III endonuclease, cleaves them to produce mature miRNAs. Mature miRNAs complex with the RNA-induced silencing complex (RISC) and bind to the 3’ UTR of mRNA to induce predominantly mRNA degradation [25, 26].

MiRNAs regulate a wide range of biological processes including cell proliferation and differentiation, apoptosis and metabolism, thus it is not surprising that altered miRNA expression have been shown to associated with various conditions including cancer, obesity, T2D and cardiovascular disease [30]. MiRNAs play a critical role in the pathophysiology of metabolic disease, and their aberrant expression is observed in tissues associated with disease. For example, various in vitro, in vivo animal models and studies in diabetic patients have demonstrated the altered expression of miRNAs that regulate insulin secretion, adipocyte differentiation, lipid metabolism, inflammation and glucose homeostasis in dysfunctional pancreatic beta cells and insulin-resistant target tissues, such as adipose, liver and muscle during T2D [23]. Increasingly evidence show that correcting aberrant miRNA expression can prevent or treat T2D, making them attractive therapeutic targets [31].

The identification of circulating miRNAs in biological fluids such as whole blood, serum, plasma and urine has sparked research efforts to investigate their feasibility as diagnostic or prognostic biomarkers of disease [32, 33, 34]. Circulating miRNAs are speculated to reflect tissue expression, to play a central role in cell-to-cell communication and to be associated with disease progression [33, 34, 35]. Other attributes that make miRNAs attractive biomarkers is their stability and robust expression [22], even in degraded RNA samples [36]. Technological advances and the development of various platforms for miRNA profiling have bolstered the popularity of miRNAs [37], and have enabled relative easy and cost-effective methods of quantification using sensitive techniques such as quantitative real time PCR (qRT-PCR) [22, 32, 38]. Circulating miRNAs are thought to be released from cells as exosomes, microvesicles, apoptotic bodies, or are non-vesicle bound and encapsulated in protein or lipid complexes [32]. A number of studies [39, 40, 41], have demonstrated that circulating miRNAs are associated with glucose homeostasis and are dysregulated during T2D progression. Recently, we showed that the expression of miR-27b is increased in peripheral blood cells and serum of South African women with impaired glucose tolerance compared to normoglycaemia [42] and identified novel miRNAs associated with dysglycaemia in these women [43]. Putative gene targets of these novel miRNAs were enriched in biological processes involved in key aspects of glucose regulation, and receiver operating characteristic (ROC) curve analysis demonstrated that the diagnostic utility of these miRNAs were similar to fasting insulin [43]. Intriguingly, Parrizas et al. showed that an exercise intervention was able to reverse the aberrant expression of miR-192 and miR-193b induced by impaired glucose tolerance [44], while Luo et al. [45] showed that platelet-derived miR-126 was altered during T2D progression, and that glucose lowering treatment was able to normalise its expression. These studies provide support for the use of miRNAs as diagnostic and prognostic biomarkers to monitor treatment response.

Advertisement

5. MicroRNAs, pregnancy and gestational diabetes

MiRNAs play an important role as metabolic and developmental regulators during pregnancy. They respond to changing physiological conditions during pregnancy, while their dysregulation contributes to pregnancy-related disorders [46]. Thus far over 600 placental miRNAs have been identified [47]. Altered placental miRNA expression has been demonstrated in several pregnancy related disorders. In 2007, Pineles et al. were the first to demonstrate altered miRNA expression during preeclampsia. They reported that the expression of two miRNAs, miR-210 and miR-182, were increased during preeclampsia [48]. In 2009, using microarrays, Hu et al. and Zhu et al. identified seven and 34 miRNAs, respectively, that are dysregulated in preeclamptic compared to normal pregnancies [49, 50]. Subsequently, other studies have reported differential miRNA expression during preeclampsia and importantly provide experimental evidence to support the involvement of these miRNAs in disease pathophysiology [51, 52]. Altered miRNA expression has also been observed in other pregnancy complications such as macrosomia [53], preterm delivery and small for gestational age [54].

Placental-derived miRNAs in maternal blood have potential as biomarkers for pregnancy monitoring [54, 55]. It is suggested that miRNAs from placental tissue are exported into the maternal circulation via exosomes, and that these miRNAs reflect the physiological status of pregnancy and may thus have diagnostic potential [56]. Many of studies have reported that maternal circulating miRNAs are associated with placental weight [57], placental dysfunction [58, 59] and pregnancy complications [54, 60, 61, 62, 63]. MiR-517c was increased in pregnancies complicated with placental abruption [59], miR-515, miR-516a, miR-516b, miR-518b, miR-519d, miR-520a, miR-520h, miR-525, miR-526b and miR-1323 were increased during preeclampsia [63], miR-516, miR-517, miR-520a, miR-525 and miR-526a were upregulated during preeclampsia, gestational hypertension and foetal growth restriction [60], miR-517a was increased and miR-518b was decreased during placenta previa [58], and miR-346 and miR-582 were increased during preeclampsia, preterm delivery and small for gestational age patients compared to normal controls [54]. Importantly, Hromadnikova et al. showed that upregulation of plasma miR-517, miR-518b and miR-520h during the first trimester was associated with the development of preeclampsia, and provided evidence to suggest that miR-517 could predict preeclampsia [61]. Furthermore, a number of studies have reported that circulating miRNAs are associated with macrosomia [62, 64, 65]. Jiang and colleagues demonstrated that maternal expression of miR-21, and to a lesser extent miR-20a in serum samples from pregnant women in the third trimester was associated with macrosomia [62], Hu et al. reported that macrosomia was associated with decreased serum expression of miR-376a [64], while Ge et al. reported that the expression of miR-18a, miR-141, miR-143, miR-200c and miR-221 were decreased, and miR-16, miR-30a and miR-523 were increased in the plasma of pregnant women with foetal macrosomia compared to normal controls [65], further supporting the use of miRNAs as predictive biomarkers for pregnancy complications.

Growing evidence implicate miRNAs in the pathogenesis of GDM [47] and suggest that maternal miRNA expression may be used as biomarkers to predict GDM. Indeed, many GDM associated miRNAs are also expressed in placentas of women with Type 1 diabetes and T2D, confirming that miRNAs expressed during GDM play an important role in metabolic regulation and reflects some of the shared aetiology between these different types of diabetes [66]. Interestingly, a subset of miRNAs were distinct for each type of diabetes, illustrating their potential to differentiate between GDM and other manifestations of diabetes. Several other studies have reported that placental miRNA expression is altered in women with GDM. Zhao et al. reported that miR-518d is upregulated in placentas of women with GDM compared to controls, and further showed that increased expression of miR-518d correlated with decreased protein expression of peroxisome proliferator-activated receptor-α (PPARα) [67], a major regulatory transcription factor in lipid homeostasis and energy metabolism [68, 69]. Li et al. identified nine miRNAs that are dysregulated in placentas of women with GDM, the expression of miR-508 was increased and miR-9, miR-27a, miR-30d, miR-33a, miR-92a, miR-137, miR-362 and miR-502 were decreased. Importantly, the decreased expression of these miRNAs correlated with increased protein expression of their gene targets, epidermal growth factor receptor (EGFR), phosphoinositide 3-Kinase (PI3K) and protein kinase B (Akt), key proteins in placental development and foetal growth [70]. Other studies showed that miR-98 [71] and miR-503 [72] are upregulated and miR-143 is downregulated [73] in placentas of women with GDM compared to women with normoglycaemic pregnancies. Intriguingly, the expression of miR-143 differentiated between GDM managed by diet or medication [71], further supporting the clinical value of miRNAs.

Xu et al. showed that the increased expression of miR-503 in the placentas of women with GDM compared to normoglycaemic pregnancies, are reflected in plasma [72]. The studies that have quantified circulating miRNA expression during GDM are summarised in Table 1. In 2011, Zhao et al. were the first to profile the expression of serum miRNAs during GDM [74]. Using Taqman low density arrays, followed by confirmation with individual qRT-PCR, they found that serum expression of miR-29a, miR-132 and miR-222 were decreased during GDM. Importantly, these results were validated in an internal and external cohort. Notably, serum for miRNA profiling in the discovery cohort was collected at 16–19 weeks of pregnancy, while GDM was diagnosed at 24–28 weeks of pregnancy, thus illustrating the potential of these miRNAs as screening tools for GDM. Recently, Pheiffer et al. reported that the expression of miR-29a, miR-132 and miR-222 were similarly decreased in the serum of South African women with GDM, however, only the latter was statistically significant [75]. Conflictingly, Tagoma et al. showed that miR-222 was increased in plasma of women with GDM compared to normoglycaemic pregnancies [76]. Moreover, Wander et al., observed no differences in the expression of miR-29a and miR-222 in the plasma of American women with or without GDM [77], thus illustrating the heterogeneity of miRNA expression.

GDM/
controls
Biological source Detection method Up-
regulated
Down-regulated No change Normalisation Ref
24/24*
(16–19 weeks gestation)
Serum Taqman low density array
qRT-PCR
miR-29a
miR-132
miR-222
Cel-miR-39 [74]
28/53
(13–31 weeks gestation)
Serum qRT-PCR miR-20a
miR-222
miR-16
miR-17
miR-19a
miR-19b
miR-29a
miR-132
Cel-miR-39 [75]
13/9
(23–31 weeks gestation)
Plasma qRT-PCR let-7e
let-7 g
miR-100
miR-101
miR-146a miR-8a
miR-195
miR-222
miR-23b
miR-30b
miR-30c
miR-30d
miR-342
miR-423
miR-92a
Cel-miR-39 [76]
36/80
(7–23 weeks gestation)
Plasma qRT-PCR miR-155
miR-21
miR-146b miR-517
miR-222
miR-210
miR-518a
miR-29a
miR-223
miR-126
Cel-miR-39 and miR-423 [77]
10/10
(16–19 weeks gestation)
Plasma Sequencing qRT-PCR miR-16
miR-17
miR-19a
miR-19b
miR-20a
miR-221 [78]
85 GDM and 72 controls (16–20, 20–24 and 24–28 weeks gestation) Plasma qRT-PCR miR-16
miR-17
miR-20a
miR-19a
miR-19b
RNU6 [79]
21/10
(24–33 weeks gestation)
Plasma qRT-PCR miR-330 miR-548c miR-374
miR-320
[92]
30/30 (24–32 weeks gestation) Whole blood Sequencing
qRT-PCR
miR-340 RNU6B [93]
20/20
NK
Whole
blood
qRT-PCR miR-494 RNU6 [94]
67/74
(16–20, 20–24 and 24–28)
Serum qRT-PCR miR-183
miR-200b miR-125b
miR-1290
Cel-miR-39 [95]
11/12 (third trimester) Plasma qRT-PCR miR-137 RNU6 [81]
25/25
NK
Plasma qRT-PCR miR-503 NK [72]

Table 1.

Studies investigating microRNA expression in maternal blood during gestational diabetes mellitus.

Validated in internal (36 GDM/36 controls) and two external cohorts (16 GDM/16 controls each).


GDM: gestational diabetes mellitus; qRT-PCR: quantitative real time polymerase chain reaction; NK: not known.

In 2015, Zhu et al. used high-throughput sequencing and qRT-PCR to investigate miRNAs in plasma samples of Chinese women with or without GDM [78]. Five miRNAs (miR-16, miR-17, miR-19a, miR-19b and miR-20a) were significantly upregulated in women with GDM compared to controls. Furthermore, the differential expression of these miRNAs were observed at 16–19 weeks of pregnancy, before GDM diagnosis, once again illustrating the diagnostic value of miRNAs [78]. Cao et al. similarly demonstrated increased expression of plasma miR-16, miR-17 and miR-20a in a larger cohort of Chinese women, however, they did not observe differences in the expression of miR-19a and miR-19b [79]. More recently, Pheiffer et al. reported conflicting results. The expression of all five miRNAs were decreased in South African women with GDM, however, only the decreased expression of miR-20a was statistically significant [75]. Interestingly, regression analysis showed that miR-20a was a significant predictor of GDM, while age and body mass index were not.

Although these miRNAs were identified in plasma or serum, bioinformatics [75, 78] and experimental [74] functional analyses provided support for their biological relevance and role in the pathogenesis of GDM. Other studies also confirm the importance of miRNAs during GDM. For example, in 2014, Shi et al. reported that the expression of miR-222 is increased in omental tissue from women with GDM compared to women with normoglycaemic pregnancies, and conducted elegant in vitro experiments to demonstrate that miR-222 potentially regulates oestrogen-induced insulin resistance during GDM. As shown in Table 1, many more miRNAs have been reported to exhibit altered expression in maternal blood during GDM, however, these were investigated in single studies only.

Advertisement

6. Gestational diabetes and foetal microRNA expression

Dysregulated miRNA expression has been reported in human umbilical vein endothelial cells (HUVECs) of foetuses exposed to GDM. Floris et al. reported that impaired HUVEC function during GDM is associated with altered miR-101 expression [80]. Several other miRNAs, miR-137 [81], miR-let-7a, miR-let-7g, miR-30c, miR-126, miR-130b, miR-148a and miR-452 [82] were upregulated in HUVECs from infants born to mothers with GDM, suggesting that miRNAs reflect the adverse in utero environment imposed by GDM. Tryggestad et al. further showed that two of these miRNAs, miR-130b and miR-148a, target and decrease the expression of 5′ Adenosine monophosphate-activated protein kinase (AMPKα1), whose protein expression is decreased in placenta exposed to GDM [82]. Recently, altered miRNA expression in offspring blood was shown to be associated with birth weight [83]. MiR-33b and miR-375 were overexpressed during macrosomia, while miR-454 was overexpressed in blood of both low birth weight and macrosomic compared to normal birth weight offspring [83]. Aberrant miR-346 and miR-582 expression in cord blood were shown to be associated with foetal complications [54]. Taken together, these studies provide evidence that GDM induces dysregulated miRNA expression in offspring, which may predispose them to metabolic disease in later life. Thus, miRNAs offer potential to predict disease in offspring, which could facilitate intervention strategies to prevent future disease.

Advertisement

7. Challenges of microRNA profiling

Despite their stability and relative ease of quantification, analysis of circulating miRNAs present several pre-analytical and analytical challenges [84] that must be addressed before they can be used clinically. Many studies have reported that miRNA expression is affected by sample type, method of miRNA extraction, and quantification and data normalisation strategies. Differences in miRNA expression between whole blood and serum [42], between different cell types in whole blood [85, 86], between plasma and serum [87, 88] and between placenta, plasma and cord blood [54] have been described. Furthermore, miRNA expression varies according to the extraction kit used [88]. Currently, qRT-PCR is considered the gold standard for miRNA analysis, however variations between qRT-PCR platforms [87] and between qRT-PCR and other measurement platforms [22, 42, 87, 88] have been widely reported. Furthermore, data normalisation is a significant challenge during miRNA profiling, particularly extracellular miRNAs [89]. Currently, there is no consensus on the best normalisation strategy to use when profiling circulating miRNAs, although strategies based on exogenous spike-in-controls such as C. elegans miR-39 have been shown to be less variable than using endogenous miRNAs [88]. Moreover, heterogeneous miRNA expression is observed between populations, mediated by both genetic and environmental factors [90, 91]. During pregnancy, gestation time is also reported to affect miRNA expression [55]. Lastly, miRNAs are non-specific. For example, a single miRNA can regulate up to 200 different genes [27, 28], thus miRNAs found to be associated with GDM, may possibly be involved in other conditions as well.

Advertisement

8. Perspectives and recommendations for future research

MiRNAs offer great potential as biomarkers for GDM. However, they face many challenges that need to be addressed before they can become clinically applicable. Standardisation of pre-analytical and analytical methods for miRNA research may minimise the lack of reproducibility between studies and should be prioritised in miRNA research [22]. MiRNAs are epigenetic mechanisms that are regulated by various factors [90, 91], which need to be considered in miRNA studies. Large prospective cohort studies should be conducted to elucidate how biological, genetic and environmental factors affect miRNA expression, and to identify plausible diagnostic or prognostic candidates. Moreover, due to their non-specificity [27, 28], it is recommended that a panel of miRNAs, either alone, or in combination with other risk factors, should be used to increase the specificity of risk stratification models for GDM.

Advertisement

9. Conclusions

In this review the current status of miRNAs as biomarkers for GDM was discussed, together with recommendations for research. We provide evidence to show that miRNAs possess tremendous potential as routine screening tools, which could facilitate earlier diagnosis and management of GDM with dietary modifications or therapeutic intervention. A growing number of studies have demonstrated their clinical utility, and technological advances can lead to the development of inexpensive, point-of-care miRNA diagnostic tests in the future. However, at present miRNA profiling during GDM remains inconclusive, largely due to the irreproducibility of results between studies. Many technical, analytical and biological challenges hamper miRNA research, and must be addressed before these small RNA molecules, which are master regulators of gene expression, can become clinically applicable.

Advertisement

Acknowledgments

This work was funded by the South African Medical Research Council.

Conflict of interest

The authors have no conflict of interest to declare.

References

  1. 1. Ferrara A. Increasing prevalence of gestational diabetes mellitus. Diabetes Care. 2007;30:S141-S146. DOI: 10.2337/dc07-s206
  2. 2. International Diabetes Federation. IDF Diabetes Atlas. 8th ed. Brussels, Belgium: International Diabetes Federation; 2017. http://www.diabetesatlas.org
  3. 3. Agarwal MM. Gestational diabetes mellitus: Screening with fasting plasma glucose. World Journal of Diabetes. 2016;7:279-289. DOI: 10.4239/wjd.v7.i14.279
  4. 4. Bellamy L, Casas J-P, Hingorani AD, Williams D. Type 2 diabetes mellitus after gestational diabetes: A systematic review and meta-analysis. Lancet. 2009;373:1773-1779. DOI: 10.1016/S0140-6736(09)60731-5
  5. 5. Mitanchez D, Yzydorczyk C, Siddeek B, Boubred F, Benahmed M, Simeoni U. The offspring of the diabetic mother—Short- and long-term implications. Best Practice & Research. Clinical Obstetrics & Gynaecology. 2015;29:256-269. DOI: 10.1016/j.bpobgyn.2014.08.004
  6. 6. Ratner RE. Prevention of type 2 diabetes in women with previous gestational diabetes. Diabetes Care. 2007;30:S242-S245. DOI: 10.2337/dc07-s223
  7. 7. Falavigna M, Schmidt MI, Trujillo J, Alves LF, Wendland ER, Torloni MR, et al. Effectiveness of gestational diabetes treatment: A systematic review with quality of evidence assessment. Diabetes Research and Clinical Practice. 2012;98:396-405. DOI: 10.1016/j.diabres.2012.09.002
  8. 8. Duran A, Sáenz S, Torrejón MJ, Bordiú E, Del Valle L, Galindo M, et al. Introduction of IADPSG criteria for the screening and diagnosis of gestational diabetes mellitus results in improved pregnancy outcomes at a lower cost in a large cohort of pregnant women: The St. Carlos Gestational Diabetes Study. Diabetes Care. 2014;37:2442-2450. DOI: 10.2337/dc14-0179
  9. 9. Hernandez TL, Van Pelt RE, Anderson MA, Daniels LJ, West NA, Donahoo WT, et al. A higher-complex carbohydrate diet in gestational diabetes mellitus achieves glucose targets and lowers postprandial lipids: A randomized crossover study. Diabetes Care. 2014;37:1254-1262. DOI: 10.2337/dc13-2411
  10. 10. Kelley KW, Carroll DG, Meyer A. A review of current treatment strategies for gestational diabetes mellitus. Drugs in Context. 2015;4:212282. DOI: 10.7573/dic.212282
  11. 11. Santangelo C, Zicari A, Mandosi E, Scazzocchio B, Mari E, Morano S, et al. Could gestational diabetes mellitus be managed through dietary bioactive compounds? Current knowledge and future perspectives. British Journal of Nutrition. 2016;115:1129-1144. DOI: 10.1017/S0007114516000222
  12. 12. World Health Organization. Definition, Diagnosis and Classification of Diabetes Mellitus and its Complications. Geneva: World Health Organization; 1999
  13. 13. Xu Y, Shen S, Sun L, Yang H, Jin B, Cao X. Metabolic syndrome risk after gestational diabetes: A systematic review and meta-analysis. PLoS One. 2014;9:e87863. DOI: 10.1371/journal.pone.0087863
  14. 14. Li J-W, He S-Y, Liu P, Luo L, Zhao L, Xiao Y-B. Association of gestational diabetes mellitus (GDM) with subclinical atherosclerosis: A systemic review and meta-analysis. BMC Cardiovascular Disorders. 2014;14:132. DOI: 10.1186/1471-2261-14-132
  15. 15. Park YM, O’Brien KM, Zhao S, Weinberg CR, Baird DD, Sandler DP. Gestational diabetes mellitus may be associated with increased risk of breast cancer. British Journal of Cancer. 2017;116:960-963. DOI: 10.1038/bjc.2017.34
  16. 16. Silveira PP, Portella AK, Goldani MZ, Barbieri MA. Developmental origins of health and disease (DOHaD). The Journal of Pediatrics. 2007;83:494-504. DOI: 10.2223/JPED.1728
  17. 17. Barker DJ. Maternal nutrition, fetal nutrition, and disease in later life. Nutrition. 1997;13:807-813
  18. 18. Chu SY, Callaghan WM, Kim SY, Schmid CH, Lau J, England LJ, et al. Maternal obesity and risk of gestational diabetes mellitus. Diabetes Care. 2007;30:2070-2076. DOI: 10.2337/dc06-2559a
  19. 19. Adam S, Pheiffer C, Dias S, Rheeder P. Association between gestational diabetes and biomarkers: A role in diagnosis. Biomarkers. 2018;23:386-391. DOI: 10.1080/1354750X.2018.1432690
  20. 20. Cosson E, Benbara A, Pharisien I, Nguyen MT, Revaux A, Lormeau B, et al. Diagnostic and prognostic performances over 9 years of a selective screening strategy for gestational diabetes mellitus in a cohort of 18,775 subjects. Diabetes Care. 2013;36:598-603. DOI: 10.2337/dc12-1428
  21. 21. Hulka B. Overview of biological markers. In: Hulka BS, Griffith JD, Wilcosky TC, editors. Biological Markers in Epidemiology. New York: Oxford University Press; 1990. pp. 3-15
  22. 22. Etheridge A, Lee I, Hood L, Galas D, Wang K. Extracellular microRNA: A new source of biomarkers. Mutation Research. 2011;717:85-90. DOI: 10.1016/j.mrfmmm.2011.03.004
  23. 23. Guay C, Roggli E, Nesca V, Jacovetti C, Regazzi R. Diabetes mellitus, a microRNA-related disease? Translational Research. 2011;157:253-264. DOI: 10.1016/j.trsl.2011.01.009
  24. 24. Lee RC, Feinbaum RL, Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75:843-854
  25. 25. Bartel DP. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell. 2004;116:281-297
  26. 26. Guo H, Ingolia NT, Weissman JS, Bartel DP. Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature. 2010;466:835-840. DOI: 10.1038/nature09267
  27. 27. Krek A, Grün D, Poy MN, Wolf R, Rosenberg L, Epstein EJ, et al. Combinatorial microRNA target predictions. Nature Genetics. 2005;37:495-500. DOI: 10.1038/ng1536
  28. 28. Lewis BP, Burge CB, Bartel DP. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell. 2005;120:15-20. DOI: 10.1016/j.cell.2004.12.035
  29. 29. Friedman RC, Farh KK-H, Burge CB, Bartel DP. Most mammalian mRNAs are conserved targets of microRNAs. Genome Research. 2009;19:92-105. DOI: 10.1101/gr.082701.108
  30. 30. Ardekani AM, Naeini MM. The role of MicroRNAs in human diseases. Avicenna Journal of Medical Biotechnology. 2010;2:161-179
  31. 31. Regazzi R. MicroRNAs as therapeutic targets for the treatment of diabetes mellitus and its complications. Expert Opinion on Therapeutic Targets. 2018;22:153-160. DOI: 10.1080/14728222.2018.1420168
  32. 32. Creemers EE, Tijsen AJ, Pinto YM. Circulating microRNAs: Novel biomarkers and extracellular communicators in cardiovascular disease? Circulation Research. 2012;110:483-495. DOI: 10.1161/CIRCRESAHA.111.247452
  33. 33. Guay C, Regazzi R. Exosomes as new players in metabolic organ cross-talk. Diabetes, Obesity & Metabolism. 2017;19:S137-S146. DOI: 10.1111/dom.13027
  34. 34. Turchinovich A, Weiz L, Burwinkel B. Extracellular miRNAs: The mystery of their origin and function. Trends in Biochemical Sciences. 2012;37:460-465. DOI: 10.1016/j.tibs.2012.08.003
  35. 35. Guay C, Regazzi R. New emerging tasks for microRNAs in the control of β-cell activities. Biochimica et Biophysica Acta. 2016;1861:2121-2129. DOI: 10.1016/j.bbalip.2016.05.003
  36. 36. Jung M, Schaefer A, Steiner I, Kempkensteffen C, Stephan C, Erbersdobler A, et al. Robust microRNA stability in degraded RNA preparations from human tissue and cell samples. Clinical Chemistry. 2010;56:998-1006. DOI: 10.1373/clinchem.2009.141580
  37. 37. Pritchard CC, Cheng HH, Tewari M. MicroRNA profiling: Approaches and considerations. Nature Reviews. Genetics. 2012;13:358-369. DOI: 10.1038/nrg3198
  38. 38. Mayeux R. Biomarkers: Potential uses and limitations. NeuroRx. 2004;1:182-188. DOI: 10.1602/neurorx.1.2.182
  39. 39. Zampetaki A, Kiechl S, Drozdov I, Willeit P, Mayr U, Prokopi M, et al. Plasma microRNA profiling reveals loss of endothelial miR-126 and other microRNAs in type 2 diabetes. Circulation Research. 2010;107:810-817. DOI: 10.1161/CIRCRESAHA.110.226357
  40. 40. Zampetaki A, Mayr M. MicroRNAs in vascular and metabolic disease. Circulation Research. 2012;110:508-522. DOI: 10.1161/CIRCRESAHA.111.247445
  41. 41. Karolina DS, Armugam A, Tavintharan S, Wong MT, Lim SC, Sum CF, et al. MicroRNA 144 impairs insulin signaling by inhibiting the expression of insulin receptor substrate 1 in type 2 diabetes mellitus. PLoS One. 2011;6:e22839. DOI: 10.1371/journal.pone.0022839
  42. 42. Dias HS, Muller C, Louw J, Pheiffer C. MicroRNA expression varies according to glucose tolerance, measurement platform, and biological source. BioMed Research International. 2017;2017:1080157. DOI: 10.1155/2017/1080157
  43. 43. Pheiffer C, Dias S, Willmer T, Pace R, Aagaard K, Louw J. Altered microRNA expression during impaired glucose tolerance and high-fat diet feeding. Experimental and Clinical Endocrinology & Diabetes. 11 Jun 2018. DOI: 10.1055/a-0619-4576. [Epub ahead of print]
  44. 44. Párrizas M, Brugnara L, Esteban Y, González-Franquesa A, Canivell S, Murillo S, et al. Circulating miR-192 and miR-193b are markers of prediabetes and are modulated by an exercise intervention. The Journal of Clinical Endocrinology and Metabolism. 2015;100:E407-E415. DOI: 10.1210/jc.2014-2574
  45. 45. Luo M, Li R, Deng X, Ren M, Chen N, Zeng M, et al. Platelet-derived miR-103b as a novel biomarker for the early diagnosis of type 2 diabetes. Acta Diabetologica. 2015;52:943-949. DOI: 10.1007/s00592-015-0733-0
  46. 46. Cai M, Kolluru GK, Ahmed A. Small molecule, big prospects: MicroRNA in pregnancy and its complications. Journal of Pregnancy. 2017;2017:6972732. DOI: 10.1155/2017/6972732
  47. 47. Poirier C, Desgagné V, Guérin R, Bouchard L. MicroRNAs in pregnancy and gestational diabetes mellitus: Emerging role in maternal metabolic regulation. Current Diabetes Reports. 2017;17:35. DOI: 10.1007/s11892-017-0856-5
  48. 48. Pineles BL, Romero R, Montenegro D, Tarca AL, Han YM, Kim YM, et al. Distinct subsets of microRNAs are expressed differentially in the human placentas of patients with preeclampsia. American Journal of Obstetrics and Gynecology. 2007;196:261.e1-261.e6. DOI: 10.1016/j.ajog.2007.01.008
  49. 49. Zhu X, Han T, Sargent IL, Yin G, Yao Y. Differential expression profile of microRNAs in human placentas from preeclamptic pregnancies vs normal pregnancies. American Journal of Obstetrics and Gynecology. 2009;200:661.e1-661.e7. DOI: 10.1016/j.ajog.2008.12.045
  50. 50. Hu Y, Li P, Hao S, Liu L, Zhao J, Hou Y. Differential expression of microRNAs in the placentae of Chinese patients with severe pre-eclampsia. Clinical Chemistry and Laboratory Medicine. 2009;47:923-929. DOI: 10.1515/CCLM.2009.228
  51. 51. Wang W, Feng L, Zhang H, Hachy S, Satohisa S, Laurent LC, et al. Preeclampsia up-regulates angiogenesis-associated MicroRNA (i.e., miR-17, -20a, and -20b) that target Ephrin-B2 and EPHB4 in human placenta. The Journal of Clinical Endocrinology and Metabolism. 2012;97:E1051-E1059. DOI: 10.1210/jc.2011-3131.
  52. 52. Niu Z-R, Han T, Sun X-L, Luan L-X, Gou W-L, Zhu X-M. MicroRNA-30a-3p is overexpressed in the placentas of patients with preeclampsia and affects trophoblast invasion and apoptosis by its effects on IGF-1. American Journal of Obstetrics and Gynecology. 2018;218:249.e1-249.e12. DOI: 10.1016/j.ajog.2017.11.568
  53. 53. Li J, Chen L, Tang Q, Wu W, Gu H, Liu L, et al. The role, mechanism and potentially novel biomarker of microRNA-17-92 cluster in macrosomia. Scientific Reports. 2015;5:17212. DOI: 10.1038/srep17212
  54. 54. Tsai P-Y, Li S-H, Chen W-N, Tsai H-L, Su M-T. Differential miR-346 and miR-582-3p expression in association with selected maternal and fetal complications. International Journal of Molecular Sciences. 2017;18.pii:E1570. DOI: 10.3390/ijms18071570
  55. 55. Chim SS, Shing TK, Hung EC, Leung T-Y, Lau T-K, Chiu RW, et al. Detection and characterization of placental microRNAs in maternal plasma. Clinical Chemistry. 2008;54:482-490. DOI: 10.1373/clinchem.2007.097972
  56. 56. Luo S-S, Ishibashi O, Ishikawa G, Ishikawa T, Katayama A, Mishima T, et al. Human villous trophoblasts express and secrete placenta-specific microRNAs into maternal circulation via exosomes. Biology of Reproduction. 2009;81:717-729. DOI: 10.1095/biolreprod.108.075481
  57. 57. Miura K, Morisaki S, Abe S, Higashijima A, Hasegawa Y, Miura S, et al. Circulating levels of maternal plasma cell-free pregnancy-associated placenta-specific microRNAs are associated with placental weight. Placenta. 2014;35:848-851. DOI: 10.1016/j.placenta.2014.06.002
  58. 58. Hasegawa Y, Miura K, Higashijima A, Abe S, Miura S, Yoshiura K, et al. Increased levels of cell-free miR-517a and decreased levels of cell-free miR-518b in maternal plasma samples from placenta previa pregnancies at 32 weeks of gestation. Reproductive Sciences. 2015;22:1569-1576. DOI: 10.1177/1933719115589407
  59. 59. Miura K, Higashijima A, Murakami Y, Fuchi N, Tsukamoto O, Abe S, et al. Circulating levels of pregnancy-associated, placenta-specific microRNAs in pregnant women with placental abruption. Reproductive Sciences. 2017;24:148-155. DOI: 10.1177/1933719116653837
  60. 60. Hromadnikova I, Kotlabova K, Ondrackova M, Kestlerova A, Novotna V, Hympanova L, et al. Circulating C19MC microRNAs in preeclampsia, gestational hypertension, and fetal growth restriction. Mediators of Inflammation. 2013;2013:186041. DOI: 10.1155/2013/186041
  61. 61. Hromadnikova I, Kotlabova K, Ivankova K, Krofta L. First trimester screening of circulating C19MC microRNAs and the evaluation of their potential to predict the onset of preeclampsia and IUGR. PLoS One. 2017;12:e0171756. DOI: 10.1371/journal.pone.0171756
  62. 62. Jiang H, Wen Y, Hu L, Miao T, Zhang M, Dong J. Serum microRNAs as diagnostic biomarkers for macrosomia. Reproductive Sciences. 2015;22:664-671. DOI: 10.1177/1933719114561557
  63. 63. Miura K, Higashijima A, Murakami Y, Tsukamoto O, Hasegawa Y, Abe S, et al. Circulating chromosome 19 miRNA cluster microRNAs in pregnant women with severe pre-eclampsia. The Journal of Obstetrics and Gynaecology Research. 2015;41:1526-1532. DOI: 10.1111/jog.12749
  64. 64. Hu L, Han J, Zheng F, Ma H, Chen J, Jiang Y, et al. Early second-trimester serum microRNAs as potential biomarker for nondiabetic macrosomia. BioMed Research International. 2014;2014:394125. DOI: 10.1155/2014/394125
  65. 65. Ge Q, Zhu Y, Li H, Tian F, Xie X, Bai Y. Differential expression of circulating miRNAs in maternal plasma in pregnancies with fetal macrosomia. International Journal of Molecular Medicine. 2015;35:81-91. DOI: 10.3892/ijmm.2014.1989
  66. 66. Collares CV, Evangelista AF, Xavier DJ, Rassi DM, Arns T, Foss-Freitas MC, et al. Identifying common and specific microRNAs expressed in peripheral blood mononuclear cell of type 1, type 2, and gestational diabetes mellitus patients. BMC Research Notes. 2013;6:491. DOI: 10.1186/1756-0500-6-491
  67. 67. Zhao C, Zhang T, Shi Z, Ding H, Ling X. MicroRNA-518d regulates PPARα protein expression in the placentas of females with gestational diabetes mellitus. Molecular Medicine Reports. 2014;9:2085-2090. DOI: 10.3892/mmr.2014.2058
  68. 68. Abbott BD. Review of the expression of peroxisome proliferator-activated receptors alpha (PPAR alpha), beta (PPAR beta), and gamma (PPAR gamma) in rodent and human development. Reproductive Toxicology. 2009;27:246-257. DOI: 10.1016/j.reprotox.2008.10.001
  69. 69. Pawlak M, Lefebvre P, Staels B. Molecular mechanism of PPARα action and its impact on lipid metabolism, inflammation and fibrosis in non-alcoholic fatty liver disease. Journal of Hepatology. 2015;62:720-733. DOI: 10.1016/j.jhep.2014.10.039
  70. 70. Li J, Song L, Zhou L, Wu J, Sheng C, Chen H, et al. A microRNA signature in gestational diabetes mellitus associated with risk of macrosomia. Cellular Physiology and Biochemistry. 2015;37:243-252. DOI: 10.1159/000430349
  71. 71. Cao J-L, Zhang L, Li J, Tian S, Lv X-D, Wang X-Q, et al. Up-regulation of miR-98 and unraveling regulatory mechanisms in gestational diabetes mellitus. Scientific Reports. 2016;6:32268. DOI: 10.1038/srep32268
  72. 72. Xu K, Bian D, Hao L, Huang F, Xu M, Qin J, et al. MicroRNA-503 contribute to pancreatic beta cell dysfunction by targeting the mTOR pathway in gestational diabetes mellitus. EXCLI Journal. 2017;16:1177-1187. DOI: 10.17179/excli2017-738
  73. 73. Muralimanoharan S, Maloyan A, Myatt L. Mitochondrial function and glucose metabolism in the placenta with gestational diabetes mellitus: Role of miR-143. Clinical Science. 2016;130:931-941. DOI: 10.1042/CS20160076
  74. 74. Zhao C, Dong J, Jiang T, Shi Z, Yu B, Zhu Y, et al. Early second-trimester serum miRNA profiling predicts gestational diabetes mellitus. PLoS One. 2011;6:e23925. DOI: 10.1371/journal.pone.0023925
  75. 75. Pheiffer C, Dias S, Rheeder P, Adam S. Decreased expression of circulating miR-20a-5p in south African women with gestational diabetes mellitus. Molecular Diagnosis & Therapy. 2018;22:345-352. DOI: 10.1007/s40291-018-0325-0
  76. 76. Tagoma A, Alnek K, Kirss A, Uibo R, Haller-Kikkatalo K. MicroRNA profiling of second trimester maternal plasma shows upregulation of miR-195-5p in patients with gestational diabetes. Gene. 2018;672:137-142. DOI: 10.1016/j.gene.2018.06.004
  77. 77. Wander PL, Boyko EJ, Hevner K, Parikh VJ, Tadesse MG, Sorensen TK, et al. Circulating early- and mid-pregnancy microRNAs and risk of gestational diabetes. Diabetes Research and Clinical Practice. 2017;132:1-9. DOI: 10.1016/j.diabres.2017.07.024
  78. 78. Zhu Y, Tian F, Li H, Zhou Y, Lu J, Ge Q. Profiling maternal plasma microRNA expression in early pregnancy to predict gestational diabetes mellitus. International Journal of Gynaecology and Obstetrics. 2015;130:49-53. DOI: 10.1016/j.ijgo.2015.01.010
  79. 79. Cao Y-L, Jia Y-J, Xing B-H, Shi D-D, Dong X-J. Plasma microRNA-16-5p, -17-5p and -20a-5p: Novel diagnostic biomarkers for gestational diabetes mellitus. The Journal of Obstetrics and Gynaecology Research. 2017;43:974-981. DOI: 10.1111/jog.13317
  80. 80. Floris I, Descamps B, Vardeu A, Mitić T, Posadino AM, Shantikumar S, et al. Gestational diabetes mellitus impairs fetal endothelial cell functions through a mechanism involving microRNA-101 and histone methyltransferase enhancer of zester homolog-2. Arteriosclerosis, Thrombosis, and Vascular Biology. 2015;35:664-674. DOI: 10.1161/ATVBAHA.114.304730
  81. 81. Peng H-Y, Li H-P, Li M-Q. High glucose induces dysfunction of human umbilical vein endothelial cells by upregulating miR-137 in gestational diabetes mellitus. Microvascular Research. 2018;118:90-100. DOI: 10.1016/j.mvr.2018.03.002
  82. 82. Tryggestad JB, Vishwanath A, Jiang S, Mallappa A, Teague AM, Takahashi Y, et al. Influence of gestational diabetes mellitus on human umbilical vein endothelial cell miRNA. Clinical Science. 2016;130:1955-1967. DOI: 10.1042/CS20160305
  83. 83. Rodil-Garcia P, Arellanes-Licea EDC, Montoya-Contreras A, Salazar-Olivo LA. Analysis of microRNA expression in newborns with differential birth weight using newborn screening cards. International Journal of Molecular Sciences. 2017;18(12):pii:E2552. DOI: 10.3390/ijms18122552
  84. 84. Kroh EM, Parkin RK, Mitchell PS, Tewari M. Analysis of circulating microRNA biomarkers in plasma and serum using quantitative reverse transcription-PCR (qRT-PCR). Methods. 2010;50:298-301. DOI: 10.1016/j.ymeth.2010.01.032
  85. 85. Leidinger P, Backes C, Dahmke IN, Galata V, Huwer H, Stehle I, et al. What makes a blood cell based miRNA expression pattern disease specific?–A miRNome analysis of blood cell subsets in lung cancer patients and healthy controls. Oncotarget. 2014;5:9484-9497. DOI: 10.18632/oncotarget.2419
  86. 86. Leidinger P, Backes C, Meder B, Meese E, Keller A. The human miRNA repertoire of different blood compounds. BMC Genomics. 2014;15:474. DOI: 10.1186/1471-2164-15-474
  87. 87. Wang K, Yuan Y, Cho J-H, McClarty S, Baxter D, Galas DJ. Comparing the microRNA spectrum between serum and plasma. PLoS One. 2012;7:e41561. DOI: 10.1371/journal.pone.0041561
  88. 88. Vigneron N, Meryet-Figuière M, Guttin A, Issartel J-P, Lambert B, Briand M, et al. Towards a new standardized method for circulating miRNAs profiling in clinical studies: Interest of the exogenous normalisation to improve miRNA signature accuracy. Molecular Oncology. 2016;10:981-992. DOI: 10.1016/j.molonc.2016.03.005
  89. 89. Schwarzenbach H, da Silva AM, Calin G, Pantel K. Which is the accurate data normalisation strategy for microRNA quantification? Clinical Chemistry. 2015;61:1333-1342. DOI: 10.1373/clinchem.2015.239459
  90. 90. Huang RS, Gamazon ER, Ziliak D, Wen Y, Im HK, Zhang W, et al. Population differences in microRNA expression and biological implications. RNA Biology. 2011;8:692-701. DOI: 10.4161/rna.8.4.16029
  91. 91. Gulyaeva LF, Kushlinskiy NE. Regulatory mechanisms of microRNA expression. Journal of Translational Medicine. 2016;14:143. DOI: 10.1186/s12967-016-0893-x

Written By

Carmen Pheiffer, Stephanie Dias, Paul Rheeder and Sumaiya Adam

Submitted: 17 July 2018 Reviewed: 18 October 2018 Published: 19 November 2018