Next Article in Journal
Fluorescent Molecular Logic Gates and Pourbaix Sensors in Polyacrylamide Hydrogels
Next Article in Special Issue
Dietary Antioxidant Curcumin Mitigates CuO Nanoparticle-Induced Cytotoxicity through the Oxidative Stress Pathway in Human Placental Cells
Previous Article in Journal
Isolation and Identification of an α-Galactosidase-Producing Lactosphaera pasteurii Strain and Its Enzymatic Expression Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Micro-RNAs in Human Placenta: Tiny Molecules, Immense Power

1
National Clinical Research Center for Child Health of the Children’s Hospital, Zhejiang University School of Medicine, Hangzhou 310052, China
2
Department of Obstetrics, Tongde Hospital of Zhejiang Province, Hangzhou 310012, China
3
State Key Laboratory of Ophthalmology, Zhongshan Ophthalmic Center, Sun Yat-sen University, Guangzhou 510060, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(18), 5943; https://doi.org/10.3390/molecules27185943
Submission received: 4 August 2022 / Revised: 30 August 2022 / Accepted: 9 September 2022 / Published: 13 September 2022
(This article belongs to the Special Issue Key Molecules in the Pregnancy)

Abstract

:
Micro-RNAs (miRNAs) are short non-coding single-stranded RNAs that modulate the expression of various target genes after transcription. The expression and distribution of kinds of miRNAs have been characterized in human placenta during different gestational stages. The identified miRNAs are recognized as key mediators in the regulation of placental development and in the maintenance of human pregnancy. Aberrant expression of miRNAs is associated with compromised pregnancies in humans, and dysregulation of those miRNAs contributes to the occurrence and development of related diseases during pregnancy, such as pre-eclampsia (PE), fetal growth restriction (FGR), gestational diabetes mellitus (GDM), recurrent miscarriage, preterm birth (PTB) and small-for-gestational-age (SGA). Thus, having a better understanding of the expression and functions of miRNAs in human placenta during pregnancy and thereby developing novel drugs targeting the miRNAs could be a potentially promising method in the prevention and treatment of relevant diseases in future. Here, we summarize the current knowledge of the expression pattern and function regulation of miRNAs in human placental development and related diseases.

1. Introduction

Small non-coding RNAs (ncRNAs) are a group of short non-coding RNAs that play vital roles in many biological processes and diseases [1]. Among those ncRNAs, micro-RNA (miRNA) is a small endogenous single-stranded RNA that regulates the expression of a variety of target genes after transcription. The first ncRNA was characterized in 1965 in baker’s yeast, but the first miRNA, named lin-4, was discovered by Lee et al. in 1993 [2]. As a non-coding RNA with 22 nucleotides in length, lin-4 miRNA modulates the timing of post-embryonic development through suppressing protein expression levels of lin-14 in Caenorhabditis elegans [2]. Later in 2000, let-7, the second miRNA, was identified in nematodes [3]. Same as lin-4 and let-7, miRNAs share a similar mechanism in function by their binding with partial complementarity towards the 3′-untranslated region (3′-UTR) of the potential target genes, respectively. To date, more than one thousand human miRNAs have been characterized so far, and each of them has the potential to control the expression of hundreds of target genes. Interestingly, while one identified characteristics of the miRNAs was the highly conserved sequences throughout species, miRNAs are distributed in a cell-dependent and tissue-specific manner [4]. Considering the nature of miRNAs, which control target gene expression but do not produce proteins themselves, it can be hypothesized that miRNAs may function as key mediators in biologically evolutionary processes and may be particularly involved in the evolution of the complexity among higher mammals. Owing to previous work in the past decades, it is now widely accepted and well believed that the tiny but powerful miRNAs function as important regulators of target genes to control and mediate various physiological or pathophysiological events among organisms including humans [5,6,7], such as the regulation of human placental development and maintenance of human pregnancy.
The placenta is a temporary but the largest organ in the fetus during pregnancy [8], while it functions as the important interface between the maternal and her fetal environments and plays a key role in health support of both the fetus as well as its mother. It is even believed that the placenta has lifelong consequences to both the fetal and maternal future well-being through various functions [9]. Firstly, the placenta participates in the exchange of multiple gases (such as oxygen), kinds of nutrients (such as amino acids) and different waste products (such as metabolic waste) between the mother and the developing fetus [10,11,12]. In addition, unlike murine placenta, the human placenta during pregnancy also functions as a major endocrine organ that produces a wide array of pregnancy-associated proteins including hormones, cytokines and growth factors to sustain fetal growth and maternal physiology, such as estradiol, progesterone and insulin growth factor 2 (IGF2) [13,14,15,16]. Moreover, the placenta serves as a natural barrier to prevent the growing fetus from being affected by the maternal immune attack of the feto–placental unit [17]. A proper placental development is indispensable to the successful pregnancy maintenance and is pivotal to embryo development, while abnormal placental development has come to be the primary defect in most major diseases associated with human pregnancy, such as pre-eclampsia (PE), gestational diabetes mellitus (GDM), fetal growth restriction (FGR), recurrent miscarriage, preterm birth (PTB) as well as small-for-gestational-age (SGA) [18,19,20,21,22,23,24].
Recently, accumulating evidence has indicated that human placental development is precisely controlled by multiple miRNAs. In both in vitro and in vivo studies using cell models and placental tissues, expression of various miRNAs is detected, whose dysregulation gives rise to the aberrant functions of placenta and is strongly associated with pregnancy-related diseases [25,26,27]. The results provide the possible evidence that treatment with molecules or drugs targeting the candidate miRNAs may contribute to the prevention of those diseases. Here, based on recent work from our and other groups, we will summarize the current stage of research on miRNAs in human placenta, and particularly we will discuss the expression, regulation and function of miRNAs in human placenta, as well as their potential involvement in human pregnancy-associated disorders.

2. Biogenesis and Working Mechanisms of miRNAs

Since their discovery in the early 1990s, large numbers of miRNAs have been characterized not only in plants but also in animals [28]. MiRNAs are a dominating class of small RNAs with ~22 nucleotides (nt) in length in almost all somatic tissues and are generated by two RNase III proteins, including Drosha and Dicer [29]. MicroRNA biogenesis initiates from transcription of a miRNA gene and comprises several post-transcriptional modifications that give rise to the maturation of miRNAs, including processing of miRNA by Drosha (a nuclear protein of ~160 kDa) and Dicer (an endonuclease of RNase III type with ~200 kDa at protein weight) in the nucleus and in the cytoplasm, respectively, argonaute protein (AGO) loading and RNA decay [30,31]. MiRNA genes are initially transcribed into long primary transcripts (pri-miRNAs) by RNA polymerase II (Pol II), and are simultaneously regulated by other molecules, such as epigenetic regulators and RNA Pol II-associated transcription factors [32]. The pri-miRNA is characterized by a local hairpin structure and miRNA sequences are embedded within this structure [33,34]. In addition, the hairpin structure is capped with 7-methylguanosine and is polyadenylated [35]. Then, by the nuclear RNase III enzyme Drosha and its co-activator DGCR8, the generated pri-miRNAs are cleaved into precursor miRNAs (pre-miRNAs), which are about 70 nt in length with a characteristic stem–loop structure that is imperfect [36,37]. Next, the produced pre-miRNAs are carried into the cytoplasm from the nucleus by protein Exportin 5 and further undergo cleavage in the presence of a second RNase III enzyme Dicer to form a miRNA duplex [38,39]. Finally, the generated RNA duplex is unwound and the miRNA acting as a guide molecule is subsequently loaded onto an AGO protein to generate an effector complex called the miRNA-induced silencing complex (RISC) for the recognition of potential targets [40] (Figure 1). Of note, RISC assembly mainly consists of two consecutive steps: (1) the loading of the RNA duplex; (2) subsequent unwinding of the RNA duplex [41].
Alterations in RNA sequence or structure will impact the maturation and turnover of miRNAs [42]. The mature miRNAs regulate gene expression at the post-transcriptional level through suppressing protein translation of target genes and/or inducing degradation of target mRNAs. MiRNA-binding sites are at most cases located in the 3′UTR of messenger RNAs (mRNAs), and the region at the 5′-end of miRNAs that spans from nucleotide position 2 to 7 is vital for their target recognition and has been termed the “seed sequence” [43,44]. Effects of mature miRNAs mainly rely on their complementary binding of the “seed sequence” to target sites located at the 3′UTR of correspondingly targeted mRNAs [38]. The nucleotide position of the seed sequence is found to be important for proper target recognition. Imperfect binding of miRNA to their partially complementary sequences located at the 3′UTRs of the targeting mRNAs gives rise to repression of subsequent protein translation; however, near-perfect binding of those miRNAs to their complementary sites within 3′UTR can result in potential cleavage and further degradation of the targeting mRNAs [45,46]. MiRNAs not only recognize 3′ UTR, some miRNAs also bind to targeting sites located at the 5′-untranslated region (5′-UTR). Recently, it has also been indicated that miRNAs negatively regulate gene expression in proliferating cells but up-regulate gene expression in quiescent cells [47]. Given that more than 60% of human protein-coding genes contain at least one conserved miRNA-binding site and that numerous non-conserved sites also exist [48], it is likely that the majority of genes coding proteins may be under the precise control of miRNAs. Thus, it is natural to think that the biogenesis and function of miRNAs are precisely regulated, and that the dysregulation of miRNAs is closely associated with human diseases.

3. Human Placental Development and Structure

The development of human placenta is initiated by the implantation of the blastocyst [49], and the placenta is developed from the trophectoderm (TE) [50]. During 5 days post-fertilization, the pre-implantation embryo is segregated into two different lineages, including the TE and the inner cell mass (ICM). The TE undergoes fusion and consequently forms the primary syncytium [51,52], which intrudes the maternal surface epithelium and further into the underlying maternal endometrium. Subsequently, TE cells undergo differentiation in a temporal and spatial manner and they further continue growing to intrude into the entire decidualized endometrium (decidua) up until the inner third of the myometrium as well as the maternal vasculature [53,54]. The progenitor cells of cytotrophoblasts localize at the basement membrane of multiple placental villi, and those cells thereafter differentiate along two main directions: (1) villous; and (2) extra-villous trophoblasts, both of which occur spontaneously [55,56].
Human placental villi are observed to be surrounded by a layer of mono-nucleated villous cytotrophoblasts (VCTs), which gradually undergoes fusion to form a syncytiotrophoblast (SCT) layer that is multi-nucleated. This multi-nucleated syncytium layer directly interacts with maternal blood and is fundamentally involved in the secretion and transportation of gas, nutrients and waste products across the materno–fetal interface [57,58]. The specifically multi-nucleated SCT layer also acts a key mediator in sustaining human pregnancy through regulating the secretion of pregnancy-related hormones, such as human placental lactogen (hPL), estradiol and human gonadotropin (hCG) [59]. While the non-proliferative villi interact with the maternal decidua at the tips of anchoring villi, VCT proliferates and grows to form cell columns. In the extra-villous pathway, while the trophoblast moves away from the niche of progenitors at the column base, it undergoes a process that is similar to the cellular transformation called epithelial-to-mesenchymal transition (EMT) and further differentiates into extra-villous trophoblasts (EVTs), which can be further divided into two kinds of EVTs that have distinct roles in the regulation of maternal decidua, including (1) interstitial EVTs (iEVTs); and (2) endovascular EVTs (enEVTs) [60,61]. iEVTs intrude into the maternal decidual stroma to remodel the maternal spiral arteries and continue to move as far as the maternal myometrium [62,63]. However, different from iEVTs, enEVTs, acting as a key player, move downwards inside of the spiral arteries and transiently replace the endothelium [64]. All these key steps of human placental development are precisely regulated at a temporal and spatial manner throughout pregnancy, whose disruption can cause improper development and aberrant functions of the placenta and consequently contribute to potential gestational complications.

4. Detection of miRNAs in Human Placenta

Owing to the detection methods such as Northern blot assays [65], quantitative real-time polymerase chain reaction (qRT-PCR) [66], microarrays [67], in situ hybridization (ISH) [68] as well as cloning of short RNAs expressed endogenously [69] and deep sequencing [70], we have obtained exciting findings in miRNA-related studies in human placenta. However, despite the merits with unique features, each method also possesses unsatisfied points. For example, Northern blot assays are able to examine the expression of miRNAs of interest and further provide useful hints on the distribution and existence of individual miRNA precursors within a specific tissue, but the sensitivity and specificity in practice are low. Although qRT-PCR provides high sensitivity in measuring miRNA expression, the throughput is not high enough, resulting in low efficiency.
MicroRNA microarray is a newly developed method with its capacity of powerful high-throughput in profiling the expression of miRNAs (particularly the large miRNAs) at the same time [71]; however, the results are always additionally needed to be verified by some other method such as qRT-PCR to make further confirmation. Different from the above methods, miRNA ISH provides with the possibility to study and observe the localization of some miRNAs of interest in specific cells or tissues; however, the sensitivity of this detection is not perfect due to its low binding affinity [72]. Nevertheless, as a powerful tool in miRNA exploration, cloning of short RNAs that are endogenously expressed is still widely accepted and utilized currently in miRNA-related studies. In addition, as a high-throughput approach for the identification and expression profiling of various miRNAs, deep sequencing enables sequencing of the entire miRNA transcriptome within each sample at a given time [73]. Thus, given the advantages and limitations in each method, ideally, more than single one method should be combined for experiments in one study, particularly in accurate miRNA expression detection, to increase the credibility and reliability of the results obtained.

5. Expression of Micro-RNAs in Human Placenta

It is now believed that many kinds of miRNAs can be efficiently produced by human placenta, particularly by placental trophoblasts [74]. So far, more than 500 miRNAs have been found to be expressed in human placenta [75], and Morales-Prieto et al. have reported that the expression of 762 miRNAs was successfully detected in trophoblast cells isolated from human placenta [76]. The expression of miRNAs in trophoblasts is mediated by various factors, such as hypoxia [77] (summarized in Table 1). Nowadays, hypoxia has emerged to be a common risk that may disrupt human placental development and function during pregnancy, which is associated with potentially increased incidence of gestational diseases. The expression of many miRNAs is controlled by hypoxia in human placenta [78,79,80]. For example, hypoxia up-regulates miRNA-210 (miR-210, well-known as a sensor of hypoxia) expression, which facilitates mitochondrial function in human trophoblast cells [81]. Similarly, the expression of miR-141 is induced by hypoxia and the elevated expression of miR-141 affects trophoblast cell apoptosis, invasion and vascularization [82]. Moreover, miR-218 expression is enhanced by hypoxia to suppress trophoblast cell function [83]. In addition to hypoxia, other factors including environmental factors and epigenetic modification also exert effects on the expression of miRNAs during human pregnancy [84,85,86]. Environmental factors, particularly referred to as environmental stimuli or environmental toxicants, serve as mediators in regulating placental gene expression [84]. As an endocrine disruptor group, phthalates arouse oxidative stress (OS) in human placenta. It is found that maternal exposure to phthalates results in alterations in the expression of the placental miRNA profile [87]. Similarly, mono-(2-ethylhexyl) phthalate induces miRNAs and triggers apoptosis in HTR8/SVneo cells (a kind of first trimester placental cell line), including miR-17-5p, miR-155-5p, miR-126-3p and miR-16 [88,89]. In addition, Bisphenol A (BPA), up-regulates the expression levels of several miRNAs in different trophoblast cell lines, such as miR-146a [90]. Likewise, phenol exposure is associated with alterations of miRNA expression in human placenta, including miR-142-3p, miR15a-5p and miR-185 [91]. Similar phenomena can also be observed from metals and air toxics, which are closely associated with pregnancy-related complications [92].
Epigenetic regulation, involving heritable changes in expression of genes or phenotype of cells without alterations of the underlying DNA sequence, has been shown to participate in controlling the expression of human placental miRNAs. Particularly, the status of DNA methylation has been proved to act as a key mediator in the expression of those miRNAs [93]. One famous example is the C19MC microRNA locus that is imprinted in human placenta [94]. Down-regulated C19MC methylation status is associated with the increased maternal size, which in turn, gives rise to the larger body size of the child [95].
Of note, recently, series of studies have additionally revealed that the presence of single-nucleotide polymorphism (SNP) in miRNA genes also participates in the regulation of miRNA expression and maturation during human pregnancy [96]. For example, Srivastava et al. indicated that miRNA gene polymorphism is closely associated with recurrent pregnancy loss (RPL) in a latest meta-analysis report, where 27 miRNAs with 29 different SNPs are clarified [96]. Similarly, Ezat et al. found that three different miRNA variants, including miR-126 rs4636297, miR-143 rs41291957 and miR-143 rs353292, are correlated with the risk of idiopathic RPL (iRPL) [97].
Table 1. Summary of factors that affect expression of miRNAs in human trophoblasts.
Table 1. Summary of factors that affect expression of miRNAs in human trophoblasts.
FactorsAffected miRNAs in Human Trophoblasts
HypoxiaUp-regulatedmiR-210 [81], miR-141 [82], miR-218 [83]
Environment FactorsPhthalateUp-regulatedmiR-17-5p [88], miR-155-5p [88], miR-126-3p [88], miR-16 [89]
Bisphenol AUp-regulatedmiR-146a [90]
PhenolUp-regulatedmiR-142-3p [91], miR15a-5p [91], miR-185 [91]
Epigenetic RegulationMethylationUp-regulatedC19MC [95]
Single-Nucleotide Polymorphism (SNP)-27 miRNAs [96], miR-126 [97], miR-143 [97], miR-143 [97]
Evidence is emerging to suggest that a great deal of miRNAs are highly expressed in human placenta, and interestingly, some miRNA genes are expressed in a time-dependent and/or tissue-specific manner during different developmental stages of human placenta, which is in agreement with their functions in modulating placental development and trophoblast functions. Particularly, some miRNAs are specifically or preferentially expressed in the placenta, such as C19MC and C14MC clusters [98], which are located in chromosome 19q13.41 and 14q32, respectively. C19MC is the largest miRNA cluster identified to date, and C14MC is encoded by maternally imprinted genes. Moreover, a recent study has pointed out that C19MC also plays a key role in human placental morphogenesis [99]. Same as the miRNA clusters of C19MC and C14MC, the expression levels of some miRNAs exhibit variability within the different stages of human placental development. A previous study reveals that the expression of C19MC miRNAs in trophoblast cells is increased significantly from the first trimester to the third trimester, while intriguingly, C14MC miRNA expression presents the opposite pattern [76]. Similarly, the expression levels of the miR-371-3 cluster that is adjacent to C19MC are found to be elevated from the first to third trimester. In addition to the clusters, other miRNAs also exhibit variations in different gestational stages, such as miR-378a-5p, miR-376c and miR-371-3 [100,101,102], indicating that different miRNAs are regulated precisely and developmentally with stage-specific functions during human pregnancy.
Recent studies have clarified that kinds of miRNAs exert effects on human trophoblast cell proliferation, apoptosis, migration and invasion, suggesting human placental development is controlled by various miRNAs. In an in vitro study using the first trimester placental explants, small interfering RNA (siRNA)-mediated knockdown of Dicer significantly promotes cytotrophoblast cell proliferation and also significantly induces the expression of two molecules in pro-mitogenic signaling within cytotrophoblasts, including ERK and SHP-2 [103], indicating the role of Dicer-dependent miRNAs in regulating human early placental development. Likewise, some other miRNAs have been found to affect trophoblast cell proliferation and apoptosis. For example, miR-378a-5p promotes trophoblast cell survival by regulating Nodal [101]. Similarly, miR-376c elevates trophoblast cell proliferation by impairing transforming growth factor-β (TGF-β) and nodal signaling [100], miR-210-3p regulates trophoblast cell proliferation by fibroblast growth factor 1 (FGF1) [104], miR-133b boosts trophoblast cell proliferation by restricting SGK1 [105], and miR-518b enhances human trophoblast cell proliferation through reglating Rap1b and activating the Ras-MAPK pathway [106]. In contrast, miR-155 inhibits proliferation of human extra-villous trophoblast cells via the down-regulation of cyclin D1 expression [107], miR-424 suppresses trophoblast cell proliferation through regulating ERRγ [108], miR-184 promotes apoptosis of human trophoblast cells by targeting WIG1 [109], and miR-93 restrains trophoblast cell proliferation and induces trophoblast cell apoptosis by targeting BCL2L2 [110].
As stated above, movement (including migration and invasion) of EVTs towards the maternal decidua and myometrium are fundamental for effective placentation [111,112]. In EVTs of the human first trimester placenta, Dicer is observed to be abundantly expressed [103], suggesting miRNAs play a key role in trophoblast cell migration and invasion. A series of studies recently has uncovered a number of miRNAs that participate in regulating human trophoblast cell migration and invasion. Here are some examples: miR-34a targeting MYC inhibits human trophoblast cell invasion [113]; miR-27a suppresses trophoblast cell migration and invasion through mediating SMAD2 [114]; miR-320a negatively regulates trophoblast cell invasion via targeting estrogen-related receptor-gamma [115]; miR-431 dampens trophoblast cell migration and invasion by regulating ZEB1 [116]; miR-30a-3p modulates the invasive capacity in JEG-3 cells (a model for first trimester human trophoblasts) [117,118]; miR-7 negatively modulates human trophoblast migration and invasion by suppression of EMT [119]; miR-204 restrains human trophoblast cell invasion by targeting matrix metalloproteinase-9 (MMP-9) [120]; miR-135b down-regulates HTR-8/SVneo cell (a human trophoblastic cell line) invasion by directly inhibiting CXCL12 [121,122]; miR-145-5p inhibits HTR-8/SVneo cell invasion by down-regulation of Cyr61 [123]. In contrast, miR-378-5p enhances trophoblast cell migration and invasion via modulating Nodal [101], miR-3935 promotes human trophoblast cell migration and invasion through up-regulating EMT [124], and miR-302a potentiates trophoblast cell migration by regulating the expression levels of VEGFA [125]. Therefore, miRNAs expressed in human placenta exert both positive and negative effects on migration and invasion in trophoblast cells by regulating the expression of downstream molecules and the activity of different signaling pathways. Despite the small fraction of miRNAs that have already been characterized and investigated in migration and invasion in human trophoblast cells, further work is needed to make deeper explorations and gain a comprehensive understanding of miRNAs in regulating these cellular processes.

6. Micro-RNAs and Human Gestational Disorders

Aberrant expression levels of miRNAs have been both in vitro and in vivo detected in kinds of gestational complications, such as PE, FGR, GDM, and a number of studies have demonstrated the dysregulation of miRNA expression in compromised human pregnancies (summarized in Table 2).

6.1. Preeclampsia (PE)

PE, characterized by maternal hypertension and proteinuria, has been recognized as one of the most feared complications during human pregnancy, which negatively affects both the mother and fetus [126,127]. Many recent studies have provided the evidence that miRNAs participate in the regulation of physiological function and pathological development of placenta in pregnant women with PE. The expression of miRNAs varies between the placentas from women with PE and normal placentas and the expression of a variety of miRNAs have been observed to be increased or decreased in placentas of PE in clinic [128,129,130]. Particularly, aberrant expression of those miRNAs in human trophoblast cells gives rise to their arrested proliferation and inadequate invasion, which further lead to the failure in sufficient remodeling of maternal spiral arteries and possibly consequent deficiency in angiogenesis [74]. The miRNAs with significantly up-regulated expression in PE placenta, including miR-17 [68], miR-155 [107], miR-431 [116], miR-30a-3p [117], miR-206 [129], miR-210 [131], miR-20a [132], miR-20b [133], miR-29b [134], miR-16 [135], miR-675 [136], miR-141 [137], miR-142-3p [138], miR-125b [139], miR-200p-3b [140], miR-137 [141], miR-202-3p [142], miR-146a [143], miR-155 [144], miR-517a/b [145] and miR-517c [145], have been proved to negatively regulate trophoblast cell proliferation, migration, invasion and apoptosis. The effect of miRNAs on trophoblast cell is commonly dependent on the regulation of their corresponding target genes. For example, the up-regulated expression of miR-17 is associated with PE, and miR-17 impedes trophoblast cell function by targeting EPHB4 and ephrin-B2, which have been shown to be critical for human early placental development [68]. Similarly, in most cases, miRNA-induced alterations in trophoblast cell function such as growth and mobility rest with the expression mediation of related genes. For example, miR-125b inhibits migration and invasion of extra-villous trophoblastic cells by suppressing its target STAT3 expression, while STAT3 plays a key role in cell infiltration and vascular proliferation [139]. In addition, the up-regulated expression of miR-17~92 and miR-106a~363 clusters are demonstrated to inhibit trophoblast differentiation [146]. On the flip side, a variety of miRNAs with down-regulated expression, including miR-378a-5p [101], miR-376c [147], miR-195 [148], miR-335 [149], miR-126 [150] and miR-125a [151], exert positive effects on proliferation, migration, invasion and differentiation in human trophoblast cells, and contribute to anti-apoptosis effects. More recently, we have revealed that miR-3935, whose expression is significantly decreased in preeclamptic placental tissues, promotes human trophoblast cell EMT through regulating TRAF6-RGS2 signaling [124]. More recently, it has been suggested that miRNAs are selectively packaged into placental-derived exosomes (a kind of extracellular vesicles) and that miRNA species are differentially expressed in exosomes from women with PE and those from controls with normal pregnancy [152,153], indicating that placental exosomes may serve as mediators in the pathogenesis of PE and as early biomarkers of PE diagnosis. Given that placenta-derived exosomes participate in communication between maternal and fetal cells, it is thereby supposed that those exosomal microRNAs act as mediators in cellular biological processes including proliferation, apoptosis, invasion and migration, which have been identified to be dysregulated in PE. For example, the expression of miR-153-3p, miR-653-5p, and miR-325 is decreased in exosomes of a PE group [152], whereas the levels of miR-222-3p, miR-486-1-5p, miR-486-2-5p, miR-155, miR-136, miR-494 and miR-495 are up-regulated in exosomes from PE patients [152,153,154,155]. In addition to the expression regulation, miRNA SNPs have been shown to be associated with PE susceptibility, such as the polymorphism of miR-152 rs12940701, [156], miR-146a rs2910164 [157], pri-miR-26a1 rs7372209 [158], miR-155 rs767649 [159] and miR-146a rs2910164 [160] genes.
Table 2. Summary of aberrant expression of miRNA in human gestational disorders.
Table 2. Summary of aberrant expression of miRNA in human gestational disorders.
Disorder NameMiRNAs (Reference)
Pre-eclampsia
(PE)
Up-regulatedproliferation ↓miR-155 [107], miR-20a [132], miR-16 [135], miR-675 [136], miR-142-3p [138], miR-200p-3b [140], miR-137 [141], miR-146a [143]
migration ↓miR-155 [107], miR-431 [116], miR-20b [133], miR-16 [135], miR-125b [139], miR-200p-3b [140], miR-137 [141], miR-146a [143]
invasion ↓miR-431 [116], miR-30a-3p [117], miR-20a [132], miR-20b [133], miR-141 [137], miR-142-3p [138], miR-125b [139], miR-146a [143], miR-517a/b [145], miR-517c [145]
apoptosis ↑miR-30a-3p [117], miR-29b [134], miR-16 [135], miR-200p-3b [140]
differentiation ↓miR-17~92 clusters [146], miR-106a~363 clusters [146]
other/targets ↓miR-17 [68], miR-206 [129], miR-210 [131], miR-202-3p [142], miR-155 [144]
in exosomesmiR-222-3p [152], miR-486-1-5p [153], miR-486-2-5p [153], miR-155 [154], miR-136 [155], miR-494 [155], miR-495 [155]
Down-regulatedproliferation ↑miR-378a-5p [101], miR-376c [147], miR-335 [149], miR-126 [150]
migration ↑miR-378a-5p [101], miR-3935 [124], miR-376c [147], miR-126 [150],
invasion ↑miR-378a-5p [101], miR-3935 [124]
apoptosis ↓miR-335 [149], miR-125a [151]
differentiation ↑miR-126 [150]
other/targets ↑miR-195 [148]
in exosomesmiR-153-3p [152], miR-653-5p [152], miR-325 [152]
Fetal Growth Restriction
(FGR)
Up-regulatedoxidative stress ↓miR-199a-5p [161]
other/targets ↓miR-424 [162], miR-519a [163], miR-1323 [164], miR-516b [164], miR-515-5p [164], miR-520h [164], miR-519d [164], miR-526b [164]
Down-regulatedother/targets ↑miR-518b [163], miR-16 [165], miR-21 [165]
Gestational Diabetes Mellitus
(GDM)
Up-regulatedother/targets ↓miR-98 [166], miR-518d [167], miR-144 [168]
in exosomesmiR-122-5p [26], miR-132-3p [26],
miR-1323 [26], miR-136-5p [26], miR-182-3p [26],
miR-210-3p [26], miR-29a-3p [26], miR-29b-3p [26],
miR-342-3p [26], miR-520h [26], miR-330-3p [169]
Down-regulatedproliferation ↑miR-296 [170], miR-96 [171], miR-137 [172]
migration ↑miR-296 [170]
invasion ↑miR-296 [170]
in exosomesmiR-125b [168], miR-516-5p [173], miR-517-3p [173], miR-518-5p [173], miR-222-3p [173], miR-16-5p [173]
Early Recurrent MiscarriageUp-regulatedapoptosis ↑miR-365 [174], miR-149 [175]
other/targets ↓miR-4497 [176]
Down-regulatedproliferation ↑miR-155-5p [177]
apoptosis ↓miR-155-5p [177]
Preterm Birth
(PTB)
Down-regulatedother/targets ↑miR-338 [178], miR-29b-3p [179]
Small-for-Gestational-Age (SGA)Down-regulatedother/targets ↑miR-21 [165], miR-16 [165]

6.2. Fetal Growth Restriction (FGR)

Aberrant expression of miRNAs has been also observed and verified in other gestational disorders, including FGR. FGR is defined as the failure of a fetus to reach to growth potential and the occurrence of FGR is closely associated with multiple factors, such as maternal factors, fetal factors, placental factors as well as external factors [180]. The previous studies have revealed that the increased levels of some miRNAs are associated with the pathogenesis of FGR, such as miR-424 [162], miR-199a-5p [161], miR-519a [163], miR-515-5p, miR-1323, miR-520h, miR-519d, miR-516b and miR-526b [164], whereas the expression of some miRNAs is markedly reduced in infants with FGR, including miR-518b [163], miR-16 and miR-21 [165]. Additionally, miRNA SNPs, such as miR-146a G/C rs2910164 variation [181], exert inhibitory effects on fetal growth. However, a previous report showed that there is no significant difference in exosomal miRNA expression between women with FRG and controls [152].

6.3. Gestational Diabetes Mellitus (GDM)

GDM is characterized as glucose intolerance (any degree) with onset or first recognition during human pregnancy, which has emerged as one of the complications with high incidence during human pregnancy [182]. Several studies have demonstrated that miRNA expression is also associated with GDM occurrence, and aberrant expression of miRNA in human placenta gives rise to the pathogenesis of GDM, such as miR-296, whose expression is improperly down-regulated in placenta with GDM in clinic and affects human placental development through the alterations in migration and invasion of trophoblast cells by regulating HIF3A [170]. Similarly, the expression of miR-96 is observed to be suppressed in GDM samples and the down-regulated miR-96 exerts protective effects on pancreatic β-cell function via regulating PAK1 in GDM [171]. In addition, high glucose inhibits HTR-8/SVneo cell viability and proliferation through regulation of miR-137 expression, which in turn modulates the downstream PRKAA1/IL6 axis [122,172]. Conversely, up-regulation expression of miR-98 in human placental tissues with GDM is closely associated with the global DNA methylation status through targeting MECP2 [166], and the increased expression levels of miR-518d in the placentas of females with GDM are negatively correlated with the expression levels of PPARα protein [167]. Similarly, it has been found that the expression of exosomal miRNAs is also altered between women with GDM and normal pregnancy, suggesting the changes in the expression of certain miRNAs could be used for diagnosing GDM [183]. For example, miR-125b and miR-144 are always dysfunctional in the circulating exosomes and placenta of women with GDM [168], and the expression of miR-516-5p, miR-517-3p, miR-518-5p, miR-222-3p and miR-16-5p is down-regulated in exosomes from patients with GDM, particularly at the third trimester of gestation [173]. In contrast, the expression of miRNAs, such as miR-330-3p [169], miR-122-5p, miR-132-3p, miR-1323, miR-136-5p, miR-182-3p, miR-210-3p, miR-29a-3p, miR-29b-3p, miR-342-3p and miR-520h [26], is elevated in exosomes from GDM patients. Intriguingly, recent studies have also indicated that miR-binding SNPs such as SLC30A8 rs2466293 and INSR rs1366600 increase GDM susceptibility [184,185]. Similarly, Wang et al. reported that SNPs loci rs2466293, rs1366600, and rs1063192 located in the miR-181a-2-3p, miR-106, and miR-323b-5p binding regions, respectively, all increase the risk of GDM [185,186], whereas the C allele of miR-27a rs895819 could reduce the risk of GDM [187].

6.4. Other Disorders

In addition to PE, FGR and GDM, miRNAs such as miR-365, miR-4497, miR-155-5p and miR-149 are found to be associated with early recurrent miscarriage [174,175,176,177]. Moreover, in placenta of women with PTB, the expression of miRNAs is found to be altered, such as miR-338 [178] and miR-29b-3p [179], and interestingly, a comparison analysis study of the exosomal miRNA profile between term group and PTB group revealed a total of 173 miRNAs with significant alterations across gestation [188,189], indicating the exosomal miRNA expression pattern may predict the risk in PTB. Furthermore, the down-regulated expression of miR-21 and miR-16 in human placenta is shown to be predictive of SGA status [165], and a latest study reported that SGA children have a distinct exosomal miRNA expression profile compared with appropriate-for-gestational-age children [190], also providing the evidence that exosomal miRNAs may serve as novel biomarkers for SGA. On the other hand, SNPs recently have also been shown to be associated with these disorders. For example, polymorphisms in miRNAs such as miR-27a, miR-449b, miR-150 and miR-1179 are associated with recurrent miscarriage [191,192], while rs1537515 located in the binding sites of miR-1224-3p and miR-3150-5p increases the PTB susceptibility [193], and similarly, miR-1304-3p rs1537516 increases the risk of PTB [193]. Moreover, Silva et al. recently reported that reduced miR-146a rs2910164 presents increased frequency in SGA children [194].

7. Perspectives

Detection of specific miRNAs in different gestational stages and in the maternal circulation provides with the possibility that miRNAs may be used as pregnant-related biomarkers to monitor the status of human normal pregnancy and potential gestational diseases in clinic. However, quantification of miRNAs, particularly the exosome miRNAs, is challenging due to their small amount. Thus, it is important to improve current experimental methodologies and to hopefully make comparisons of the experimental data obtained with available databases on line as well as with in silico prediction models to achieve a complex view on the role of miRNAs in different gestational diseases and to further discover suitable biomarkers. Owing to the great work in past decades, we have obtained a better understanding of miRNAs in human pregnancy. Nevertheless, despite the exciting results on miRNA research in human placenta, our current knowledge on how the thousands of miRNAs contribute to human placental development and function in pregnancy processes is still very limited. Studies in the future are in need to continue to search for novel miRNA candidates, to investigate the regulation pattern of those miRNAs expression and to further decipher the molecular mechanisms underlying their regulation in human placenta.

Author Contributions

Writing—original draft preparation, M.J. and C.T.; writing—original draft preparation, Q.X., J.L. and S.X.; writing—review and editing, C.T.; funding acquisition, M.J. and C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was funded by Zhejiang Province Public Welfare Technology Application Research Project (No. LGF19H040010) to M.J., and by Starting Research Foundation of the Children’s Hospital, Zhejiang University School of Medicine (No. 481), Foundation for the Top-Notch Youth Talent Cultivation Project of Independent Design Project of National Clinical Research Center for Child Health (No. Q21A0006), and National Natural Science Foundation of China (No. 31801207) to C.T.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bhatti, G.K.; Khullar, N.; Sidhu, I.S.; Navik, U.S.; Reddy, A.P.; Reddy, P.H.; Bhatti, J.S. Emerging role of non-coding RNA in health and disease. Metab. Brain Dis. 2021, 36, 1119–1134. [Google Scholar] [CrossRef] [PubMed]
  2. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 1993, 75, 843–854. [Google Scholar] [CrossRef]
  3. Reinhart, B.J.; Slack, F.J.; Basson, M.; Pasquinelli, A.E.; Bettinger, J.C.; Rougvie, A.E.; Horvitz, H.R.; Ruvkun, G. The 21-nucleotide let-7 RNA regulates developmental timing in Caenorhabditis elegans. Nature 2000, 403, 901–906. [Google Scholar] [CrossRef]
  4. Mori, M.A.; Ludwig, R.G.; Garcia-Martin, R.; Brandao, B.B.; Kahn, C.R. Extracellular miRNAs: From Biomarkers to Mediators of Physiology and Disease. Cell Metab. 2019, 30, 656–673. [Google Scholar] [CrossRef] [PubMed]
  5. Li, M.; Yu, B. Recent advances in the regulation of plant miRNA biogenesis. RNA Biol. 2021, 18, 2087–2096. [Google Scholar] [CrossRef]
  6. Do, D.N.; Dudemaine, P.L.; Mathur, M.; Suravajhala, P.; Zhao, X.; Ibeagha-Awemu, E.M. miRNA Regulatory Functions in Farm Animal Diseases, and Biomarker Potentials for Effective Therapies. Int. J. Mol. Sci. 2021, 22, 3080. [Google Scholar] [CrossRef]
  7. Avital, G.; Franca, G.S.; Yanai, I. Bimodal Evolutionary Developmental miRNA Program in Animal Embryogenesis. Mol. Biol. Evol. 2018, 35, 646–654. [Google Scholar] [CrossRef]
  8. Turco, M.Y.; Moffett, A. Development of the human placenta. Development 2019, 146, dev163428. [Google Scholar] [CrossRef]
  9. Hoffmann, A.; Spengler, D. The Mitochondrion as Potential Interface in Early-Life Stress Brain Programming. Front. Behav. Neurosci. 2018, 12, 306. [Google Scholar] [CrossRef]
  10. Tang, C.; Pan, Y.; Luo, H.; Xiong, W.; Zhu, H.; Ruan, H.; Wang, J.; Zou, C.; Tang, L.; Iguchi, T.; et al. Hedgehog signaling stimulates the conversion of cholesterol to steroids. Cell. Signal. 2015, 27, 487–497. [Google Scholar] [CrossRef] [Green Version]
  11. Cui, C.; Wu, C.; Wang, J.; Zheng, X.; Ma, Z.; Zhu, P.; Guan, W.; Zhang, S.; Chen, F. Leucine supplementation during late gestation globally alters placental metabolism and nutrient transport via modulation of the PI3K/AKT/mTOR signaling pathway in sows. Food Funct. 2022, 13, 2083–2097. [Google Scholar] [CrossRef] [PubMed]
  12. Soncin, F.; Parast, M.M. Role of Hippo signaling pathway in early placental development. Proc. Natl. Acad. Sci. USA 2020, 117, 20354–20356. [Google Scholar] [CrossRef] [PubMed]
  13. Hong, S.H.; Kim, S.C.; Park, M.N.; Jeong, J.S.; Yang, S.Y.; Lee, Y.J.; Bae, O.N.; Yang, H.S.; Seo, S.; Lee, K.S.; et al. Expression of steroidogenic enzymes in human placenta according to the gestational age. Mol. Med. Rep. 2019, 19, 3903–3911. [Google Scholar] [CrossRef] [PubMed]
  14. Walker, O.S.; Ragos, R.; Wong, M.K.; Adam, M.; Cheung, A.; Raha, S. Reactive oxygen species from mitochondria impacts trophoblast fusion and the production of endocrine hormones by syncytiotrophoblasts. PLoS ONE 2020, 15, e0229332. [Google Scholar] [CrossRef]
  15. Walker, O.S.; Ragos, R.; Gurm, H.; Lapierre, M.; May, L.L.; Raha, S. Delta-9-tetrahydrocannabinol disrupts mitochondrial function and attenuates syncytialization in human placental BeWo cells. Physiol. Rep. 2020, 8, e14476. [Google Scholar] [CrossRef]
  16. Zhu, H.; Ren, Q.; Yan, Z.; Xu, S.; Luo, J.; Wu, X.; Tang, C. Human HAND1 Inhibits the Conversion of Cholesterol to Steroids in Trophoblasts. J. Genet. Genom. 2021, 49, 350–363. [Google Scholar] [CrossRef]
  17. Ander, S.E.; Diamond, M.S.; Coyne, C.B. Immune responses at the maternal-fetal interface. Sci. Immunol. 2019, 4, eaat6114. [Google Scholar] [CrossRef]
  18. Jung, E.; Romero, R.; Yeo, L.; Gomez-Lopez, N.; Chaemsaithong, P.; Jaovisidha, A.; Gotsch, F.; Erez, O. The etiology of preeclampsia. Am. J. Obstet. Gynecol. 2022, 226, S844–S866. [Google Scholar] [CrossRef]
  19. Turbeville, H.R.; Sasser, J.M. Preeclampsia beyond pregnancy: Long-term consequences for mother and child. Am. J. Physiol. Renal Physiol. 2020, 318, F1315–F1326. [Google Scholar] [CrossRef]
  20. American College of Obstetricians and Gynecologists’ Committee on Practice Bulletins–Obstetrics and the Society for Maternal-Fetal Medicine. ACOG Practice Bulletin No. 204: Fetal Growth Restriction. Obstet. Gynecol. 2019, 133, e97–e109. [Google Scholar] [CrossRef]
  21. Wilcox, A.J.; Cortese, M.; McConnaughey, D.R.; Moster, D.; Basso, O. The limits of small-for-gestational-age as a high-risk category. Eur. J. Epidemiol. 2021, 36, 985–991. [Google Scholar] [CrossRef] [PubMed]
  22. Walani, S.R. Global burden of preterm birth. Int. J. Gynaecol. Obstet. 2020, 150, 31–33. [Google Scholar] [CrossRef] [PubMed]
  23. Szmuilowicz, E.D.; Josefson, J.L.; Metzger, B.E. Gestational Diabetes Mellitus. Endocrinol. Metab. Clin. N. Am. 2019, 48, 479–493. [Google Scholar] [CrossRef] [PubMed]
  24. Devall, A.J.; Papadopoulou, A.; Podesek, M.; Haas, D.M.; Price, M.J.; Coomarasamy, A.; Gallos, I.D. Progestogens for preventing miscarriage: A network meta-analysis. Cochrane Database Syst. Rev. 2021, 4, CD013792. [Google Scholar] [PubMed]
  25. Gonzalez, T.L.; Eisman, L.E.; Joshi, N.V.; Flowers, A.E.; Wu, D.; Wang, Y.; Santiskulvong, C.; Tang, J.; Buttle, R.A.; Sauro, E.; et al. High-throughput miRNA sequencing of the human placenta: Expression throughout gestation. Epigenomics 2021, 13, 995–1012. [Google Scholar] [CrossRef] [PubMed]
  26. Gillet, V.; Ouellet, A.; Stepanov, Y.; Rodosthenous, R.S.; Croft, E.K.; Brennan, K.; Abdelouahab, N.; Baccarelli, A.; Takser, L. miRNA Profiles in Extracellular Vesicles From Serum Early in Pregnancies Complicated by Gestational Diabetes Mellitus. J. Clin. Endocrinol. Metab. 2019, 104, 5157–5169. [Google Scholar] [CrossRef]
  27. Vaiman, D. Genes, epigenetics and miRNA regulation in the placenta. Placenta 2017, 52, 127–133. [Google Scholar] [CrossRef]
  28. Kozomara, A.; Birgaoanu, M.; Griffiths-Jones, S. miRBase: From microRNA sequences to function. Nucleic Acids Res. 2019, 47, D155–D162. [Google Scholar] [CrossRef]
  29. Yoshida, T.; Asano, Y.; Ui-Tei, K. Modulation of MicroRNA Processing by Dicer via Its Associated dsRNA Binding Proteins. Non-Coding RNA 2021, 7, 57. [Google Scholar] [CrossRef]
  30. Saliminejad, K.; Khorram Khorshid, H.R.; Soleymani Fard, S.; Ghaffari, S.H. An overview of microRNAs: Biology, functions, therapeutics, and analysis methods. J. Cell. Physiol. 2019, 234, 5451–5465. [Google Scholar] [CrossRef]
  31. Rani, V.; Sengar, R.S. Biogenesis and mechanisms of microRNA-mediated gene regulation. Biotechnol. Bioeng. 2022, 119, 685–692. [Google Scholar] [CrossRef] [PubMed]
  32. Parker, M.P.; Peterson, K.R.; Slawson, C. O-GlcNAcylation and O-GlcNAc Cycling Regulate Gene Transcription: Emerging Roles in Cancer. Cancers 2021, 13, 1666. [Google Scholar] [CrossRef] [PubMed]
  33. Sheng, P.; Fields, C.; Aadland, K.; Wei, T.; Kolaczkowski, O.; Gu, T.; Kolaczkowski, B.; Xie, M. Dicer cleaves 5′-extended microRNA precursors originating from RNA polymerase II transcription start sites. Nucleic Acids Res. 2018, 46, 5737–5752. [Google Scholar] [CrossRef] [PubMed]
  34. Lauressergues, D.; Couzigou, J.M.; Clemente, H.S.; Martinez, Y.; Dunand, C.; Becard, G.; Combier, J.P. Primary transcripts of microRNAs encode regulatory peptides. Nature 2015, 520, 90–93. [Google Scholar] [CrossRef]
  35. Roshandel, E.; Noorazar, L.; Farhadihosseinabadi, B.; Mehdizadeh, M.; Kazemi, M.H.; Parkhideh, S. PI3 kinase signaling pathway in hematopoietic cancers: A glance in miRNA’s role. J. Clin. Lab. Anal. 2021, 35, e23725. [Google Scholar] [CrossRef] [PubMed]
  36. Mizutani, A.; Seimiya, H. Tankyrase promotes primary precursor miRNA processing to precursor miRNA. Biochem. Biophys. Res. Commun. 2020, 522, 945–951. [Google Scholar] [CrossRef] [PubMed]
  37. Kian, R.; Moradi, S.; Ghorbian, S. Role of components of microRNA machinery in carcinogenesis. Exp. Oncol. 2018, 40, 2–9. [Google Scholar] [CrossRef]
  38. Michlewski, G.; Caceres, J.F. Post-transcriptional control of miRNA biogenesis. RNA 2019, 25, 1–16. [Google Scholar] [CrossRef]
  39. Clancy, J.W.; Zhang, Y.; Sheehan, C.; D’Souza-Schorey, C. An ARF6-Exportin-5 axis delivers pre-miRNA cargo to tumour microvesicles. Nat. Cell Biol. 2019, 21, 856–866. [Google Scholar] [CrossRef]
  40. Dandewad, V.; Vindu, A.; Joseph, J.; Seshadri, V. Import of human miRNA-RISC complex into Plasmodium falciparum and regulation of the parasite gene expression. J. Biosci. 2019, 44, 50. [Google Scholar] [CrossRef]
  41. Abedini, F.; Ebrahimi, M.; Hosseinkhani, H. Technology of RNA Interference in Advanced Medicine. Microrna 2018, 7, 74–84. [Google Scholar] [CrossRef] [PubMed]
  42. Gebert, L.F.R.; MacRae, I.J. Regulation of microRNA function in animals. Nat. Rev. Mol. Cell Biol. 2019, 20, 21–37. [Google Scholar] [CrossRef] [PubMed]
  43. Correia de Sousa, M.; Gjorgjieva, M.; Dolicka, D.; Sobolewski, C.; Foti, M. Deciphering miRNAs’ Action through miRNA Editing. Int. J. Mol. Sci. 2019, 20, 6249. [Google Scholar] [CrossRef] [PubMed]
  44. Ha, M.; Kim, V.N. Regulation of microRNA biogenesis. Nat. Rev. Mol. Cell Biol. 2014, 15, 509–524. [Google Scholar] [CrossRef]
  45. Kern, F.; Krammes, L.; Danz, K.; Diener, C.; Kehl, T.; Kuchler, O.; Fehlmann, T.; Kahraman, M.; Rheinheimer, S.; Aparicio-Puerta, E.; et al. Validation of human microRNA target pathways enables evaluation of target prediction tools. Nucleic Acids Res. 2021, 49, 127–144. [Google Scholar] [CrossRef]
  46. Roos, D.; de Boer, M. Mutations in cis that affect mRNA synthesis, processing and translation. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 166166. [Google Scholar] [CrossRef]
  47. Ezhilarasan, D. MicroRNA interplay between hepatic stellate cell quiescence and activation. Eur. J. Pharmacol. 2020, 885, 173507. [Google Scholar] [CrossRef]
  48. Chu, Y.W.; Chang, K.P.; Chen, C.W.; Liang, Y.T.; Soh, Z.T.; Hsieh, L.C. miRgo: Integrating various off-the-shelf tools for identification of microRNA-target interactions by heterogeneous features and a novel evaluation indicator. Sci. Rep. 2020, 10, 1466. [Google Scholar] [CrossRef]
  49. Moser, G.; Guettler, J.; Forstner, D.; Gauster, M. Maternal Platelets-Friend or Foe of the Human Placenta? Int. J. Mol. Sci. 2019, 20, 5639. [Google Scholar] [CrossRef]
  50. Knofler, M.; Haider, S.; Saleh, L.; Pollheimer, J.; Gamage, T.; James, J. Human placenta and trophoblast development: Key molecular mechanisms and model systems. Cell. Mol. Life Sci. 2019, 76, 3479–3496. [Google Scholar] [CrossRef] [Green Version]
  51. Zhai, J.; Xiao, Z.; Wang, Y.; Wang, H. Human embryonic development: From peri-implantation to gastrulation. Trends Cell Biol. 2022, 32, 18–29. [Google Scholar] [CrossRef] [PubMed]
  52. Roberts, R.M.; Ezashi, T.; Schulz, L.C.; Sugimoto, J.; Schust, D.J.; Khan, T.; Zhou, J. Syncytins expressed in human placental trophoblast. Placenta 2021, 113, 8–14. [Google Scholar] [CrossRef] [PubMed]
  53. Popovic, M.; Chuva de Sousa Lopes, S.M. Emerging in vitro platforms and omics technologies for studying the endometrium and early embryo-maternal interface in humans. Placenta 2022, 125, 36–46. [Google Scholar] [CrossRef]
  54. Burton, G.J.; Jauniaux, E. Placentation in the Human and Higher Primates. Adv. Anat. Embryol. Cell Biol. 2021, 234, 223–254. [Google Scholar] [PubMed]
  55. Aplin, J.D.; Jones, C.J.P. Cell dynamics in human villous trophoblast. Hum. Reprod. Update 2021, 27, 904–922. [Google Scholar] [CrossRef]
  56. Sheridan, M.A.; Zhao, X.; Fernando, R.C.; Gardner, L.; Perez-Garcia, V.; Li, Q.; Marsh, S.G.E.; Hamilton, R.; Moffett, A.; Turco, M.Y. Characterization of primary models of human trophoblast. Development 2021, 148, dev199749. [Google Scholar] [CrossRef] [PubMed]
  57. Spiliopoulos, M.; Kuo, C.Y.; Eranki, A.; Jacobs, M.; Rossi, C.T.; Iqbal, S.N.; Fisher, J.P.; Fries, M.H.; Kim, P.C.W. Characterizing placental stiffness using ultrasound shear-wave elastography in healthy and preeclamptic pregnancies. Arch. Gynecol. Obstet. 2020, 302, 1103–1112. [Google Scholar] [CrossRef]
  58. Stenhouse, C.; Seo, H.; Wu, G.; Johnson, G.A.; Bazer, F.W. Insights into the Regulation of Implantation and Placentation in Humans, Rodents, Sheep, and Pigs. Adv. Exp. Med. Biol. 2022, 1354, 25–48. [Google Scholar]
  59. Cattini, P.A.; Jin, Y.; Jarmasz, J.S.; Noorjahan, N.; Bock, M.E. Obesity and regulation of human placental lactogen production in pregnancy. J. Neuroendocrinol. 2020, 32, e12859. [Google Scholar] [CrossRef]
  60. Xu, L.; Li, Y.; Sang, Y.; Li, D.J.; Du, M. Crosstalk Between Trophoblasts and Decidual Immune Cells: The Cornerstone of Maternal-Fetal Immunotolerance. Front. Immunol. 2021, 12, 642392. [Google Scholar] [CrossRef]
  61. Treissman, J.; Yuan, V.; Baltayeva, J.; Le, H.T.; Castellana, B.; Robinson, W.P.; Beristain, A.G. Low oxygen enhances trophoblast column growth by potentiating differentiation of the extravillous lineage and promoting LOX activity. Development 2020, 147, dev181263. [Google Scholar] [CrossRef] [PubMed]
  62. James, J.L.; Boss, A.L.; Sun, C.; Allerkamp, H.H.; Clark, A.R. From stem cells to spiral arteries: A journey through early placental development. Placenta 2022, 125, 68–77. [Google Scholar] [CrossRef]
  63. Albrecht, E.D.; Pepe, G.J. Regulation of Uterine Spiral Artery Remodeling: A Review. Reprod. Sci. 2020, 27, 1932–1942. [Google Scholar] [CrossRef]
  64. Haram, K.; Mortensen, J.H.; Myking, O.; Roald, B.; Magann, E.F.; Morrison, J.C. Early development of the human placenta and pregnancy complications. J. Matern. Fetal Neonatal. Med. 2020, 33, 3538–3545. [Google Scholar] [CrossRef] [PubMed]
  65. Mouillet, J.F.; Chu, T.; Nelson, D.M.; Mishima, T.; Sadovsky, Y. MiR-205 silences MED1 in hypoxic primary human trophoblasts. FASEB J. 2010, 24, 2030–2039. [Google Scholar] [CrossRef] [PubMed]
  66. Hornakova, A.; Kolkova, Z.; Holubekova, V.; Loderer, D.; Lasabova, Z.; Biringer, K.; Halasova, E. Diagnostic Potential of MicroRNAs as Biomarkers in the Detection of Preeclampsia. Genet. Test. Mol. Biomark. 2020, 24, 321–327. [Google Scholar] [CrossRef] [PubMed]
  67. Gunel, T.; Hosseini, M.K.; Gumusoglu, E.; Kisakesen, H.I.; Benian, A.; Aydinli, K. Expression profiling of maternal plasma and placenta microRNAs in preeclamptic pregnancies by microarray technology. Placenta 2017, 52, 77–85. [Google Scholar] [CrossRef]
  68. Wang, W.; Feng, L.; Zhang, H.; Hachy, S.; Satohisa, S.; Laurent, L.C.; Parast, M.; Zheng, J.; Chen, D.B. Preeclampsia up-regulates angiogenesis-associated microRNA (i.e., miR-17, -20a, and -20b) that target ephrin-B2 and EPHB4 in human placenta. J. Clin. Endocrinol. Metab. 2012, 97, E1051–E1059. [Google Scholar] [CrossRef]
  69. Stribling, D.; Lei, Y.; Guardia, C.M.; Li, L.; Fields, C.J.; Nowialis, P.; Opavsky, R.; Renne, R.; Xie, M. A noncanonical microRNA derived from the snaR-A noncoding RNA targets a metastasis inhibitor. RNA 2021, 27, 694–709. [Google Scholar] [CrossRef]
  70. Vashukova, E.S.; Glotov, A.S.; Fedotov, P.V.; Efimova, O.A.; Pakin, V.S.; Mozgovaya, E.V.; Pendina, A.A.; Tikhonov, A.V.; Koltsova, A.S.; Baranov, V.S. Placental microRNA expression in pregnancies complicated by superimposed preeclampsia on chronic hypertension. Mol. Med. Rep. 2016, 14, 22–32. [Google Scholar] [CrossRef]
  71. Pomin, V.H.; Wang, X. Synthetic Oligosaccharide Libraries and Microarray Technology: A Powerful Combination for the Success of Current Glycosaminoglycan Interactomics. ChemMedChem 2018, 13, 648–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Yaylak, B.; Akgul, B. Experimental MicroRNA Detection Methods. Methods Mol. Biol. 2022, 2257, 33–55. [Google Scholar] [PubMed]
  73. Huang, Y.; Ren, H.T.; Xiong, J.L.; Gao, X.C.; Sun, X.H. Identification and characterization of known and novel microRNAs in three tissues of Chinese giant salamander base on deep sequencing approach. Genomics 2017, 109, 258–264. [Google Scholar] [CrossRef] [PubMed]
  74. Hayder, H.; O’Brien, J.; Nadeem, U.; Peng, C. MicroRNAs: Crucial regulators of placental development. Reproduction 2018, 155, R259–R271. [Google Scholar] [CrossRef]
  75. Morales-Prieto, D.M.; Ospina-Prieto, S.; Schmidt, A.; Chaiwangyen, W.; Markert, U.R. Elsevier Trophoblast Research Award Lecture: Origin, evolution and future of placenta miRNAs. Placenta 2014, 35, S39–S45. [Google Scholar] [CrossRef]
  76. Morales-Prieto, D.M.; Chaiwangyen, W.; Ospina-Prieto, S.; Schneider, U.; Herrmann, J.; Gruhn, B.; Markert, U.R. MicroRNA expression profiles of trophoblastic cells. Placenta 2012, 33, 725–734. [Google Scholar] [CrossRef]
  77. Chen, P.S.; Chiu, W.T.; Hsu, P.L.; Lin, S.C.; Peng, I.C.; Wang, C.Y.; Tsai, S.J. Pathophysiological implications of hypoxia in human diseases. J. Biomed. Sci. 2020, 27, 63. [Google Scholar] [CrossRef]
  78. Hu, X.Q.; Zhang, L. Hypoxia and Mitochondrial Dysfunction in Pregnancy Complications. Antioxidants 2021, 10, 405. [Google Scholar] [CrossRef]
  79. Maslen, C.L. Recent Advances in Placenta-Heart Interactions. Front. Physiol. 2018, 9, 735. [Google Scholar] [CrossRef]
  80. Burton, G.J.; Cindrova-Davies, T.; Yung, H.W.; Jauniaux, E. Hypoxia and Reproductive Health: Oxygen and development of the human placenta. Reproduction 2021, 161, F53–F65. [Google Scholar] [CrossRef]
  81. Jaszczuk, I.; Koczkodaj, D.; Kondracka, A.; Kwasniewska, A.; Winkler, I.; Filip, A. The role of miRNA-210 in pre-eclampsia development. Ann. Med. 2022, 54, 1350–1356. [Google Scholar] [CrossRef] [PubMed]
  82. Wu, D.; Chen, X.; Wang, L.; Chen, F.; Cen, H.; Shi, L. Hypoxia-induced microRNA-141 regulates trophoblast apoptosis, invasion, and vascularization by blocking CXCL12beta/CXCR2/4 signal transduction. Biomed. Pharmacother. 2019, 116, 108836. [Google Scholar] [CrossRef] [PubMed]
  83. Fang, M.; Du, H.; Han, B.; Xia, G.; Shi, X.; Zhang, F.; Fu, Q.; Zhang, T. Hypoxia-inducible microRNA-218 inhibits trophoblast invasion by targeting LASP1: Implications for preeclampsia development. Int. J. Biochem. Cell Biol. 2017, 87, 95–103. [Google Scholar] [CrossRef] [PubMed]
  84. Addo, K.A.; Palakodety, N.; Hartwell, H.J.; Tingare, A.; Fry, R.C. Placental microRNAs: Responders to environmental chemicals and mediators of pathophysiology of the human placenta. Toxicol. Rep. 2020, 7, 1046–1056. [Google Scholar] [CrossRef] [PubMed]
  85. Ali, A.; Hadlich, F.; Abbas, M.W.; Iqbal, M.A.; Tesfaye, D.; Bouma, G.J.; Winger, Q.A.; Ponsuksili, S. MicroRNA-mRNA Networks in Pregnancy Complications: A Comprehensive Downstream Analysis of Potential Biomarkers. Int. J. Mol. Sci. 2021, 22, 2313. [Google Scholar] [CrossRef]
  86. Eaves, L.A.; Phookphan, P.; Rager, J.E.; Bangma, J.; Santos, H.P., Jr.; Smeester, L.; O’Shea, T.M.; Fry, R.C. A role for microRNAs in the epigenetic control of sexually dimorphic gene expression in the human placenta. Epigenomics 2020, 12, 1543–1558. [Google Scholar] [CrossRef]
  87. Zhong, J.; Baccarelli, A.A.; Mansur, A.; Adir, M.; Nahum, R.; Hauser, R.; Bollati, V.; Racowsky, C.; Machtinger, R. Maternal Phthalate and Personal Care Products Exposure Alters Extracellular Placental miRNA Profile in Twin Pregnancies. Reprod. Sci. 2019, 26, 289–294. [Google Scholar] [CrossRef]
  88. Meruvu, S.; Zhang, J.; Choudhury, M. Mono-(2-ethylhexyl) Phthalate Increases Oxidative Stress Responsive miRNAs in First Trimester Placental Cell Line HTR8/SVneo. Chem. Res. Toxicol. 2016, 29, 430–435. [Google Scholar] [CrossRef]
  89. Meruvu, S.; Zhang, J.; Bedi, Y.S.; Choudhury, M. Mono-(2-ethylhexyl) phthalate induces apoptosis through miR-16 in human first trimester placental cell line HTR-8/SVneo. Toxicol. In Vitro 2016, 31, 35–42. [Google Scholar] [CrossRef]
  90. Avissar-Whiting, M.; Veiga, K.R.; Uhl, K.M.; Maccani, M.A.; Gagne, L.A.; Moen, E.L.; Marsit, C.J. Bisphenol A exposure leads to specific microRNA alterations in placental cells. Reprod. Toxicol. 2010, 29, 401–406. [Google Scholar] [CrossRef]
  91. LaRocca, J.; Binder, A.M.; McElrath, T.F.; Michels, K.B. First-Trimester Urine Concentrations of Phthalate Metabolites and Phenols and Placenta miRNA Expression in a Cohort of U.S. Women. Environ. Health Perspect. 2016, 124, 380–387. [Google Scholar] [CrossRef] [PubMed]
  92. Huang, Q.; Liu, Y.; Dong, S. Emerging roles of long non-coding RNAs in the toxicology of environmental chemicals. J. Appl. Toxicol. 2018, 38, 934–943. [Google Scholar] [CrossRef] [PubMed]
  93. Yao, Q.; Chen, Y.; Zhou, X. The roles of microRNAs in epigenetic regulation. Curr. Opin. Chem. Biol. 2019, 51, 11–17. [Google Scholar] [CrossRef] [PubMed]
  94. Noguer-Dance, M.; Abu-Amero, S.; Al-Khtib, M.; Lefevre, A.; Coullin, P.; Moore, G.E.; Cavaille, J. The primate-specific microRNA gene cluster (C19MC) is imprinted in the placenta. Hum. Mol. Genet. 2010, 19, 3566–3582. [Google Scholar] [CrossRef] [PubMed]
  95. Prats-Puig, A.; Xargay-Torrent, S.; Carreras-Badosa, G.; Mas-Pares, B.; Bassols, J.; Petry, C.J.; Girardot, M.; Francis, D.E.Z.; Ibanez, L.; Dunger, D.B.; et al. Methylation of the C19MC microRNA locus in the placenta: Association with maternal and chilhood body size. Int. J. Obes. 2020, 44, 13–22. [Google Scholar] [CrossRef] [PubMed]
  96. Srivastava, P.; Bamba, C.; Chopra, S.; Mandal, K. Role of miRNA polymorphism in recurrent pregnancy loss: A systematic review and meta-analysis. Biomark Med. 2022, 16, 101–115. [Google Scholar] [CrossRef] [PubMed]
  97. Ezat, S.A.; Haji, A.I. Study of association between different microRNA variants and the risk of idiopathic recurrent pregnancy loss. Arch. Gynecol. Obstet. 2022, 306, 1281–1286. [Google Scholar] [CrossRef]
  98. Inno, R.; Kikas, T.; Lillepea, K.; Laan, M. Coordinated Expressional Landscape of the Human Placental miRNome and Transcriptome. Front. Cell Dev. Biol. 2021, 9, 697947. [Google Scholar] [CrossRef]
  99. Mouillet, J.F.; Goff, J.; Sadovsky, E.; Sun, H.; Parks, T.; Chu, T.; Sadovsky, Y. Transgenic expression of human C19MC miRNAs impacts placental morphogenesis. Placenta 2020, 101, 208–214. [Google Scholar] [CrossRef]
  100. Fu, G.; Ye, G.; Nadeem, L.; Ji, L.; Manchanda, T.; Wang, Y.; Zhao, Y.; Qiao, J.; Wang, Y.L.; Lye, S.; et al. MicroRNA-376c impairs transforming growth factor-beta and nodal signaling to promote trophoblast cell proliferation and invasion. Hypertension 2013, 61, 864–872. [Google Scholar] [CrossRef]
  101. Luo, L.; Ye, G.; Nadeem, L.; Fu, G.; Yang, B.B.; Honarparvar, E.; Dunk, C.; Lye, S.; Peng, C. MicroRNA-378a-5p promotes trophoblast cell survival, migration and invasion by targeting Nodal. J. Cell Sci. 2012, 125, 3124–3132. [Google Scholar]
  102. Gottlieb, A.; Flor, I.; Nimzyk, R.; Burchardt, L.; Helmke, B.; Langenbuch, M.; Spiekermann, M.; Feidicker, S.; Bullerdiek, J. The expression of miRNA encoded by C19MC and miR-371-3 strongly varies among individual placentas but does not differ between spontaneous and induced abortions. Protoplasma 2021, 258, 209–218. [Google Scholar] [CrossRef]
  103. Forbes, K.; Farrokhnia, F.; Aplin, J.D.; Westwood, M. Dicer-dependent miRNAs provide an endogenous restraint on cytotrophoblast proliferation. Placenta 2012, 33, 581–585. [Google Scholar] [CrossRef] [PubMed]
  104. Li, L.; Huang, X.; He, Z.; Xiong, Y.; Fang, Q. miRNA-210-3p regulates trophoblast proliferation and invasiveness through fibroblast growth factor 1 in selective intrauterine growth restriction. J. Cell. Mol. Med. 2019, 23, 4422–4433. [Google Scholar] [CrossRef]
  105. Wang, D.; Na, Q.; Song, G.Y.; Wang, L. Human umbilical cord mesenchymal stem cell-derived exosome-mediated transfer of microRNA-133b boosts trophoblast cell proliferation, migration and invasion in preeclampsia by restricting SGK1. Cell Cycle 2020, 19, 1869–1883. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, M.; Wang, Y.; Lu, H.; Wang, H.; Shi, X.; Shao, X.; Li, Y.X.; Zhao, Y.; Wang, Y.L. miR-518b Enhances Human Trophoblast Cell Proliferation Through Targeting Rap1b and Activating Ras-MAPK Signal. Front. Endocrinol. 2018, 9, 100. [Google Scholar] [CrossRef]
  107. Dai, Y.; Qiu, Z.; Diao, Z.; Shen, L.; Xue, P.; Sun, H.; Hu, Y. MicroRNA-155 inhibits proliferation and migration of human extravillous trophoblast derived HTR-8/SVneo cells via down-regulating cyclin D1. Placenta 2012, 33, 824–829. [Google Scholar] [CrossRef] [PubMed]
  108. Zou, Z.; He, Z.; Cai, J.; Huang, L.; Zhu, H.; Luo, Y. Potential role of microRNA-424 in regulating ERRgamma to suppress trophoblast proliferation and invasion in fetal growth restriction. Placenta 2019, 83, 57–62. [Google Scholar] [CrossRef]
  109. Zhang, Y.; Zhou, J.; Li, M.Q.; Xu, J.; Zhang, J.P.; Jin, L.P. MicroRNA-184 promotes apoptosis of trophoblast cells via targeting WIG1 and induces early spontaneous abortion. Cell Death Dis. 2019, 10, 223. [Google Scholar] [CrossRef]
  110. Liu, H.N.; Tang, X.M.; Wang, X.Q.; Gao, J.; Li, N.; Wang, Y.Y.; Xia, H.F. MiR-93 Inhibits Trophoblast Cell Proliferation and Promotes Cell Apoptosis by Targeting BCL2L2 in Recurrent Spontaneous Abortion. Reprod. Sci. 2020, 27, 152–162. [Google Scholar] [CrossRef]
  111. Matsubara, K.; Matsubara, Y.; Uchikura, Y.; Takagi, K.; Yano, A.; Sugiyama, T. HMGA1 Is a Potential Driver of Preeclampsia Pathogenesis by Interference with Extravillous Trophoblasts Invasion. Biomolecules 2021, 11, 822. [Google Scholar] [CrossRef] [PubMed]
  112. Nakashima, A.; Shima, T.; Aoki, A.; Kawaguchi, M.; Yasuda, I.; Tsuda, S.; Yoneda, S.; Yamaki-Ushijima, A.; Cheng, S.B.; Sharma, S.; et al. Molecular and immunological developments in placentas. Hum. Immunol. 2021, 82, 317–324. [Google Scholar] [CrossRef]
  113. Sun, M.; Chen, H.; Liu, J.; Tong, C.; Meng, T. MicroRNA-34a inhibits human trophoblast cell invasion by targeting MYC. BMC Cell Biol. 2015, 16, 21. [Google Scholar] [CrossRef] [PubMed]
  114. Zheng, W.; Chen, A.; Yang, H.; Hong, L. MicroRNA-27a inhibits trophoblast cell migration and invasion by targeting SMAD2: Potential role in preeclampsia. Exp. Ther. Med. 2020, 20, 2262–2269. [Google Scholar] [CrossRef] [PubMed]
  115. Gao, T.; Deng, M.; Wang, Q. MiRNA-320a inhibits trophoblast cell invasion by targeting estrogen-related receptor-gamma. J. Obstet. Gynaecol. Res. 2018, 44, 756–763. [Google Scholar] [CrossRef] [PubMed]
  116. Yang, X.; Meng, T. MicroRNA-431 affects trophoblast migration and invasion by targeting ZEB1 in preeclampsia. Gene 2019, 683, 225–232. [Google Scholar] [CrossRef]
  117. 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. Am. J. Obstet. Gynecol. 2018, 218, 249.e1–249.e12. [Google Scholar] [CrossRef]
  118. Mandl, M.; Haas, J.; Bischof, P.; Nohammer, G.; Desoye, G. Serum-dependent effects of IGF-I and insulin on proliferation and invasion of human first trimester trophoblast cell models. Histochem. Cell Biol. 2002, 117, 391–399. [Google Scholar] [CrossRef]
  119. Shih, J.C.; Lin, H.H.; Hsiao, A.C.; Su, Y.T.; Tsai, S.; Chien, C.L.; Kung, H.N. Unveiling the role of microRNA-7 in linking TGF-beta-Smad-mediated epithelial-mesenchymal transition with negative regulation of trophoblast invasion. FASEB J. 2019, 33, 6281–6295. [Google Scholar] [CrossRef]
  120. Yu, Y.; Wang, L.; Liu, T.; Guan, H. MicroRNA-204 suppresses trophoblast-like cell invasion by targeting matrix metalloproteinase-9. Biochem. Biophys. Res. Commun. 2015, 463, 285–291. [Google Scholar] [CrossRef]
  121. Tamaru, S.; Mizuno, Y.; Tochigi, H.; Kajihara, T.; Okazaki, Y.; Okagaki, R.; Kamei, Y.; Ishihara, O.; Itakura, A. MicroRNA-135b suppresses extravillous trophoblast-derived HTR-8/SVneo cell invasion by directly down regulating CXCL12 under low oxygen conditions. Biochem. Biophys. Res. Commun. 2015, 461, 421–426. [Google Scholar] [CrossRef] [PubMed]
  122. Han, D.; Jiang, L.; Gu, X.; Huang, S.; Pang, J.; Wu, Y.; Yin, J.; Wang, J. SIRT3 deficiency is resistant to autophagy-dependent ferroptosis by inhibiting the AMPK/mTOR pathway and promoting GPX4 levels. J. Cell. Physiol. 2020, 235, 8839–8851. [Google Scholar] [CrossRef] [PubMed]
  123. Wen, Z.; Chen, Y.; Long, Y.; Yu, J.; Li, M. Tumor necrosis factor-alpha suppresses the invasion of HTR-8/SVneo trophoblast cells through microRNA-145-5p-mediated downregulation of Cyr61. Life Sci. 2018, 209, 132–139. [Google Scholar] [CrossRef] [PubMed]
  124. Jin, M.; Xu, S.; Li, J.; Yao, Y.; Tang, C. MicroRNA-3935 promotes human trophoblast cell epithelial-mesenchymal transition through tumor necrosis factor receptor-associated factor 6/regulator of G protein signaling 2 axis. Reprod. Biol. Endocrinol. 2021, 19, 134. [Google Scholar] [CrossRef]
  125. Wu, M.; Zhao, Y.; Li, L.; Wang, G.; Xing, L. Exosomal microRNA302a promotes trophoblast migration and proliferation, and represses angiogenesis by regulating the expression levels of VEGFA in preeclampsia. Mol. Med. Rep. 2021, 24, 864. [Google Scholar] [CrossRef]
  126. Rana, S.; Lemoine, E.; Granger, J.P.; Karumanchi, S.A. Preeclampsia: Pathophysiology, Challenges, and Perspectives. Circ. Res. 2019, 124, 1094–1112. [Google Scholar] [CrossRef]
  127. Armaly, Z.; Jadaon, J.E.; Jabbour, A.; Abassi, Z.A. Preeclampsia: Novel Mechanisms and Potential Therapeutic Approaches. Front. Physiol. 2018, 9, 973. [Google Scholar] [CrossRef]
  128. Yang, H.L.; Zhang, H.Z.; Meng, F.R.; Han, S.Y.; Zhang, M. Differential expression of microRNA-411 and 376c is associated with hypertension in pregnancy. Braz. J. Med. Biol. Res. 2019, 52, e7546. [Google Scholar] [CrossRef]
  129. Akehurst, C.; Small, H.Y.; Sharafetdinova, L.; Forrest, R.; Beattie, W.; Brown, C.E.; Robinson, S.W.; McClure, J.D.; Work, L.M.; Carty, D.M.; et al. Differential expression of microRNA-206 and its target genes in preeclampsia. J. Hypertens. 2015, 33, 2068–2074. [Google Scholar] [CrossRef]
  130. Gunel, T.; Kamali, N.; Hosseini, M.K.; Gumusoglu, E.; Benian, A.; Aydinli, K. Regulatory effect of miR-195 in the placental dysfunction of preeclampsia. J. Matern. Fetal Neonatal Med. 2020, 33, 901–908. [Google Scholar] [CrossRef]
  131. Muralimanoharan, S.; Maloyan, A.; Mele, J.; Guo, C.; Myatt, L.G.; Myatt, L. MIR-210 modulates mitochondrial respiration in placenta with preeclampsia. Placenta 2012, 33, 816–823. [Google Scholar] [CrossRef]
  132. Wang, Y.; Zhang, Y.; Wang, H.; Wang, J.; Zhang, Y.; Wang, Y.; Pan, Z.; Luo, S. Aberrantly up-regulated miR-20a in pre-eclampsic placenta compromised the proliferative and invasive behaviors of trophoblast cells by targeting forkhead box protein A1. Int. J. Biol. Sci. 2014, 10, 973–982. [Google Scholar] [CrossRef] [PubMed]
  133. Jin, M.; Li, H.; Xu, H.; Huo, G.; Yao, Y. MicroRNA-20b inhibits trophoblast cell migration and invasion by targeting MMP-2. Int. J. Clin. Exp. Pathol. 2017, 10, 10901–10909. [Google Scholar] [PubMed]
  134. Du, J.; Ji, Q.; Dong, L.; Meng, Y.; Xin, G. HDAC4 Knockdown Induces Preeclampsia Cell Autophagy and Apoptosis by miR-29b. Reprod. Sci. 2021, 28, 334–342. [Google Scholar] [CrossRef]
  135. Yuan, Y.; Zhao, L.; Wang, X.; Lian, F.; Cai, Y. Ligustrazine-induced microRNA-16-5p inhibition alleviates preeclampsia through IGF-2. Reproduction 2020, 160, 905–917. [Google Scholar] [CrossRef]
  136. Beards, F.; Jones, L.E.; Charnock, J.; Forbes, K.; Harris, L.K. Placental Homing Peptide-microRNA Inhibitor Conjugates for Targeted Enhancement of Intrinsic Placental Growth Signaling. Theranostics 2017, 7, 2940–2955. [Google Scholar] [CrossRef]
  137. Ospina-Prieto, S.; Chaiwangyen, W.; Herrmann, J.; Groten, T.; Schleussner, E.; Markert, U.R.; Morales-Prieto, D.M. MicroRNA-141 is upregulated in preeclamptic placentae and regulates trophoblast invasion and intercellular communication. Transl. Res. 2016, 172, 61–72. [Google Scholar] [CrossRef] [PubMed]
  138. Mao, Y.; Hou, B.; Shan, L.; Sun, X.; Wang, L. Aberrantly up-regulated miR-142-3p inhibited the proliferation and invasion of trophoblast cells by regulating FOXM1. Placenta 2021, 104, 253–260. [Google Scholar] [CrossRef] [PubMed]
  139. Tang, J.; Wang, D.; Lu, J.; Zhou, X. MiR-125b participates in the occurrence of preeclampsia by regulating the migration and invasion of extravillous trophoblastic cells through STAT3 signaling pathway. J. Recept. Signal. Transduct. Res. 2021, 41, 202–208. [Google Scholar] [CrossRef]
  140. Liu, H.; Wang, X. MiR-200b-3p is upregulated in the placental tissues from patients with preeclampsia and promotes the development of preeclampsia via targeting profilin 2. Cell Cycle 2022, 21, 1945–1957. [Google Scholar] [CrossRef]
  141. Lu, T.M.; Lu, W.; Zhao, L.J. MicroRNA-137 Affects Proliferation and Migration of Placenta Trophoblast Cells in Preeclampsia by Targeting ERRalpha. Reprod. Sci. 2017, 24, 85–96. [Google Scholar] [CrossRef] [PubMed]
  142. Singh, K.; Williams, J., III; Brown, J.; Wang, E.T.; Lee, B.; Gonzalez, T.L.; Cui, J.; Goodarzi, M.O.; Pisarska, M.D. Up-regulation of microRNA-202-3p in first trimester placenta of pregnancies destined to develop severe preeclampsia, a pilot study. Pregnancy Hypertens. 2017, 10, 7–9. [Google Scholar] [CrossRef] [PubMed]
  143. Peng, P.; Song, H.; Xie, C.; Zheng, W.; Ma, H.; Xin, D.; Zhan, J.; Yuan, X.; Chen, A.; Tao, J.; et al. miR-146a-5p-mediated suppression on trophoblast cell progression and epithelial-mesenchymal transition in preeclampsia. Biol. Res. 2021, 54, 30. [Google Scholar] [CrossRef] [PubMed]
  144. Nizyaeva, N.V.; Kulikova, G.V.; Nagovitsyna, M.N.; Kan, N.E.; Prozorovskaya, K.N.; Shchegolev, A.I.; Sukhikh, G.T. Expression of MicroRNA-146a and MicroRNA-155 in Placental Villi in Early- and Late-Onset Preeclampsia. Bull. Exp. Biol. Med. 2017, 163, 394–399. [Google Scholar] [CrossRef] [PubMed]
  145. Anton, L.; Olarerin-George, A.O.; Hogenesch, J.B.; Elovitz, M.A. Placental expression of miR-517a/b and miR-517c contributes to trophoblast dysfunction and preeclampsia. PLoS ONE 2015, 10, e0122707. [Google Scholar] [CrossRef]
  146. Kumar, P.; Luo, Y.; Tudela, C.; Alexander, J.M.; Mendelson, C.R. The c-Myc-regulated microRNA-17~92 (miR-17~92) and miR-106a~363 clusters target hCYP19A1 and hGCM1 to inhibit human trophoblast differentiation. Mol. Cell Biol. 2013, 33, 1782–1796. [Google Scholar] [CrossRef]
  147. Baker, A.H.; Delles, C. Is microRNA-376c a biomarker or mediator of preeclampsia? Hypertension 2013, 61, 767–769. [Google Scholar] [CrossRef]
  148. Sandrim, V.C.; Eleuterio, N.; Pilan, E.; Tanus-Santos, J.E.; Fernandes, K.; Cavalli, R. Plasma levels of increased miR-195-5p correlates with the sFLT-1 levels in preeclampsia. Hypertens. Pregnancy 2016, 35, 150–158. [Google Scholar] [CrossRef]
  149. Jiang, L.; Li, M.; Gan, H.; Yang, L.; Lin, F.; Hu, M. miR-335 targets CRIM1 to promote the proliferation and inhibit the apoptosis of placental trophoblast cells in preeclamptic rats. Am. J. Transl. Res. 2021, 13, 4676–4685. [Google Scholar]
  150. Yan, T.; Liu, Y.; Cui, K.; Hu, B.; Wang, F.; Zou, L. MicroRNA-126 regulates EPCs function: Implications for a role of miR-126 in preeclampsia. J. Cell Biochem. 2013, 114, 2148–2159. [Google Scholar] [CrossRef]
  151. Gu, Y.; Meng, J.; Zuo, C.; Wang, S.; Li, H.; Zhao, S.; Huang, T.; Wang, X.; Yan, J. Downregulation of MicroRNA-125a in Placenta Accreta Spectrum Disorders Contributes Antiapoptosis of Implantation Site Intermediate Trophoblasts by Targeting MCL1. Reprod. Sci. 2019, 26, 1582–1589. [Google Scholar] [CrossRef]
  152. Li, H.; Ouyang, Y.; Sadovsky, E.; Parks, W.T.; Chu, T.; Sadovsky, Y. Unique microRNA Signals in Plasma Exosomes from Pregnancies Complicated by Preeclampsia. Hypertension 2020, 75, 762–771. [Google Scholar] [CrossRef] [PubMed]
  153. Salomon, C.; Guanzon, D.; Scholz-Romero, K.; Longo, S.; Correa, P.; Illanes, S.E.; Rice, G.E. Placental Exosomes as Early Biomarker of Preeclampsia: Potential Role of Exosomal MicroRNAs Across Gestation. J. Clin. Endocrinol. Metab. 2017, 102, 3182–3194. [Google Scholar] [CrossRef]
  154. Shen, L.; Li, Y.; Li, R.; Diao, Z.; Yany, M.; Wu, M.; Sun, H.; Yan, G.; Hu, Y. Placentaassociated serum exosomal miR155 derived from patients with preeclampsia inhibits eNOS expression in human umbilical vein endothelial cells. Int. J. Mol. Med. 2018, 41, 1731–1739. [Google Scholar]
  155. Motawi, T.M.K.; Sabry, D.; Maurice, N.W.; Rizk, S.M. Role of mesenchymal stem cells exosomes derived microRNAs; miR-136, miR-494 and miR-495 in pre-eclampsia diagnosis and evaluation. Arch. Biochem. Biophys. 2018, 659, 13–21. [Google Scholar] [CrossRef]
  156. Rokni, M.; Salimi, S.; Sohrabi, T.; Asghari, S.; Teimoori, B.; Saravani, M. Association between miRNA-152 polymorphism and risk of preeclampsia susceptibility. Arch. Gynecol. Obstet. 2019, 299, 475–480. [Google Scholar] [CrossRef]
  157. Maharaj, N.R.; Ramkaran, P.; Pillay, S.; Chuturgoon, A.A. MicroRNA-146a rs2910164 is associated with severe preeclampsia in Black South African women on HAART. BMC Genet. 2017, 18, 5. [Google Scholar] [CrossRef] [PubMed]
  158. Eskandari, F.; Rezaei, M.; Mohammadpour-Gharehbagh, A.; Teimoori, B.; Yaghmaei, M.; Narooei-Nejad, M.; Salimi, S. The association of pri-miRNA- 26a1 rs7372209 polymorphism and Preeclampsia susceptibility. Clin. Exp. Hypertens. 2019, 41, 268–273. [Google Scholar] [CrossRef] [PubMed]
  159. Ayoub, S.E.; Shaker, O.G.; Abdelwahed, M.Y.; Ahmed, N.A.; Abdelhameed, H.G.; Bosilah, A.H.; Mohammed, S.R. Association of MicroRNA-155rs767649 Polymorphism with Susceptibility to Preeclampsia. Int. J. Mol. Cell. Med. 2019, 8, 247–257. [Google Scholar]
  160. Abo-Elmatty, D.M.; Mehanna, E.T. MIR146A rs2910164 (G/C) Polymorphism is Associated with Incidence of Preeclampsia in Gestational Diabetes Patients. Biochem. Genet. 2019, 57, 222–233. [Google Scholar] [CrossRef]
  161. Meng, M.; Cheng, Y.K.Y.; Wu, L.; Chaemsaithong, P.; Leung, M.B.W.; Chim, S.S.C.; Sahota, D.S.; Li, W.; Poon, L.C.Y.; Wang, C.C.; et al. Whole genome miRNA profiling revealed miR-199a as potential placental pathogenesis of selective fetal growth restriction in monochorionic twin pregnancies. Placenta 2020, 92, 44–53. [Google Scholar] [CrossRef]
  162. Huang, L.; Shen, Z.; Xu, Q.; Huang, X.; Chen, Q.; Li, D. Increased levels of microRNA-424 are associated with the pathogenesis of fetal growth restriction. Placenta 2013, 34, 624–627. [Google Scholar] [CrossRef] [PubMed]
  163. Wang, D.; Na, Q.; Song, W.W.; Song, G.Y. Altered Expression of miR-518b and miR-519a in the placenta is associated with low fetal birth weight. Am. J. Perinatol. 2014, 31, 729–734. [Google Scholar] [CrossRef]
  164. Higashijima, A.; Miura, K.; Mishima, H.; Kinoshita, A.; Jo, O.; Abe, S.; Hasegawa, Y.; Miura, S.; Yamasaki, K.; Yoshida, A.; et al. Characterization of placenta-specific microRNAs in fetal growth restriction pregnancy. Prenat. Diagn. 2013, 33, 214–222. [Google Scholar] [CrossRef] [PubMed]
  165. Maccani, M.A.; Padbury, J.F.; Marsit, C.J. miR-16 and miR-21 expression in the placenta is associated with fetal growth. PLoS ONE 2011, 6, e21210. [Google Scholar] [CrossRef] [PubMed]
  166. Cao, J.L.; Zhang, L.; Li, J.; Tian, S.; Lv, X.D.; Wang, X.Q.; Su, X.; Li, Y.; Hu, Y.; Ma, X.; et al. Up-regulation of miR-98 and unraveling regulatory mechanisms in gestational diabetes mellitus. Sci. Rep. 2016, 6, 32268. [Google Scholar] [CrossRef] [PubMed]
  167. Zhao, C.; Zhang, T.; Shi, Z.; Ding, H.; Ling, X. MicroRNA-518d regulates PPARalpha protein expression in the placentas of females with gestational diabetes mellitus. Mol. Med. Rep. 2014, 9, 2085–2090. [Google Scholar] [CrossRef]
  168. Zhang, L.; Zhang, T.; Sun, D.; Cheng, G.; Ren, H.; Hong, H.; Chen, L.; Jiao, X.; Du, Y.; Zou, Y.; et al. Diagnostic value of dysregulated microribonucleic acids in the placenta and circulating exosomes in gestational diabetes mellitus. J. Diabetes Investig. 2021, 12, 1490–1500. [Google Scholar] [CrossRef]
  169. Sebastiani, G.; Guarino, E.; Grieco, G.E.; Formichi, C.; Delli Poggi, C.; Ceccarelli, E.; Dotta, F. Circulating microRNA (miRNA) Expression Profiling in Plasma of Patients with Gestational Diabetes Mellitus Reveals Upregulation of miRNA miR-330-3p. Front. Endocrinol. 2017, 8, 345. [Google Scholar] [CrossRef]
  170. Sun, D.G.; Tian, S.; Zhang, L.; Hu, Y.; Guan, C.Y.; Ma, X.; Xia, H.F. The miRNA-29b Is Downregulated in Placenta During Gestational Diabetes Mellitus and May Alter Placenta Development by Regulating Trophoblast Migration and Invasion Through a HIF3A-Dependent Mechanism. Front. Endocrinol. 2020, 11, 169. [Google Scholar] [CrossRef]
  171. Li, L.; Wang, S.; Li, H.; Wan, J.; Zhou, Q.; Zhou, Y.; Zhang, C. microRNA-96 protects pancreatic beta-cell function by targeting PAK1 in gestational diabetes mellitus. Biofactors 2018, 44, 539–547. [Google Scholar] [CrossRef]
  172. Peng, H.Y.; Li, M.Q.; Li, H.P. High glucose suppresses the viability and proliferation of HTR8/SVneo cells through regulation of the miR137/PRKAA1/IL6 axis. Int. J. Mol. Med. 2018, 42, 799–810. [Google Scholar] [PubMed]
  173. Herrera-Van Oostdam, A.S.; Toro-Ortiz, J.C.; Lopez, J.A.; Noyola, D.E.; Garcia-Lopez, D.A.; Duran-Figueroa, N.V.; Martinez-Martinez, E.; Portales-Perez, D.P.; Salgado-Bustamante, M.; López Hernández, Y. Placental exosomes isolated from urine of patients with gestational diabetes exhibit a differential profile expression of microRNAs across gestation. Int. J. Mol. Med. 2020, 46, 546–560. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, W.; Shen, W.W.; Cao, X.M.; Ding, W.Y.; Yan, L.P.; Gao, L.J.; Li, X.L.; Zhong, T.Y. Novel mechanism of miRNA-365-regulated trophoblast apoptosis in recurrent miscarriage. J. Cell. Mol. Med. 2017, 21, 2412–2425. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, P.; Chen, X.; Chang, Y.; Wang, Y.; Xu, X.; Guo, Y.; Cui, H. Inhibition of microRNA-149 protects against recurrent miscarriage through upregulating RUNX2 and activation of the PTEN/Akt signaling pathway. J. Obstet. Gynaecol. Res. 2020, 46, 2534–2546. [Google Scholar] [CrossRef]
  176. Tang, H.; Pan, L.; Xiong, Y.; Wang, L.; Cui, Y.; Liu, J.; Tang, L. Down-regulation of the Sp1 transcription factor by an increase of microRNA-4497 in human placenta is associated with early recurrent miscarriage. Reprod. Biol. Endocrinol. 2021, 19, 21. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, Q.; Tian, P.; Xu, H. MicroRNA-155-5p regulates survival of human decidua stromal cells through NF-kappaB in recurrent miscarriage. Reprod. Biol. 2021, 21, 100510. [Google Scholar] [CrossRef]
  178. Montenegro, D.; Romero, R.; Kim, S.S.; Tarca, A.L.; Draghici, S.; Kusanovic, J.P.; Kim, J.S.; Lee, D.C.; Erez, O.; Gotsch, F.; et al. Expression patterns of microRNAs in the chorioamniotic membranes: A role for microRNAs in human pregnancy and parturition. J. Pathol. 2009, 217, 113–121. [Google Scholar] [CrossRef] [PubMed]
  179. Marttila, S.; Rovio, S.; Mishra, P.P.; Seppala, I.; Lyytikainen, L.P.; Juonala, M.; Waldenberger, M.; Oksala, N.; Ala-Korpela, M.; Harville, E.; et al. Adulthood blood levels of hsa-miR-29b-3p associate with preterm birth and adult metabolic and cognitive health. Sci. Rep. 2021, 11, 9203. [Google Scholar] [CrossRef] [PubMed]
  180. Fu, G.; Brkic, J.; Hayder, H.; Peng, C. MicroRNAs in Human Placental Development and Pregnancy Complications. Int. J. Mol. Sci. 2013, 14, 5519–5544. [Google Scholar] [CrossRef]
  181. Naidoo, P.; Naidoo, R.N.; Ramkaran, P.; Muttoo, S.; Asharam, K.; Chuturgoon, A.A. Maternal miRNA-146a G/C rs2910164 variation, HIV/AIDS and nitrogen oxide pollution exposure collectively affects foetal growth. Hum. Exp. Toxicol. 2019, 38, 82–94. [Google Scholar] [CrossRef] [PubMed]
  182. Coustan, D.R. Gestational diabetes mellitus. Clin. Chem. 2013, 59, 1310–1321. [Google Scholar] [CrossRef]
  183. Yang, X.; Wu, N. MicroRNAs and Exosomal microRNAs May Be Possible Targets to Investigate in Gestational Diabetes Mellitus. Diabetes Metab. Syndr. Obes. 2022, 15, 321–330. [Google Scholar] [CrossRef]
  184. Zhang, Y.; Bai, R.; Liu, C.; Ma, C.; Chen, X.; Yang, J.; Sun, D. MicroRNA single-nucleotide polymorphisms and diabetes mellitus: A comprehensive review. Clin. Genet. 2019, 95, 451–461. [Google Scholar] [CrossRef]
  185. Wang, X.; Li, W.; Ma, L.; Ping, F.; Liu, J.; Wu, X.; Mao, J.; Wang, X.; Nie, M. Investigation of miRNA-binding site variants and risk of gestational diabetes mellitus in Chinese pregnant women. Acta Diabetol. 2017, 54, 309–316. [Google Scholar] [CrossRef]
  186. Wang, X.; Li, W.; Ma, L.; Gao, J.; Liu, J.; Ping, F.; Nie, M. Association study of the miRNA-binding site polymorphisms of CDKN2A/B genes with gestational diabetes mellitus susceptibility. Acta Diabetol. 2015, 52, 951–958. [Google Scholar] [CrossRef]
  187. Hakimian, M.; Ghorbian, S. Negative associations between the has-miR-27a and hsa-miR-125a gene variations and prostate cancer susceptibility. Mol. Biol. Rep. 2020, 47, 4209–4214. [Google Scholar] [CrossRef]
  188. Menon, R.; Debnath, C.; Lai, A.; Guanzon, D.; Bhaatnagar, S.; Kshetrapal, P.K.; Sheller-Miller, S.; Salomon, C.; Garbhini Study, T. Circulating Exosomal miRNA Profile During Term and Preterm Birth Pregnancies: A Longitudinal Study. Endocrinology 2019, 160, 249–275. [Google Scholar] [CrossRef]
  189. Menon, R.; Shahin, H. Extracellular vesicles in spontaneous preterm birth. Am. J. Reprod. Immunol. 2021, 85, e13353. [Google Scholar] [CrossRef]
  190. Jeong, H.R.; Han, J.A.; Kim, H.; Lee, H.J.; Shim, Y.S.; Kang, M.J.; Yoon, J.S.; Ryu, S.; Hwang, I.T. Exosomal miRNA Profile in Small-for-Gestational-Age Children: A Potential Biomarker for Catch-Up Growth. Genes 2022, 13, 938. [Google Scholar] [CrossRef]
  191. Rah, H.; Chung, K.W.; Ko, K.H.; Kim, E.S.; Kim, J.O.; Sakong, J.H.; Kim, J.H.; Lee, W.S.; Kim, N.K. miR-27a and miR-449b polymorphisms associated with a risk of idiopathic recurrent pregnancy loss. PLoS ONE 2017, 12, e0177160. [Google Scholar] [CrossRef] [PubMed]
  192. Park, H.S.; Kim, E.S.; Ahn, E.H.; Kim, J.O.; An, H.J.; Kim, J.H.; Lee, Y.; Lee, W.S.; Kim, Y.R.; Kim, N.K. The microRNApolymorphisms inmiR-150 and miR-1179 are associated with risk of idiopathic recurrent pregnancy loss. Reprod. Biomed. Online 2019, 39, 187–195. [Google Scholar] [CrossRef] [PubMed]
  193. Zhu, Q.; Chen, Y.; Dai, J.; Wang, B.; Liu, M.; Wang, Y.; Tao, J.; Li, H. Methylenetetrahydrofolate reductase polymorphisms at 3′-untranslated region are associated with susceptibility to preterm birth. Transl. Pediatr. 2015, 4, 57–62. [Google Scholar] [PubMed]
  194. Silva, L.R.; Melo, A.S.; Salomao, K.B.; Mazin, S.C.; Tone, L.G.; Cardoso, V.C.; Dos Reis, R.M.; Furtado, C.L.M.; Ferriani, R.A. MIR146A and ADIPOQ genetic variants are associated with birth weight in relation to gestational age: A cohort study. J. Assist. Reprod Genet. 2022, 39, 1873–1886. [Google Scholar] [CrossRef]
Figure 1. Model of canonical microRNA biogenesis.
Figure 1. Model of canonical microRNA biogenesis.
Molecules 27 05943 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jin, M.; Xu, Q.; Li, J.; Xu, S.; Tang, C. Micro-RNAs in Human Placenta: Tiny Molecules, Immense Power. Molecules 2022, 27, 5943. https://doi.org/10.3390/molecules27185943

AMA Style

Jin M, Xu Q, Li J, Xu S, Tang C. Micro-RNAs in Human Placenta: Tiny Molecules, Immense Power. Molecules. 2022; 27(18):5943. https://doi.org/10.3390/molecules27185943

Chicago/Turabian Style

Jin, Meiyuan, Qiang Xu, Jiayong Li, Shouying Xu, and Chao Tang. 2022. "Micro-RNAs in Human Placenta: Tiny Molecules, Immense Power" Molecules 27, no. 18: 5943. https://doi.org/10.3390/molecules27185943

Article Metrics

Back to TopTop