Abstract
Micro-RNAs (miRNAs) regulate diverse cellular processes such as cell differentiation, proliferation and apoptosis. Mutation in miRNAs results in various pathological conditions such as inflammation, viral infections, neurodegeneration, autoimmunity, and so on. We have evaluated the association of miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) among patients with recurrent miscarriage (RM) and controls from North India. All the 200 patients with RM reported to experience at least 3 unexplained miscarriages before 20th week of gestation. Three hundred fertile women with no history of RMs were taken as controls. Both patients and controls were genotyped by the polymerase chain reaction amplification followed by restriction fragment length polymorphism. Variant alleles and genotypes of miR-499 A > G (Single Nucleotide Polymorphism Database [dbSNP] ID rs3746444) were found to be significant risks associated with patients having RM (odds ratio [OR] = 1.98; 95% confidence interval [CI] = 1.40-2.81; P value = .0001) and controls (OR = 3.64; 95% CI = 1.33-9.94; P value = .0109). A significant susceptible effect was found at allelic level in miR-196aT > C (dbSNP ID rs11614913) and miR-499 A > G (dbSNP ID rs3746444).
Keywords: recurrent miscarriage, single-nucleotide polymorphism, micro-RNA
Introduction
Recurrent miscarriage (RM) is defined as the loss of 3 or more consecutive pregnancies before 20th week of gestation, affecting 1% of couples trying to conceive. There are various factors involved in RMs. These are uterine anomalies, chromosomal abnormalities, endocrine dysfunction, thrombophilia, immune disorders, environmental factors, maternal infections, and so on. Among immune disorders, it is known that inflammation may regulate micro-RNA (miRNA) biogenesis. There are specific transcription factors found in the Toll-like receptor ligands, cytokines, and so on, which may alter the miRNA expression level.1 It has been reported that various cytokines may be responsible for the deregulation of Dicer expression resulting in an aberrant premature miRNA (pre-miRNA) processing.2
It has been reported that there are placenta-derived RNAs that can be detected in maternal plasma.3,4 The latter offers the possibility for noninvasive profiling of placental gene expression and the detection of diseases associated with placental dysfunction.5 Fetal miRNAs in maternal plasma can be used to obtain valuable information about the fetus or pregnancy, either for prenatal diagnosis and monitoring or for the detection of pregnancy disorders such as fetal growth restriction and so on.6 The small noncoding RNA molecules consisting of 21 to 24 nucleotides act as negative regulators of gene expression and belong to the family of miRNA. They are also involved in the posttranscriptional regulation by binding to complementary sequences on target messenger RNAs (mRNAs) and promote mRNA degradation or translational repression.7 The binding of miRNAs to mRNAs occurs in the cytoplasm in conjunction with the RNA-induced silencing complex and suppresses mRNA translation.8 Till date, more than 1500 human miRNAs have been identified and it has been estimated that miRNA targets more than 5300 human genes, which may modulate several developmental and physiological processes including pregnancy.
Most of the miRNAs are placed in the intergenic regions or antisense sequences of genes9 and are usually transcribed by RNA polymerase II to generate primary miRNAs (pri-miRNAs).10 The Pri-miRNAs get converted into pre-miRNAs by the DGCR8–Drosha complex. Premature miRNAs are further transported to the cytosolic matrix in a process involving the shuttle protein exportin 5 and cleaved into the mature miRNA form by the endoribonuclease Dicer.11 Polymorphisms of pre-miRNAs were first reported in 2005 and since then several miRNA polymorphism-association studies have been reported.12,13 Some pre-miRNA polymorphisms have negative effect on mature miRNA expression,14,15 showing that pre-miRNA sequences are important for miRNA processing.
Several studies have revealed that miRNAs play important role in different regulatory pathways such as control of developmental timing, hematopoietic cell differentiation, apoptosis, cell proliferation, and organ development.16 Micro-RNAs have also been implicated in various human diseases such as cancers, human brain dysfunctions, glioma, cardiovascular disease, primary muscular disorders, diabetes, endometriosis, preeclampsia, infertility, and RMs.17–22 The association of variants of miRNAs with RM is still not very clear. Therefore, in order to investigate whether or not there is a genetic association of the miRNA with the outcome of pregnancy, we made an attempt to investigate the role of miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) gene polymorphisms in miscarriages. We selected these miRNAs on the basis that targets of miR-146a, -149, -196a2, and -499 are FAS,23 E2F1,24 HOXB8,25 and SOX6,26 respectively. FAS induces apoptosis,27 E2F1 regulates cell cycle progression,27 HOXB8 is involved in the suppression of myeloid differentiation,28 and SOX6 is associated with transrepression of fibroblast growth factor 3 (FGF-3), promoting cell growth.26 The 146aG > C single-nucleotide polymorphism (SNP) exists in the stem region opposite to the mature miR-146a sequence, the 149C > T SNP exists in the terminal loop of the pre-miR-149 sequence, the 196a2C > T SNP is located in the 3p mature miRNA region of miR-196a2, and the 499A > G SNP is in the seed region of miR-499-3p. To the best of our knowledge, this is the first study from this part of the country.
Materials and Methods
All patients with RM were selected from patients referred to the outpatient department of Sanjay Gandhi Post Graduate Institute of Medical Sciences (SGPGIMS) Lucknow and Queen Mary Hospital of King George Medical College (Lucknow, Uttar Pradesh, India) for the evaluation of RM. In the present study, we selected 200 patients with RMs who had no known cause of RM. Patients with RM had at least 3 spontaneous miscarriages (mean 4, range 3-7) and no history of successful pregnancy. All selected patients were with primary abortion, having no live child. The patient’s detailed clinical information was recorded before inclusion in this study.
All the patients were screened for various known causes of miscarriages, including parental chromosomes, day 2 hormone levels of follicle-stimulating hormone (3-11 U/L), luteinizing hormone (3-12 U/L), troponin (0.5-3 nmol/L), antiphospholipid antibodies, including lupus anticoagulant (PLR 0.8-1.05) and anticardiolipin antibodies (immunoglobulin [Ig] G 0-12 GPL units, IgM 0-5 MPL units). Because all the patients were referrals from other hospitals, karyotyping of the product of miscarriage was not possible and only the karyotyping of both partners was carried out. Other factors screened were prothrombotic risk factors, including activated protein C resistance (2.6:4.36 ratio), factor V Leiden, and prothrombin mutations; investigation of luteal phase insufficiency, prolactin dosage, glycemic curve, thyroid hormone levels; and investigation of toxoplasmosis, cytomegalovirus, rubella, HIV, group B Streptococci, Chlamydia trachomatis, hepatitis B and C, and bacterial vaginosis. The uterine cavity was investigated for cervical incompetence by hysteroscopy, hysterosalpingography, and serial ultrasound.
The control group consisted of 300 healthy parous women of the similar ethnic distribution as the patients with RM (Supplemental Table 1) with at least 2 live births and no history of miscarriages, preeclampsia, ectopic pregnancy, preterm delivery, or systemic diseases. All patients included in this study lived in Uttar Pradesh province of northern India. Blood samples (5 mL) from control and women with RM were collected in EDTA-coated collection vials and DNA was extracted with the use of Qiagen DNA extraction kits (Brand GMbH and Co KG, Cat # 51104). This study was approved by the ethics committees of SGPGIMS and Chhatrapati Shahuji Maharaj Medical University. Written informed consent was obtained from each participant before registering them for the study.
Table 1.
Gene | SNP ID | Phenotype | Na | HWE | MAF |
---|---|---|---|---|---|
miR-146aG > C | rs2910164 | Cases | 200 | 0.188 | 0.41 |
Control | 300 | 0.085 | 0.37 | ||
miR-149T > C | rs2292832 | Cases | 200 | 0.484 | 0.20 |
Control | 300 | 0.879 | 0.17 | ||
miR-196aT > C | rs11614913 | Cases | 200 | 0.635 | 0.43 |
Control | 300 | 0.075 | 0.38 | ||
miR-499A > G | rs3746444 | Cases | 200 | 0.126 | 0.20 |
Control | 300 | 0.264 | 0.11 |
Abbreviations: HWE, Hardy–Weinberg equilibrium, MAF, minor allele frequency; miRNA, micro-RNA; SNP, single-nucleotide polymorphism. aNumber of cases and controls.
Genotype Analyses
Genetic variants of miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) were analyzed in 200 patients with RM and 300 control participants. The genotyping of the single-nucleotide variants was based on polymerase chain reaction (PCR) amplification followed by restriction digestion-based assays.20 All the amplification reactions were carried out in oil-free PTC-200 thermal cyclers (BioRad Laboratories India Pvt Ltd). In each reaction, 50 ng genomic DNA was amplified in 10 mL PCR buffer, 67 mmol/L Tris-HCl, pH 8.8, 16 mmol/L (NH4)2SO4, 2 mmol/L MgCl2, 0.01% Tween-20, and 100 mmol/L deoxynucleotides, containing 0.5 U Taq DNA polymerase. We followed a double-blind genotyping analysis where the coded DNA samples were given to the person carrying out genotyping and decoding was done by another person. Wherever we found discrepancy, the genotyping was repeated.
Statistical Analysis
The power of this study was calculated using Quanto version 1.1 (http://hydra.usc.edu/gxe) with the variables set as follows: study design, case–control; significance level, <0.05 (2 sided); model of inheritance, log additive; minor allele frequency, 0.15; and genetic effect (odds ratio [OR]) 0.6 or 1.6. The study achieved 80% power, which was sufficient to consider an OR ≤0.6 or ≥1.6, with type 1 error = 0.05. Statistical analyses for the genotypic and allelic frequencies were performed with the use of Graphpad Prism (v3.0; Graphpad Software). Allele, genotype frequencies, heterozygosities, and likelihood ratio test for Hardy-Weinberg equilibrium were calculated using Popgen v16 (www.ualberta.ca/-fyeh/fyeh). Haplotypes were generated for miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) from multilocus diploid data based on a Gibbs sampling strategy with the use of the Arlequin v3.5 software package. Allele, genotype, and haplotype frequency differences between the RM and the control groups were tested for significance with the use of Fisher exact test with Bonferroni correction. The magnitude of the effect was estimated by ORs and their 95% confidence intervals (CIs; SPSS). Logistic regression analysis was performed with different genotypes as well as alleles as the independent factors, with the risk of RM being the dependent variable: additive model comparing variant homozygous and heterozygous genotypes individually with wild-type homozygous genotypes; recessive model comparing variant homozygous genotype with wild-type homozygous and heterozygous genotypes taken together; and dominant model comparing variant homozygous and heterozygous genotypes taken together with wild homozygous genotype. P value of <0.05 was considered to be statistically significant. All statistical analyses were adjusted for age.
Results
Different alleles and their genotypes, recessive, dominant, and additive models for all miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) in this study have been investigated. The minor allele frequencies and Hardy-Weinberg equilibrium of both cases and controls are presented in Table 1. Both cases and controls were in Hardy-Weinberg equilibrium. We analyzed the effect of the genotypes of 4 miRNA polymorphisms under dominant and recessive genetic models. The genotypes of 499 A > G were risk associated in additive, dominant, and recessive models, whereas 196a2C > T revealed significant association under recessive model (Table 2). The 149C > T and 146 A > G revealed no significance. The observed heterozygosities (ho) for SNPs of miR-146aC > G, miR-149T > C, miR-196a2T > C, and miR-499A > G were in the range of 0.004 to 0.31 for control participants and did not differ significantly from the expected heterozygosities (he) which were in the range of 0.01 to 0.20. The haplotypes were constructed for miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444). There were a total of 8 haplotypes (Table 3) among the RM and controls. The haplotype estimation of miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) together showed that the haplotype consisting of wild-type alleles of all the 4 studied variables C-T-T-A was represented more often in the control group (35.5%) than in the patients group (22.5%), revealing a significant protective effect of this haplotype (OR = 0.53, 95% CI = 0.39-0.70; P < .0001). On the other hand, the haplotypes C-T-C-G and C-C-T-A were seen more frequently in the patients group (9.5% and 6.7%, respectively) than in the control group (1.8% and 3.1%, respectively) resulting in 5-fold and 2-fold risk association with RM.
Table 2.
Genotype | Patients, n = 200 (%) | Controls, n = 300 (%) | P Value | OR (95% CI) |
---|---|---|---|---|
miR-146aG > C (dbSNP ID rs2910164) | ||||
Genotype frequency | ||||
CC | 63 (31.5%) | 108 (36.0%) | 1 | – |
CG (additive model) | 107 (53.5%) | 156 (52.0%) | .4813 | 1.17 (0.79-1.74) |
GG (additive model) | 30 (15.0%) | 36 (12.0%) | .2379 | 1.42 (0.83-2.54) |
GG + CG vs CC (dominant model) | .3361 | 1.22 (0.83-1.78) | ||
GG vs CG + CC (recessive model) | .3473 | 1.29 (0.76-2.18) | ||
Allele frequency | ||||
C | 233 (58.3%) | 372 (62.0%) | .2356 | 0.85 (0.66-1.10) |
G | 167 (41.7%) | 228 (38.0%) | .2356 | 1.16 (0.90-1.51) |
miR-149T > C (dbSNP ID rs2292832) | ||||
Genotype frequency | ||||
TT | 128 (64.0%) | 207 (69.0%) | 1 | – |
TC (additive model) | 62 (31.0%) | 84 (28.0%) | .4173 | 1.19 (0.80-1.77) |
CC (additive model) | 10 (5.0%) | 9 (3.0%) | .2323 | 1.79 (0.71-4.54) |
CC + TC vs TT (dominant model) | .2461 | 1.25 (0.85-1.82) | ||
CC vs TC + TT (recessive model) | .3396 | 1.70 (0.67-4.26) | ||
Allele frequency | ||||
T | 318 (79.5%) | 498 (83.0%) | .1825 | 0.79 (0.57-1.09) |
C | 82 (20.5%) | 102 (17.0%) | .1825 | 1.25 (0.91-1.73) |
miR-196aT > C (dbSNP ID rs11614913) | ||||
Genotype frequency | ||||
TT | 65 (32.5%) | 104 (34.6%) | 1 | – |
TC (additive model) | 95 (47.5%) | 158 (52.6%) | .9185 | 0.96 (0.64-1.43) |
CC (additive model) | 40 (20.0%) | 38 (12.6%) | .0718 | 1.68 (0.97-2.89) |
CC + TC vs TT (dominant model) | .6307 | 1.10 (0.75-1.61) | ||
CC vs TC + TT (recessive model) | .0323a | 1.72 (1.06-2.80) | ||
Allele frequency | ||||
T | 225 (56.3%) | 366 (61.0%) | .1487 | 0.82 (0.63-1.06) |
C | 175 (43.7%) | 234 (39.0%) | .1487 | 1.21 (0.94-1.57) |
miR-499A > G (dbSNP ID rs3746444) | ||||
Genotype frequency | ||||
AA | 130 (65.0%) | 237 (79.0%) | 1 | – |
AG (additive model) | 58 (29.0%) | 57 (19.0%) | .0045a | 1.85 (1.21-2.83) |
GG (additive model) | 12 (6.0%) | 6 (3.0%) | .0109a | 3.64 (1.33-9.94) |
GG + GA vs AA (dominant model) | .0006a | 2.02 (1.35-3.02) | ||
GG vs GA + AA (recessive model) | .0258a | 3.12 (1.15-8.47) | ||
Allele frequency | ||||
A | 318 (79.5%) | 531 (88.5%) | .0001a | 0.50 (0.35-0.71) |
G | 82 (20.5%) | 69 (11.5%) | .0001a | 1.98 (1.40-2.81) |
Abbreviations: CI, confidence interval; dbSNP, Single Nucleotide Polymorphism Database; NS, not significant; OR, odds ratio; RM, recurrent miscarriage.
aThe difference in frequencies between the case and control groups was analyzed for statistical significance at the 95% confidence interval using Fisher exact test with Bonferroni correction under additive, recessive, as well as dominant models of inheritance. Odds ratios were calculated and reported within the 95% confidence limits. Logistic regression analysis was performed with different genotypes as well as alleles as the independent factors, with the risk of RM being the dependent variable; additive model: comparing variant homozygous and heterozygous genotypes individually with wild-type homozygous genotypes; recessive model: comparing variant homozygous genotype with wild-type homozygous and heterozygous genotypes taken together; dominant model: variant homozygous and heterozygous genotypes taken together compared with wild-type homozygous genotype.
Table 3.
Haplotypes | Patients, n = 400 (%) | Control, n = 600 (%) | OR | 95% CI | P Value |
---|---|---|---|---|---|
CTTAb | 90 (22.5%) | 213 (35.5%)c | 0.53 | 0.39-0.70 | .0001c |
GTCAc | 61 (15.3%) | 85 (14.1%)c | 1.09 | 0.76-1.55 | .6484 |
GTTAc | 70 (17.5%) | 86 (14.3%)c | 1.26 | 0.89-1.78 | .1829 |
GTTG | 8 (2.0%) | 22 (3.6%) | 0.53 | 0.23-1.21 | .1843 |
CCCAc | 35 (8.7%) | 56 (9.3%)c | 0.93 | 0.59-1.45 | .8227 |
CTTG | 14 (3.5%) | 8 (1.3%) | 2.68 | 1.11-6.46 | .0274c |
CTCAc | 26 (6.5%) | 55 (9.1%)c | 0.68 | 0.42-1.11 | .1555 |
CTCGc | 38 (9.5%) | 11 (1.8%) | 5.62 | 2.83-11.13 | .0001c |
CCTAc | 27 (6.7%) | 19 (3.1%) | 2.21 | 1.21-4.03 | .0128c |
GCTA | 6 (1.5%) | 8 (1.3%) | 1.12 | 0.38-3.27 | 1.0000 |
CCTG | 2 (0.5%) | 10 (1.6%) | 0.29 | 0.06-1.36 | .1384 |
GTCG | 11 (2.7%) | 18 (3.0%) | 0.91 | 0.42-1.95 | .8505 |
GCCA | 3 (0.75%) | 9 (1.5%) | 0.49 | 0.13-1.84 | .3806 |
GCTG | 8 (2.0%) | – | |||
CCCG | 1 (0.25%) | – |
Abbreviations: CI, confidence interval; NS, not significant; OR, odds ratio; RM, recurrent miscarriage.
aHaplotypes were generated from multilocus diploid data based on a Gibbs sampling strategy (Arlequin v3.5).
bStatistically significant protective haplotypes.
cThe difference in haplotype frequencies between the case and control groups for the prevalent haplotypes (prevalence more than 5% in either cases or controls) was analyzed for statistical significance using Fisher exact test with Bonferroni correction, and ORs are reported within the 95% confidence limits.
dStatistically significant RM predisposing haplotypes.
We applied multifactor dimension reduction analysis and observed that the best interaction model was for miR-146aC > G, miR-149T > C, miR-196a2T > C, and miR-499 A > G with testing accuracy = 0.65, cross-validation consistency = 10/10, and permutation P = 0.0001 with RM.
Discussion
In the present study, we found miR-196a2CC, miR-499AG + GG and the miR-196a2CC/miR-499AG + GG along with some of the haplotypic combinations are associated with increased risk of RM. On haplotype analysis (prevalence >5%) of miR-146aC > G (rs2910164), miR-149T > C (rs2292832), miR-196a2T > C (rs11614913), and miR-499A > G (rs3746444) variants, we observed that the haplotype C-T-T-A showed a significant protective effect (P < 0.0001), whereas the haplotypes C-C-T-A and C-T-C-G were risk-associated haplotypes with 2- and 5-fold risks. The computational example using a stimulated data set of 500 individuals for 4 miR genes revealed the predictive ability with validation accuracy of ∼60.0% for all the 4 miRNAs. The permutation testing revealed significant P values. We have observed a low hetrozygosity rate among these miRNAs. This may be due to developmental regulation and evolutionary conservation, which imply functional importance of miRNAs. Our results indicate that these miRNAs may be important in the etiology of RM. Studied miRNAs have been reported to be associated with various types of malignancies. Ryan et al revealed that miR-146a, miR-149, miR-196a2, and miR-499 are closely associated with cellular proliferation and differentiation.11
It has been demonstrated that upregulation of miR-196a2 affects mRNA expression of the HOX family of genes and Akt signaling,29 which are linked to endometriosis30 and miscarriages.20 Also, the important role of HOX genes has been implicated in implantation.29 HOXB8, target of miR-196a2,25 is essential for the myeloid differentiation and limb development.31 The main target of miR-499 is the transcriptional repressor SOX6, which suppresses the expression of FGF-3.26 The SOX6, FGF-3, and miR-499 are possibly functionally linked to breast cancer.13 Our results are in concordance with Jeon et al where it has been suggested that miR-196a2T > C and miR-499A > G polymorphisms may be responsible for recurrent spontaneous miscarriage.20 Renthal et al demonstrated that certain miRNAs can modulate uterine quiescence and contractility during pregnancy and labor.32 The evidences that support a role for this abnormal cell proliferation in RM comprise the increased risk of abortion associated with polymorphisms of the cell cycle-related genes TP53 and MDM2.33 The potential of miRNAs to act as signaling molecules, that is, miRNAs released by certain cells, can be taken up by other cells where they can elicit regulatory effects. However, no targeted miRNA expression profiling has been performed for RM-related issues. There are placenta-specific miRNAs capable of crossing the placental barrier and have been detected in maternal plasma and an altered profile of several miRNAs has been shown in pregnancy complications. One of the extensively studied is miR-146a, which binds to a site in the FAS mRNA 3′-UTR, regulating FAS expression.23 The expression of miR-146a is regulated by the nuclear factor κB pathway and elevated miR-146a levels promote the viability of mesenchymal stem cells.23 Another important target of miR-146a is SMAD4.34 The SMAD4 plays an essential role in stabilizing the undifferentiated state of human embryonic stem cells34 by regulating the promoter of p21/WAF1/Cip1. The target genes of miR-149 are Akt1 and E2F1, which are involved in promoting cell growth and cell cycle progression.24
These data suggest that miR-196a2T > C and miR-499A > G polymorphisms could contribute to RM and may be taken into consideration while evaluating these patients. Our study is the only study from Indian subcontinent and another study from Korea revealed the importance of these miRNAs20; therefore, follow-up studies in other populations, and of other pre-miRNA polymorphisms, may be important for understanding the role of miRNAs in patients with RM.
Conclusion
In conclusion, our findings demonstrate the susceptible role of miR-196aT > C (Single Nucleotide Polymorphism Database [dbSNP] ID rs11614913) and miR-499A > G (dbSNP ID rs3746444) among patients with RM. The main limitation of the present study is that we have not carried out the expression profile of these miRNAs; hence, further studies are required to clarify the associations between the miRNAs and RM. However, there exists increasing amount of evidence accumulated during the past few years which show that miRNAs are associated with numerous diseases. Moreover, noncoding RNAs have been proposed as therapeutic agents; given their silencing capabilities, this reveals the importance of our studied miRNAs.
Supplementary Material
Acknowledgments
We are thankful to the Indian Council of Medical Research for providing Research Associateship to Ms Farah Parveen. We also acknowledge Prof Shubha Phadke and Prof Vinita Das for providing well-characterized patients with RM.
Footnotes
Declaration of Conflicting Interests: The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.
Funding: The author(s) disclosed receipt of the following financial support for the research, authorship, and/or publication of this article: The grant for this project was provided by the Indian Council of Medical Research, Government of India, New Delhi, India.
Supplemental Material: The online supplementary tables are available at http://rs.sagepub.com/supplemental.
References
- 1. O'Connell RM, Rao DS, Chaudhuri AA, Baltimore D. Physiological and pathological roles for microRNAs in the immune system. Nat Rev Immunol. 2010;10(2):111–122. [DOI] [PubMed] [Google Scholar]
- 2. Wiesen JL, Tomasi TB. Dicer is regulated by cellular stresses and interferons. Mol Immunol. 2009;46(6):1222–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Chiu RW, Lo YM. Pregnancy-associated microRNAs in maternal plasma: a channel for fetal-maternal communication? Clin Chem. 2010;56(11):1656–1657. [DOI] [PubMed] [Google Scholar]
- 4. Kotlabova K, Doucha J, Hromadnikova I. Placental-specific microRNA in maternal circulation—identification of appropriate pregnancy-associated microRNAs with diagnostic potential. J Reprod Immunol. 2011;89(2):185–191. [DOI] [PubMed] [Google Scholar]
- 5. Su MT, Lin SH, Chen YC. Genetic association studies of angiogenesis- and vasoconstriction-related genes in women with recurrent pregnancy loss: a systematic review and meta-analysis. Hum Reprod Update. 2011;17(6):803–812. [DOI] [PubMed] [Google Scholar]
- 6. Ladomery MR, Maddocks DG, Wilson ID. MicroRNAs: their discovery, biogenesis, function and potential use as biomarkers in non-invasive prenatal diagnostics. Int J Mol Epidemiol Genet. 2011;2(3):253–260. [PMC free article] [PubMed] [Google Scholar]
- 7. Morales-Prieto DM, Chaiwangyen W, Ospina-Prieto S, et al. MicroRNA expression profiles of trophoblastic cells. Placenta. 2012;33(9):725–734. [DOI] [PubMed] [Google Scholar]
- 8. Bartel DP. MicroRNAs: target recognition and regulatory functions. Cell. 2009;136(2):215–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Iwai N, Naraba H. Polymorphisms in human pre-miRNAs. Biochem Biophys Res Commun. 2005;331(4):1439–1444. [DOI] [PubMed] [Google Scholar]
- 10. Xu J, Hu Z, Xu Z, et al. Functional variant in microRNA-196a2 contributes to the susceptibility of congenital heart disease in a Chinese population. Hum Mutat. 2009;30(8):1231–1236. [DOI] [PubMed] [Google Scholar]
- 11. Ryan BM, Robles AI, Harris CC. Genetic variation in microRNA networks: the implications for cancer research. Nat Rev Cancer. 2010;10(6):389–402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Zhou B, Rao L, Peng Y, et al. Common genetic polymorphisms in pre-microRNAs were associated with increased risk of dilated cardiomyopathy. Clin Chim Acta. 2010;411(17-18):1287–1290. [DOI] [PubMed] [Google Scholar]
- 13. Hoffman AE, Zheng T, Yi C, et al. MicroRNA miR-196a-2 and breast cancer: a genetic and epigenetic association study and functional analysis. Cancer Res. 2009;69(14):5970–5977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Suzuki Y, Kim HW, Ashraf M, Haider Hkh. Diazoxide potentiates mesenchymal stem cell survival via NF-κB-dependent miR-146a expression by targeting Fas. Am J Physiol Heart Circ Physiol. 2010;299(4):H1077–H1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Lin RJ, Lin YC, Yu AL. miR-149 induces apoptosis by inhibiting Akt1 and E2F1 in human cancer cells. Mol Carcinog. 2010;49(8):719–727. [DOI] [PubMed] [Google Scholar]
- 16. Kefas B, Godlewski J, Comeau L, et al. microRNA-7 inhibits the epidermal growth factor receptor and the Akt pathway and is down-regulated in glioblastoma. Cancer Res. 2008;68(10):3566–3572. [DOI] [PubMed] [Google Scholar]
- 17. Thum T, Catalucci D, Bauersachs J. MicroRNAs: novel regulators in cardiac development and disease. Cardiovasc Res. 2008;79(4):562–570. [DOI] [PubMed] [Google Scholar]
- 18. Eisenberg I, Eran A, Nishino I, et al. Distinctive patterns of microRNA expression in primary muscular disorders. Proc Natl Acad Sci USA. 2007;104(43):17016–17021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Joglekar MV, Parekh VS, Hardikar AA. Islet-specific microRNAs in pancreas development, regeneration and diabetes. Indian J Exp Biol. 2011;49(6):401–408. [PubMed] [Google Scholar]
- 20. Jeon YJ, Choi YS, Rah H, et al. Association study of microRNA polymorphisms with risk of idiopathic recurrent spontaneous abortion in Korean women. Gene. 2012;494(2):168–173. [DOI] [PubMed] [Google Scholar]
- 21. Fish JE, Srivastava D. MicroRNAs: opening a new vein in angiogenesis research. Sci Signal. 2009;2(52):pe1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Kotlabova K, Doucha J, Hromadnikova I. Placental-specific microRNA in maternal circulation--identification of appropriate pregnancy-associated microRNAs with diagnostic potential. J Reprod Immunol. 2011;89(2):185–191. [DOI] [PubMed] [Google Scholar]
- 23. Fluhr H, Wenig H, Spratte J, et al. Non-apoptotic Fas-induced regulation of cytokines in undifferentiated and decidualized human endometrial stromal cells depends on caspase-activity. Mol Hum Reprod. 2011;17(2):127–134. [DOI] [PubMed] [Google Scholar]
- 24. Carcagno AL, Marazita MC, Ogara MF, et al. E2F1-mediated upregulation of p19INK4d determines its periodic expression during cell cycle and regulates cellular proliferation. PLoS One. 2011;6(7):e21938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Fujino T, Yamazaki Y, Largaespada DA, et al. Inhibition of myeloid differentiation by Hoxa9, Hoxb8, and Meis homeobox genes. Exp Hematol. 2001;29(7):856–863. [DOI] [PubMed] [Google Scholar]
- 26. Murakami A, Ishida S, Thurlow J, et al. SOX6 binds CtBP2 to repress transcription from the Fgf-3 promoter. Nucleic Acids Res. 2001;29(7):3347–3355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Sluijter JP, van Mil A, van Vliet P, et al. MicroRNA-1 and -499 regulate differentiation and proliferation in human-derived cardiomyocyte progenitor cells. Arterioscler Thromb Vasc Biol. 2010;30(4):859–868. [DOI] [PubMed] [Google Scholar]
- 28. Kawasaki H, Taira K. MicroRNA-196 inhibits HOXB8 expression in myeloid differentiation of HL60 cells. Nucleic Acids Symp Ser (Oxf). 2004;(48):211–212. [DOI] [PubMed] [Google Scholar]
- 29. Makker A, Goel MM, Das V. PI3K-Akt-mTOR and MAPK signaling pathways in polycystic ovarian syndrome, uterine leiomyomas and endometriosis: an update. Gynecol Endocrinol. 2012;28(3):175–181. [DOI] [PubMed] [Google Scholar]
- 30. Vitiello D, Kodaman PH, Taylor HS. HOX genes in implantation. Semin Reprod Med. 2007;25(6):431–436. [DOI] [PubMed] [Google Scholar]
- 31. Hornstein E, Mansfield JH, Yekta S, et al. The microRNA miR-196 acts upstream of Hoxb8 and Shh in limb development. Nature. 2005;438(7068):671–674. [DOI] [PubMed] [Google Scholar]
- 32. Renthal NE, Chen CC, Williams KC, et al. MiR-200 family and targets, ZEB1 and ZEB2, modulate uterine quiescence and contractility during pregnancy and labor. Proc Natl Acad Sci USA. 2010;107(48):20828–20833. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Fang Y, Kong B, Yang Q, et al. MDM2 309polymorphism is associated with missed abortion. Hum Reprod. 2009;24(6):1346–1349. [DOI] [PubMed] [Google Scholar]
- 34. Zhong H, Wang HR, Yang S, et al. Targeting Smad4 links microRNA-146a to the TGF-beta pathway during retinoid acid induction in acute promyelocytic leukemia cell line. Int J Hematol. 2010;92(1):129–135. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.