Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jan 29.
Published in final edited form as: Am J Reprod Immunol. 2009 Aug;62(2):78. doi: 10.1111/j.1600-0897.2009.00716.x

Epigenetics in the Placenta

Matthew A Maccani 1, Carmen J Marsit 1
PMCID: PMC2813777  NIHMSID: NIHMS168652  PMID: 19614624

Abstract

Epigenetics is focused on understanding the control of gene expression beyond what is encoded in the sequence of DNA. Central to growing interest in the field is the hope that more can be learned about the epigenetic regulatory mechanisms underlying processes of human development and disease. Researchers have begun to examine epigenetic alterations – such as changes in promoter DNA methylation, genomic imprinting, and expression of miRNA – to learn more about epigenetic regulation in the placenta, an organ whose proper development and function are crucial to the health growth and survival of the developing fetus. A number of studies are now making important links between alterations to appropriate epigenetic regulation in the placenta and diseases of gestation and early life. In addition, these studies are adding important insight into our understanding of trophoblast biology and differentiation as well as placental immunology. Examining epigenetic alterations in the placenta will prove especially important in the search for biomarkers of exposure, pathology, and disease risk and can provide critical insights into the biology of development and pathogenesis of disease. Thus, epigenetic alterations may aid in disease diagnosis and prognosis as well as in targeting new treatment and prevention strategies.

Keywords: DNA methylation, environmental exposure, miRNA, imprinting


Our understanding of the interplay between genes and the environment is being greatly enhanced in the post-genome era. There are only a few settings where the importance of this gene-environment interface is more profound than during intrauterine development, where the “critical windows” are narrower and where disruption or modification can influence fetal development as well as lead to programming of health throughout the life course. This phenomenon, now known as “fetal programming”, is a model of gene environment interaction and can inform the mechanistic basis of the synergistic effect(s) of the environment and the molecular character of development1, 2.

Research in fetal programming and many other disciplines is now focusing on the paradigm that gene regulation occurs beyond the DNA sequence. Most of the acquired adverse effects related to intrauterine environment cannot be due to genetic alterations. This critical role of epigenetic regulation, the mitotically and meiotically heritable control of gene expression not related to DNA sequence, during development is becoming increasingly appreciated. Thus, an understanding of changes to the cellular epigenome is at the interface of the interaction between genes and environment, and can provide a mechanistic basis for the synergistic effects. Research in model systems and now expanding to human studies has suggested that the causes and consequences of a variety of pathologies are related to environmental influence on epigenetic regulation. Examination of the specific molecular character of these epigenetic alterations in perinatal development has been less comprehensive.

Throughout the in utero development of the fetus, the placenta is of utmost importance to ensure proper growth and development. Called by some a hallmark of mammalian development 3, the placenta functions to provide the fetus with nutrients, allows for waste to be transferred and ultimately excreted by the mother, and protects the fetus from what would otherwise be a type of immune detection which would result in an attack of the placenta by the maternal immune system. In addition, the placenta has a degree of metabolic and endocrine activity, is involved in secreting hormones responsible for maintaining and regulating various stages of pregnancy, and performs biochemical reactions to protect whenever possible the fetus from exposure to toxicants or other harmful chemicals4. All of these functions of the placenta as well as placental gene expression thus respond to and are marked by environmental insults 4, 5, and in many ways, the placenta can serve as a record of in utero exposure and pathology. Various compounds and drugs, including but not limited to alcohol 6, nicotine 7, cocaine 8 9, lead 10, and phthalates 11, have been shown to cross the placenta and alter placental gene expression; some even accumulate in the placental tissue. Investigations are underway to determine changes in the genetics and epigenetics of the placenta which are characteristic of such exposures and pathological responses resulting from these exposures.

Epigenetics

Epigenetics is broadly defined as the field of research which studies changes in gene expression that are not caused by changes in the sequence of DNA 12, and the field has seen relatively rapid growth over the past few decades, accelerated by advancements in molecular biology, biotechnology, and genomics. The emergence of a new field – namely, “environmental epigenetics” 13 – combines the traditional ways of studying epigenetics with the understanding that environmental exposures affect such epigenetic mechanisms as well. As the field of epigenetics continues to grow and be defined, there remains a central focus on examination of 4 main modes of epigenetic regulation: DNA methylation, imprinting, histone modification, and small RNA-mediated control, specifically miRNAs.

DNA Methylation

DNA methylation has become the most heavily studied mode of epigenetic regulation 12. In brief, DNA methylation is carried out by one of a variety of DNA methyltransferases responsible for adding a methyl group to cytosine residues in cytosine/guanine-rich regions of DNA (called “CpG islands”). A general rule (one that is usually, but not always, true) is that when a given stretch of cytosines in a CpG island or islands located in the promoter region of a gene is methylated, that gene will be effectively silenced by methylation; such a CpG island would be termed “hypermethylated”. Conversely, when a given stretch of cytosines in a CpG island or islands located in the promoter region of a gene is not methylated, that gene will not be silenced by methylation; the CpG island in this case would be said to be “hypomethylated”. It should also be noted that it is not the methylation of DNA itself which contributes to transcriptional repression but rather the binding of various elements (proteins that act as transcriptional repressors, proteins that block the movement of RNA polymerase, etc.) to methylated stretches of DNA that most greatly contribute to the transcriptional repression characteristic of genes with methylated CpG islands in their promoters. Researchers continue to work on attempting to decipher a type of chromatin code – one that may give scientists clues as to what degree of promoter methylation as well as interaction with histone post-translation modifications may be necessary to silence a particular gene.

Throughout the development of the embryo, important resetting of methylation patterns of germline and somatic lineages occur. Methylation throughout the genome of the zygote is almost completely removed during the cleavage phase of development; in between the implantation and gastrulation phases of development, de novo methylation reestablishes the developing organism's methylation patterns which, under normal conditions, are maintained throughout the rest of the organism's life 3, 14-16. This patterning is not limited to the embryo, but also occurs in a specific fashion in the extraembryonic lineages, although the overall levels of methylation in extraembryonic cell lineages are significantly lower than that in the somatic lineage 3. Crucial for the health and survival of the organism is the need for the appropriate removal and resetting of methylation patterns during development thereby making this period a critical window during which the environment can have profound effects on the epigenetic pattern of the offspring.

Advances in technology have given researchers tools to measure changes in DNA methylation marks and patterns. Initial studies relied on the use of methylation-sensitive restriction enzymes and Southern blotting with site specific probes to characterize DNA methylation in specific genomic regions, but such techniques allowed only for examinations of methylation at specific restriction enzyme sites and required large quantities of DNA for study. Chemical modification of the DNA by sodium bisulfite, which leads to the deamination of unmethylated cytosines to uracil, but maintains methylated cytosines as cytosine allowed for the sequencing of stretches of DNA to determine cytosine methylation in a genomic region17. Relying on sodium bisulfite conversion, methylation-specific PCR (MSP) uses oligonucleotide primers which will bind and amplify sequences of sodium bisulfite modified template DNA in a methylation-specific fashion, enabling researchers to determine changes in single site DNA methylation18. Quantitative real-time PCR as well as short read sequencing such as pyrosequencing now allow for relative quantification of the methylation status at multiple CpG sites within a region, reflecting the prevalence of methylated alleles within the template DNA19-21. Sodium bisulfite conversion also lends itself to high-throughput array based approaches similar to those used for high-throughput genotyping such as Illumina Inc.'s Infinium Methylation27 assay which can simultaneously determine the cytosine methylation status at >27,000 CpG sites in human samples. Immunoprecipitation of methylated DNA followed by detection of the methylated fraction through array based approaches has become common particularly in model systems22. Next generation sequencing technologies based either on sodium bisulfite modification or on the immunoprecipitation methods are the next frontier in DNA methylation detection technologies which can be applied to both human and model system samples23, and will certainly add to our understanding of DNA methylation in the placenta.

Imprinting

Epigenetic regulation is central to the phenomenon of genomic imprinting, the parent-of-origin, allele specific expression of genes. Genes controlled through imprinting are often located and regulated coordinately in clusters. Imprinted genes are theorized to be controlled at differentially methylated regions (DMRs) by DNA methylation24. One type of DMR is one that is differentially methylated in all tissues throughout development and is commonly called an imprinting control region (ICR) because such ICRs are hypothesized to be key regulators of imprinting in their particular chromosomal domains 25. The other type of DMR is one that has differential patterns of tissue-specific methylation during stages of somatic development 25.

Both non-coding RNAs and changes in DNA methylation at sites in DMRs are responsible for the regulation of the imprint. Although DNA methylation is involved, it does not function in a manner similar to that seen in promoter regions, but instead functions to alter the binding of specific transcription factor and/or enhancer elements which control the allele specific expression of the region26, 27. The marks of imprinting are erased in germline cells, and re-established dependent on the sex of the individual (i.e. in sperm, paternal imprints become established and in oocytes, maternal imprints)28.

Imprinting has been theorized to be one of the mechanisms involved in the so-called “parent conflict” theory 29. The “parent conflict theory” suggests that paternally expressed genes strongly favor using maternal resources to benefit offspring while maternally expressed genes attempt to preserve such maternal resources and thus, are in direct conflict with one another29. In such a way, one could argue that paternally expressed (and maternally imprinted) genes would work to foster the growth of offspring while maternally expressed (and paternally imprinted) genes would function to better ensure that each offspring has approximately the same access to maternal resources as its siblings30.

Imprinted genes are thought to function in the control of embryonic development, including placental development31, as well as in functions later in life such as behaviors and metabolism32. Alterations to normal imprinting patterns lead to well characterized syndromes, including Praeder-Willi and Angelman Syndromes related to inappropriate imprinting at chromosome 15q11. Developmental environment has also been suggested to affect the appropriate establishment of imprinted genes, particularly at chromosome 11p15 leading to Beckwith-Wiedemann Syndrome, which has been linked to the use of assisted reproductive technologies33-35. Although methylation patterns are critical in the regulation of genomic imprinting, their utility to examine imprinted genes is limited by the ability to determine allelic specific methylation patterns. Most often, imprinting status is determined using allele specific PCR reactions to examine allele specific expression36, although new methods based on genome-wide SNP arrays are allowing for the examination of allelic-specific expression and alterations to normal imprinting status on the genomic scale37, 38, and may be useful for the examination of imprinting patterns in the placenta and their association with normal fetal growth and development.

Histone modification

Modifications of the chromatin environment play key roles in epigenetic regulation of gene expression as well. One of these epigenetic regulatory mechanisms involves the acetylation, methylation, phosphorylation, and ubiquitinylation of histones, leading to regulation of gene expression 39, 40. The combined effects of the modification of the amino-terminal tails of core histones by acetylation, phosphorylation, and methylation play a major role in determining gene activity 41-43. A number of classes of histone methyltransferases – key enzymes involved in the transfer of methyl groups to histones – have been discovered 42, including the H3-K4 methyltransferase 44, 45 and five H3-K9 methyltransferases 46-49. Additionally, researchers have identified a number of transcription co-activators that have characteristic histone acetyltransferase (HAT) activity and histone deacetylases (HDACs), both of which play important roles in histone modification 50.

Histone modifications can be established in particular reactions or in a sequential order 41, 43; recent studies have indicated that the sequential order of modifications may be gene specific 51, 52. Other research has suggested that site-specific combinations of covalent histone modifications may comprise a type of histone code that can not only affect the structure of chromatin but can also affect targeting of transcriptional complexes 53-55. Such histone modifications can lead to gene activation or gene silencing, depending on the effects on transcriptional complexes. Alterations to patterns of histone modification can have a number of negative consequences, such as developmental dysregulation, X-chromosome inactivation, or might lead a number of diseases 50. Research is continuing to better define how the patterns of histone modifications are utilized by the cell to control gene expression, as well as how these marks are involved in regulating additional epigenetic processes 56.

Small regulatory RNAs and microRNAs (miRNA)

In the early 1990s, researchers first published observations characterizing two small regulatory RNAs, known as lin-4 and let-7, which were shown to control the timing of larval development in C. elegans 57, 58. These RNAs, initially termed “lin-4 and let-7 RNAs”, were initially suggested to represent a class of endogenous RNAs found in worms, flies, and mammals and since have been renamed “microRNAs (miRNAs)” 59-61. Subsequent work suggested that these small regulatory RNAs could be found in plants, mammals, green algae, and viruses 62. Other classes of small RNAs have been found in plants, animals and fungi; small interfering RNAs (siRNAs) 63, 64 and Piwi-interacting RNAs (piRNAs) 65 are two examples. miRNAs are different from these other classes of small RNAs in that they are formed from transcripts that have been shown to fold back on themselves, generating characteristic hairpin structures 66; other small RNA classes are formed from longer hairpins (siRNAs) or from precursor forms lacking a double-stranded nature (piRNAs) 66. Generally, as these small regulatory RNA molecules can alter gene and protein expression without altering the underlying genetic code, they too are considered critical mechanisms in epigenetic regulation.

miRNA are transcribed by RNA Polymerase II as part of transcripts called primary miRNAs (pri-miRNAs) and include 5′ caps and 3′ poly(A) tails 67-69. The miRNA portion of the pri-miRNA then forms a hairpin 67. The pri-miRNA is then digested by the dsRNA-RNA-specific ribonuclease Drosha, and the hairpin that is ultimately released is called precursor miRNA (pre-miRNA) 70. pre-miRNA has been characterized to be 70-75 nucleotides of RNA in length with 1-4 nucleotide 3′ overhangs, 25-30 base pair stems, and small loops 70, 71. Data have also suggested that Drosha processes either the 5′ or 3′ terminus of the mature miRNA, depending on which strand of the pre-miRNA associates with the RNA-induced silencing complex (RISC) 70, 71. The pre-miRNA is then exported from the nucleus to the cytoplasm by a complex containing Exportin-5 (Exp5) 71, 72. After arrival in the cytoplasm, the pre-miRNA is cleaved by Dicer, an RNase III superfamily member 70, 71. After cleavage by Dicer, the resulting double-stranded RNA has short 3′ overhangs at either end 72. It should be noted that only one of the two strands in the post-Dicer-processed dsRNA is the true mature miRNA; some mature miRNAs are formed from the leading strand of the miRNA strand while others are formed from the lagging strand, in a mature miRNA-specific fashion 72. In order to effectively control the translation of target mRNAs, the dsRNA that Dicer has processed must be separated into two strands, and the single-stranded mature miRNA has to associate with the RISC in order to be trafficked to its mRNA target 73. Researchers have shown that determination of the active strand in the ds-RNA has a direct relationship with the stability of the ends of the dsRNA 74, 75; in brief, their work revealed that the strand with less stable base pairing of the 2-4 nucleotides at the 5′ end of the duplex associates with RISC and ultimately takes on the role as the active miRNA strand 74.

Data have suggested that miRNA regulate gene expression by base-pairing to a target mRNA transcript; the exact mechanism for this post-transcriptional gene regulation varies depending on a number of factors, the most noteworthy of which seems to be the degree of complementarity of the miRNA to its target mRNA sequence 59. The active strand of the mature miRNA associates more specifically with the Argonaute protein of the RISC and upon trafficking to the target mRNA, participates in post-transcriptional repression 73, 76. As a general rule, a miRNA with perfect complementarity to its target mRNA will cause the degradation of the mRNA transcript through Argonaute-catalyzed mRNA cleavage 73, 77, 78 while miRNA with imperfect complementarity to a target mRNA will cause translational repression by blocking or altering the normal function of machinery that would otherwise aid in translation of mRNA into protein 59. Mechanisms by which translational repression is carried out include inhibition of translation initiation and poly(A) shortening 79. Some groups have also shown data which suggest that miRNA can use a combination of both mRNA degradation and disruption of translation to carry out post-transcriptional repression 80.

miRNA have been shown to carry out important roles in a number of responses to stress and disease. Work by van Rooij and colleagues demonstrated that a cardiac-specific miRNA, miR-208, was crucial for hypertrophy of cardiomyocytes and fibrosis in response to stress and hypothyroidism 81. Findings such as this have led several to hypothesize that there may be important miRNA-based therapies for heart disease that have yet to be developed 82. Several groups have suggested roles for miRNA in preventing or even contributing to the development and progression of cancer 83. miR-21, shown to be upregulated in human brain tumor glioblastoma, has also been shown to have anti-apoptotic properties in human glioblastoma cells 84. Several other miRNAs have been described as having oncogenic or even tumor-suppressive characteristics, and many have been selected for use in developing cancer-specific therapies 83.

There exist a number of technologies and methods that have given researchers the power to measure even the smallest of changes in miRNA expression. Reverse transcription quantitative real-time PCR has been used to determine changes in the expression of particular miRNA associated with particular exposures or diseases. High-throughput assays, including miRNA microarrays, have been used to interrogate the expression of thousands of reported and predicted miRNA sequences in tissues that have a particular exposure or disease. Downstream effects of miRNA on post-transcriptional gene regulation remain more challenging and require use of bioinformatics approaches to first predict mRNA targets of specific miRNA and then confirm the effects of over- or underexpression of miRNA on that particular target. Due to the miRNAs' ultimate role in controlling protein translation, true confirmation of targets requires examination of the proteins of interest using specific antibodies, or through in-vitro approaches coupling targeted miRNA binding regions to reporter constructs. The understanding of miRNAs' role in the human placenta is in its infancy but due to the known critical role of miRNA in human development, it is certainly an attractive and exciting field of study.

DNA promoter methylation in placental genes

Researchers focusing on the placentas of non-humans have discovered that hypermethylation in the placentas of cloned cats may be associated with decreased cloning success rates 85 and that a particular cytochrome P450 gene is controlled, in part, by varying methylation status in the placentas of sheep and cattle 86. In humans as early as the mid-1980s, researchers demonstrated the effects of site-specific DNA methylation on the binding ability of DNA-binding protein in human placenta 87,contributing greatly to understanding the effects of DNA methylation patterns on the recruitment of DNA binding elements and subsequent effects on transcription 87.

Some have further analyzed mechanisms of human placentation in the context of its similarities with human cancer progression and tumorigenesis. More specifically, data have shown that the epigenetics related to normal human placental invasion and function have striking similarities to particular patterns of tumor-associated methylation involved in the coordinated set of epigenetic silencing events at play in tumorigenesis and progression 88.

Still others have focused on links between aberrant methylation patterns of placental gene promoters and disease progression. Work by Zhang and colleagues has further demonstrated that the Oct4 transcription factor whose hypermethylation is associated with downregulation of gene expression and differentiation of trophectoderm cell lineage is downregulated by increased methylation in normal placenta and in gestational trophoblastic disease (GTD), an epigenetic regulatory mechanism which may prove important in the development and progression of GTD 89. In work to better characterize the genetic and epigenetic factors underlying the onset and progression of preeclampsia, Chelbi and colleagues suggested that aberrant methylation patterns may be a typical mechanism ultimately leading to preeclampsia 90.

Others have investigated placenta-specific methylation patterns which maximize the bioavailability of essential vitamins, specifically vitamin D, at the fetomaternal interface 91. Park and colleagues have even demonstrated the association of folate and homocysteine levels and DNA methylation levels in the human placenta 92; more specifically, their data suggest that maternal levels of folate and homocysteine, as well as a particular polymorphism in MTHFR 677, influence DNA methylation patterns in the placenta during pregnancy 92.

Taken collectively, these efforts have provided the scientific and clinical community with a better understanding of the mechanisms and pathways underlying DNA methylation's involvement in epigenetic regulation of key placental processes.

Imprinting in the placenta

It is thought that genomic imprinting may play a critical role in placental biology, as the control of allelic expression is exaggerated in the placenta and alterations to these imprints have been linked to severe placenta pathologies93. At the same time, less well characterized are the role that imprinting alterations may play in more common, placental-related pathologies including intrauterine growth restriction and preeclampsia. Some groups have focused on characterizing patterns of imprinting in the human placenta in hopes of finding new biomarkers that could be affected by prenatal conditions. More specifically, Lambertini and colleagues have utilized techniques to measure loss of imprinting (LOI) in genes in human placentas and have concluded that not only is LOI common in human placentas but may also serve as a key biomarker for epigenetics affected by prenatal conditions or environment 94. Guo and colleagues have investigated gene expression and methylation patterns of imprinted regions in small for gestational age (SGA) placentas and have shown that loss of imprinting at H19 due to methylation alterations and subsequent effects on gene expression may be some causes of poor growth of the human fetus 5.

Haycock and colleagues used a mouse model to investigate the effects of imprinting control disruption at the H19/IGF2 domain as a mechanism of aberrant growth in the event of ethanol exposure during fetal development 24. Their results suggested that paternal alleles in placentas treated with ethanol exhibited a significantly lower degree of methylation than those in placentas treated with saline control. Analysis of data showing a relationship between placental weight and ethanol treatment further revealed a possible relationship between DNA methylation at the CCCTC-binding factor site on the paternal allele in placentas 24.

McMinn and colleagues conducted a rather thorough investigation of associations of intrauterine growth restriction (IUGR) with altered patterns of expression of imprinted genes 95. They performed a genome-wide survey which suggested what they termed an “unbalanced” expression of imprinted genes in IUGR placentas compared to the imprinted gene expression in non-IUGR placentas. In addition, they also demonstrated a degree of differential expression of non-imprinted genes in IUGR vs. non-IUGR placentas 95. Their research suggested that differential expression of a panel of imprinted genes may be a possible biomarker for IUGR.

Collectively, these results suggest that environmental insults and factors associated with diseases such as IUGR may impact imprinting control mechanisms in the placenta. More work needs to be done to gain a better understanding of more specific mechanistic effects of such exposures on mechanisms of imprinting in the placenta and downstream consequences in fetal growth and development.

Histone modification in the placenta

Examinations of histone post-translational modification are focused currently on both trophoblast biology and differentiation, as well as in placental immunology. Kimura and colleagues investigated the acetylation and methylation patterns of core histones within and flanking the human growth hormone (hGH) multigene cluster in human placental chromatin 40. Their data further distinguished the differences between placental and pituitary mechanisms of transcriptional control at the hGH cluster and suggested that the selective nature of placental gene activation seems to point toward unique roles for histone acetyltransferase and histone methyltransferase coactivator complexes in the regulation of placental gene expression 40.

Morris and colleagues utilized chromatin immunoprecipitation (ChIP) assays to further investigate transcription factor assembly and histone modifications that occur during gamma interferon (IFN-gamma) induction of the master regulator of major histocompatibility complex class II (MHC class II) transcription, known as CIITA 96. Treatment of most cells with IFN-gamma causes the induction of MHC class II genes; however, trophoblast cells will not upregulate MHC class II following exposure to IFN-gamma 97, 98. Such inability to upregulate MHC class II genes when exposed to IFN-gamma has been suggested to be one possible mechanism relating to maternal-fetal tolerance, in which the mother's immune system does not react to placental tissues that are expressing paternally-derived (and thus recognized as foreign) genes 96.

As described by Morris and colleagues, the major IFN-gamma responsive promoter for the expression of CIITA, Promoter IV (PIV), needs both STAT1 and IFN regulatory factor 1 (IRF-) for induction by IFN-gamma. The investigators first noted the binding of STAT1 to PIV and also noted the association of a medium degree of histone H3 and H4 acetylation 96. However, CIITA expression was not detected until IRF-1 protein was synthesized and properly bound to its necessary site. When using fetal trophoblast-like cell lines that are resistant to CIITA induction by IFN-gamma, the authors observed that the cells could not properly assemble factors which would otherwise enable them to be inducible by IFN-gamma; moreover, these cells were also unable to modify their chromatin, a characteristic which suggested that the promoter region may have been blocked. Subsequent experiments showed that PIV exhibited a strong degree of hypermethylation. Taken collectively, Morris and colleagues' data furthered the understanding of the activation of the CIITA gene in response to IFN-gamma and suggest that assembly of regulatory factors, modification of chromatin, and expression of genes are linked and may progress in a distinct series of steps 96.

Other groups, such as Chuang and colleagues, have investigated the effects of acetylation mediated by histone acetyltransferases (HATs) and the expression of important genes which play a role in the mediation of trophoblastic fusion 99. Chuang and coworkers investigated key histone deacetylases (HDACs) that are involved in the deacetylation of the human GCMa transcription factor which plays a role in regulating syncytin – a placental protein that mediates trophoblastic fusion. Their data gave support to the theory that trophoblastic fusion in placental morphogenesis is largely dependent on the proper regulation of GCMa by HAT and HDAC 99.

Collectively, research into the effects of histone modification on placental gene expression continues to unveil a better understanding of this important epigenetic regulatory mechanism. Such knowledge has enabled researchers to gain a better understanding of important phenomena such as maternal-fetal tolerance and placental differentiation and growth.

Placental miRNA

Several groups have published data demonstrating that miRNA expression is tissue-specific and that several miRNAs are expressed in the human placenta 80, 100. Since such discoveries, much interest has been generated for investigating the involvement of miRNA in placental gene regulation and the possible utility of discovering placental miRNA which can serve as clinical biomarkers of exposure or disease.

Pineles and colleagues measured miRNA levels in placentas to investigate whether placental miRNA expression patterns are associated with preeclampsia or small for gestational age diagnoses 101. This work revealed differential expression of miR-210 and miR-182 in preeclampsia versus control patients and showed key associations in miRNA expression patterns and preeclampsia, furthering efforts to find a miRNA biomarker of preeclampsia pathology 101. Zhu and coworkers found that thirty four miRNAs were differentially expressed in preeclamptic placentas compared to normal placentas further suggesting a role for miRNA in the pathogenesis of preeclampsia 102. Such biomarkers would prove especially useful in giving clinicians and pathologists sensitive molecular tools to better diagnose preeclampsia, a disorder that has great impacts on both pregnant women and their newborns.

It has long been speculated that varying oxygen levels have effects on the genetic and epigenetic control mechanisms involved in placental growth and cell survival – and ultimately, on the health and survival of the developing fetus. To investigate this further, Donker and coworkers analyzed the relationships between expression of Argonaute 2, an important RNAi enzyme, and other miRNA in trophoblasts and in environments with varying oxygen level 103. Donker and colleagues' data showed that not only is the miRNA processing machinery present and functional in human trophoblasts but that varying expression of miR-93 and miR-424 is associated with different levels of oxygen 103. Such data may prove especially helpful in determining whether particular placental abnormalities – and ultimately, fetal abnormalities – may be associated with aberrant levels of oxygen at particular critical windows of development.

Some groups have observed and detected placental miRNA in maternal plasma and have conducted comparative studies on circulating miRNA and other circulating nucleic acids (CAN) in the maternal sera of pregnant and non-pregnant women 104, 105. Chim and colleagues showed that placental miRNAs (miR-141, miR-149, miR-299-5p, and miR-135b) were highly expressed in maternal plasma in pregnant mothers and suggested that such expression patterns could serve as important biomarkers for monitoring pregnancy 104. Gilad and coworkers demonstrated that placental miRNA levels in sera from pregnant women were higher than those in sera from non-pregnant women; moreover, levels of miRNA in sera of pregnant women correlated with pregnancy stage 105. Thus, placental miRNA levels detectable in maternal serum may serve as important clinical biomarkers of pregnancy, pregnancy stage, and other pregnancy-related outcomes.

Future Directions in Placental Epigenetics

Data have suggested that environmental exposures can alter miRNA expression in in-vitro systems 106 and various tissues 13 and that altered levels of miRNA expression can be associated with particular diseases or risk factor for disease 107. As reviewed above, work has also been done to investigate biomarkers of diseases, such as cancer, growth retardation, and other records of in utero environment in placental patterns of DNA methylation, imprinting, histone modification, and miRNA expression. Studies aimed at further revealing the importance of miRNA in responding to environmental exposure and to disease may prove very useful in better understanding how such exposures and diseases affect the body. Current and past research into diseases of pregnancy – such as finding miRNA associated with preeclampsia 101, 102 or gene-specific placental methylation patterns associated with GTD 89 – might be expanded to include work to better understand associations between epigenetic factors in the placenta and diseases such as choriocarcinoma, childhood cancers, and other diseases of childhood and adolescence, as well as alterations to normal placental immunology and function.

Using the placenta as a reference tissue allows researchers to utilize an important residual tissue whose respective miRNA expression and DNA methylation patterns may prove to be powerful biomarkers possessing predictive capability for a number of diseases or disease progression. Modern advances in bioinformatics – such as pathway analysis tools and target gene prediction software – as well as advances in technology, such as microarray technology, have given researchers and clinicians tools to better detect patterns in DNA methylation, imprinting, histone modification and miRNA expression that are associated with particular exposures and diseases. Aberrant patterns of miRNA expression or DNA methylation may ultimately serve as biomarkers for exposure, disease burden, or even as “early indicator” diagnostics of increased risk for developing future disease or disorders.

Conclusion

Important advances in placental epigenetics continue to elucidate a better of understanding of the epigenetic regulatory mechanisms of in the placenta. Knowledge of such epigenetic mechanisms may be useful in identifying novel biomarkers for exposure, burden, or risk for disease. Such biomarkers may prove essential for developing new diagnostics for early diagnosis of risk factor and levels of exposure. Additionally, these aberrant patterns of miRNA expression, imprinting, DNA methylation, or histone modification may identify previously unknown pathways targeted for alteration, which, in turn, may serve as targets for novel drug treatment or prevention strategies. These epigenetic biomarkers can be brought from the benchtop to the bedside and will be useful in helping clinicians better diagnose and prevent the onset of disease.

ACKNOWLEDGMENTS

This work is supported by grants from the NIH-NCRR (P20RR018728), the NIEHS Superfund Basic Research Program (P42ES013660), the NIEHS Training Program in Environmental Pathology (T32ES007272, MAM) and the Flight Attendants Medical Research Institute.

References

  • 1.Barker DJ, Clark PM. Fetal undernutrition and disease in later life. Rev Reprod. 1997;2:105–112. doi: 10.1530/ror.0.0020105. [DOI] [PubMed] [Google Scholar]
  • 2.Hales CN, Barker DJ. Type 2 (non-insulin-dependent) diabetes mellitus: the thrifty phenotype hypothesis. Diabetologia. 1992;35:595–601. doi: 10.1007/BF00400248. [DOI] [PubMed] [Google Scholar]
  • 3.Jaenisch R. DNA methylation and imprinting: why bother? Trends Genet. 1997;13:323–329. doi: 10.1016/s0168-9525(97)01180-3. [DOI] [PubMed] [Google Scholar]
  • 4.Sood R, Zehnder JL, Druzin ML, Brown PO. Gene expression patterns in human placenta. Proc Natl Acad Sci U S A. 2006;103:5478–5483. doi: 10.1073/pnas.0508035103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Guo L, Choufani S, Ferreira J, Smith A, Chitayat D, Shuman C, Uxa R, Keating S, Kingdom J, Weksberg R. Altered gene expression and methylation of the human chromosome 11 imprinted region in small for gestational age (SGA) placentae. Dev Biol. 2008;320:79–91. doi: 10.1016/j.ydbio.2008.04.025. [DOI] [PubMed] [Google Scholar]
  • 6.Beattie JO. Transplacental alcohol intoxication. Alcohol Alcohol. 1986;21:163–166. [PubMed] [Google Scholar]
  • 7.Bush PG, Mayhew TM, Abramovich DR, Aggett PJ, Burke MD, Page KR. Maternal cigarette smoking and oxygen diffusion across the placenta. Placenta. 2000;21:824–833. doi: 10.1053/plac.2000.0571. [DOI] [PubMed] [Google Scholar]
  • 8.Bailey DN. Cocaine detection during toxicology screening of a university medical center patient population. J Toxicol Clin Toxicol. 1987;25:71–79. doi: 10.3109/15563658708992614. [DOI] [PubMed] [Google Scholar]
  • 9.Pastrakuljic A, Derewlany LO, Koren G. Maternal cocaine use and cigarette smoking in pregnancy in relation to amino acid transport and fetal growth. Placenta. 1999;20:499–512. doi: 10.1053/plac.1999.0418. [DOI] [PubMed] [Google Scholar]
  • 10.Rastogi S, Nandlike K, Fenster W. Elevated blood lead levels in pregnant women: identification of a high-risk population and interventions. J Perinat Med. 2007;35:492–496. doi: 10.1515/JPM.2007.131. [DOI] [PubMed] [Google Scholar]
  • 11.Mose T, Knudsen LE, Hedegaard M, Mortensen GK. Transplacental transfer of monomethyl phthalate and mono(2-ethylhexyl) phthalate in a human placenta perfusion system. Int J Toxicol. 2007;26:221–229. doi: 10.1080/10915810701352721. [DOI] [PubMed] [Google Scholar]
  • 12.Bird A. Perceptions of epigenetics. Nature. 2007;447:396–398. doi: 10.1038/nature05913. [DOI] [PubMed] [Google Scholar]
  • 13.Reamon-Buettner SM, Mutschler V, Borlak J. The next innovation cycle in toxicogenomics: environmental epigenetics. Mutat Res. 2008;659:158–165. doi: 10.1016/j.mrrev.2008.01.003. [DOI] [PubMed] [Google Scholar]
  • 14.Monk M, Boubelik M, Lehnert S. Temporal and regional changes in DNA methylation in the embryonic, extraembryonic and germ cell lineages during mouse embryo development. Development. 1987;99:371–382. doi: 10.1242/dev.99.3.371. [DOI] [PubMed] [Google Scholar]
  • 15.Kafri T, Ariel M, Brandeis M, Shemer R, Urven L, McCarrey J, Cedar H, Razin A. Developmental pattern of gene-specific DNA methylation in the mouse embryo and germ line. Genes Dev. 1992;6:705–714. doi: 10.1101/gad.6.5.705. [DOI] [PubMed] [Google Scholar]
  • 16.Razin A, Shemer R. DNA methylation in early development. Hum Mol Genet. 1995;4(Spec No):1751–1755. doi: 10.1093/hmg/4.suppl_1.1751. [DOI] [PubMed] [Google Scholar]
  • 17.Frommer M, McDonald LE, Millar DS, Collis CM, Watt F, Grigg GW, Molloy PL, Paul CL. A genomic sequencing protocol that yields a positive display of 5-methylcytosine residues in individual DNA strands. Proc Natl Acad Sci U S A. 1992;89:1827–1831. doi: 10.1073/pnas.89.5.1827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Herman JG, Graff JR, Myohanen S, Nelkin BD, Baylin SB. Methylation-specific PCR: a novel PCR assay for methylation status of CpG islands. Proc Natl Acad Sci U S A. 1996;93:9821–9826. doi: 10.1073/pnas.93.18.9821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Colella S, Shen L, Baggerly KA, Issa JP, Krahe R. Sensitive and quantitative universal Pyrosequencing methylation analysis of CpG sites. Biotechniques. 2003;35:146–150. doi: 10.2144/03351md01. [DOI] [PubMed] [Google Scholar]
  • 20.Uhlmann K, Brinckmann A, Toliat MR, Ritter H, Nurnberg P. Evaluation of a potential epigenetic biomarker by quantitative methyl-single nucleotide polymorphism analysis. Electrophoresis. 2002;23:4072–4079. doi: 10.1002/elps.200290023. [DOI] [PubMed] [Google Scholar]
  • 21.Eads CA, Danenberg KD, Kawakami K, Saltz LB, Blake C, Shibata D, Danenberg PV, Laird PW. MethyLight: a high-throughput assay to measure DNA methylation. Nucleic Acids Res. 2000;28:E32. doi: 10.1093/nar/28.8.e32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Mohn F, Weber M, Schubeler D, Roloff TC. Methylated DNA immunoprecipitation (MeDIP) Methods Mol Biol. 2009;507:55–64. doi: 10.1007/978-1-59745-522-0_5. [DOI] [PubMed] [Google Scholar]
  • 23.Mardis ER. ChIP-seq: welcome to the new frontier. Nat Methods. 2007;4:613–614. doi: 10.1038/nmeth0807-613. [DOI] [PubMed] [Google Scholar]
  • 24.Haycock PC, Ramsay M. Exposure of Mouse Embryos to Ethanol During Preimplantation Development: Effect on DNA Methylation in the H19 Imprinting Control Region. Biol Reprod. 2009 doi: 10.1095/biolreprod.108.074682. [DOI] [PubMed] [Google Scholar]
  • 25.Reik W, Walter J. Genomic imprinting: parental influence on the genome. Nat Rev Genet. 2001;2:21–32. doi: 10.1038/35047554. [DOI] [PubMed] [Google Scholar]
  • 26.Constancia M, Pickard B, Kelsey G, Reik W. Imprinting mechanisms. Genome Res. 1998;8:881–900. doi: 10.1101/gr.8.9.881. [DOI] [PubMed] [Google Scholar]
  • 27.Wood AJ, Oakey RJ. Genomic imprinting in mammals: emerging themes and established theories. PLoS Genet. 2006;2:e147. doi: 10.1371/journal.pgen.0020147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Swales AK, Spears N. Genomic imprinting and reproduction. Reproduction. 2005;130:389–399. doi: 10.1530/rep.1.00395. [DOI] [PubMed] [Google Scholar]
  • 29.Tilghman SM. The sins of the fathers and mothers: genomic imprinting in mammalian development. Cell. 1999;96:185–193. doi: 10.1016/s0092-8674(00)80559-0. [DOI] [PubMed] [Google Scholar]
  • 30.Badcock C, Crespi B. Imbalanced genomic imprinting in brain development: an evolutionary basis for the aetiology of autism. J Evol Biol. 2006;19:1007–1032. doi: 10.1111/j.1420-9101.2006.01091.x. [DOI] [PubMed] [Google Scholar]
  • 31.Isles AR, Holland AJ. Imprinted genes and mother-offspring interactions. Early Hum Dev. 2005;81:73–77. doi: 10.1016/j.earlhumdev.2004.10.006. [DOI] [PubMed] [Google Scholar]
  • 32.Tycko B, Morison IM. Physiological functions of imprinted genes. J Cell Physiol. 2002;192:245–258. doi: 10.1002/jcp.10129. [DOI] [PubMed] [Google Scholar]
  • 33.Chang AS, Moley KH, Wangler M, Feinberg AP, Debaun MR. Association between Beckwith-Wiedemann syndrome and assisted reproductive technology: a case series of 19 patients. Fertil Steril. 2005;83:349–354. doi: 10.1016/j.fertnstert.2004.07.964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.DeBaun MR, Niemitz EL, Feinberg AP. Association of in vitro fertilization with Beckwith-Wiedemann syndrome and epigenetic alterations of LIT1 and H19. Am J Hum Genet. 2003;72:156–160. doi: 10.1086/346031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Maher ER, Brueton LA, Bowdin SC, Luharia A, Cooper W, Cole TR, Macdonald F, Sampson JR, Barratt CL, Reik W, Hawkins MM. Beckwith-Wiedemann syndrome and assisted reproduction technology (ART) J Med Genet. 2003;40:62–64. doi: 10.1136/jmg.40.1.62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Uejima H, Lee MP, Cui H, Feinberg AP. Hot-stop PCR: a simple and general assay for linear quantitation of allele ratios. Nat Genet. 2000;25:375–376. doi: 10.1038/78040. [DOI] [PubMed] [Google Scholar]
  • 37.Bjornsson HT, Albert TJ, Ladd-Acosta CM, Green RD, Rongione MA, Middle CM, Irizarry RA, Broman KW, Feinberg AP. SNP-specific array-based allele-specific expression analysis. Genome Res. 2008;18:771–779. doi: 10.1101/gr.073254.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kerkel K, Spadola A, Yuan E, Kosek J, Jiang L, Hod E, Li K, Murty VV, Schupf N, Vilain E, Morris M, Haghighi F, Tycko B. Genomic surveys by methylation-sensitive SNP analysis identify sequence-dependent allele-specific DNA methylation. Nat Genet. 2008;40:904–908. doi: 10.1038/ng.174. [DOI] [PubMed] [Google Scholar]
  • 39.Turner BM. Cellular memory and the histone code. Cell. 2002;111:285–291. doi: 10.1016/s0092-8674(02)01080-2. [DOI] [PubMed] [Google Scholar]
  • 40.Kimura AP, Liebhaber SA, Cooke NE. Epigenetic modifications at the human growth hormone locus predict distinct roles for histone acetylation and methylation in placental gene activation. Mol Endocrinol. 2004;18:1018–1032. doi: 10.1210/me.2003-0468. [DOI] [PubMed] [Google Scholar]
  • 41.Jenuwein T, Allis CD. Translating the histone code. Science. 2001;293:1074–1080. doi: 10.1126/science.1063127. [DOI] [PubMed] [Google Scholar]
  • 42.Zhang Y, Reinberg D. Transcription regulation by histone methylation: interplay between different covalent modifications of the core histone tails. Genes Dev. 2001;15:2343–2360. doi: 10.1101/gad.927301. [DOI] [PubMed] [Google Scholar]
  • 43.Strahl BD, Allis CD. The language of covalent histone modifications. Nature. 2000;403:41–45. doi: 10.1038/47412. [DOI] [PubMed] [Google Scholar]
  • 44.Wang H, Cao R, Xia L, Erdjument-Bromage H, Borchers C, Tempst P, Zhang Y. Purification and functional characterization of a histone H3-lysine 4-specific methyltransferase. Mol Cell. 2001;8:1207–1217. doi: 10.1016/s1097-2765(01)00405-1. [DOI] [PubMed] [Google Scholar]
  • 45.Nishioka K, Chuikov S, Sarma K, Erdjument-Bromage H, Allis CD, Tempst P, Reinberg D. Set9, a novel histone H3 methyltransferase that facilitates transcription by precluding histone tail modifications required for heterochromatin formation. Genes Dev. 2002;16:479–489. doi: 10.1101/gad.967202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Rea S, Eisenhaber F, O'Carroll D, Strahl BD, Sun ZW, Schmid M, Opravil S, Mechtler K, Ponting CP, Allis CD, Jenuwein T. Regulation of chromatin structure by site-specific histone H3 methyltransferases. Nature. 2000;406:593–599. doi: 10.1038/35020506. [DOI] [PubMed] [Google Scholar]
  • 47.Yang L, Xia L, Wu DY, Wang H, Chansky HA, Schubach WH, Hickstein DD, Zhang Y. Molecular cloning of ESET, a novel histone H3-specific methyltransferase that interacts with ERG transcription factor. Oncogene. 2002;21:148–152. doi: 10.1038/sj.onc.1204998. [DOI] [PubMed] [Google Scholar]
  • 48.Schultz DC, Ayyanathan K, Negorev D, Maul GG, Rauscher FJ., 3rd SETDB1: a novel KAP-1-associated histone H3, lysine 9-specific methyltransferase that contributes to HP1-mediated silencing of euchromatic genes by KRAB zinc-finger proteins. Genes Dev. 2002;16:919–932. doi: 10.1101/gad.973302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Ogawa H, Ishiguro K, Gaubatz S, Livingston DM, Nakatani Y. A complex with chromatin modifiers that occupies E2F- and Myc-responsive genes in G0 cells. Science. 2002;296:1132–1136. doi: 10.1126/science.1069861. [DOI] [PubMed] [Google Scholar]
  • 50.Li E. Chromatin modification and epigenetic reprogramming in mammalian development. Nat Rev Genet. 2002;3:662–673. doi: 10.1038/nrg887. [DOI] [PubMed] [Google Scholar]
  • 51.Agalioti T, Lomvardas S, Parekh B, Yie J, Maniatis T, Thanos D. Ordered recruitment of chromatin modifying and general transcription factors to the IFN-beta promoter. Cell. 2000;103:667–678. doi: 10.1016/s0092-8674(00)00169-0. [DOI] [PubMed] [Google Scholar]
  • 52.Cosma MP, Tanaka T, Nasmyth K. Ordered recruitment of transcription and chromatin remodeling factors to a cell cycle- and developmentally regulated promoter. Cell. 1999;97:299–311. doi: 10.1016/s0092-8674(00)80740-0. [DOI] [PubMed] [Google Scholar]
  • 53.Rice JC, Allis CD. Histone methylation versus histone acetylation: new insights into epigenetic regulation. Curr Opin Cell Biol. 2001;13:263–273. doi: 10.1016/s0955-0674(00)00208-8. [DOI] [PubMed] [Google Scholar]
  • 54.Berger SL. Histone modifications in transcriptional regulation. Curr Opin Genet Dev. 2002;12:142–148. doi: 10.1016/s0959-437x(02)00279-4. [DOI] [PubMed] [Google Scholar]
  • 55.Lachner M, O'Sullivan RJ, Jenuwein T. An epigenetic road map for histone lysine methylation. J Cell Sci. 2003;116:2117–2124. doi: 10.1242/jcs.00493. [DOI] [PubMed] [Google Scholar]
  • 56.Calvanese V, Horrillo A, Hmadcha A, Suarez-Alvarez B, Fernandez AF, Lara E, Casado S, Menendez P, Bueno C, Garcia-Castro J, Rubio R, Lapunzina P, Alaminos M, Borghese L, Terstegge S, Harrison NJ, Moore HD, Brustle O, Lopez-Larrea C, Andrews PW, Soria B, Esteller M, Fraga MF. Cancer genes hypermethylated in human embryonic stem cells. PLoS ONE. 2008;3:e3294. doi: 10.1371/journal.pone.0003294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Lee RC, Feinbaum RL, Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75:843–854. doi: 10.1016/0092-8674(93)90529-y. [DOI] [PubMed] [Google Scholar]
  • 58.Reinhart BJ, Slack FJ, Basson M, Pasquinelli AE, Bettinger JC, Rougvie AE, Horvitz HR, Ruvkun G. The 21-nucleotide let-7 RNA regulates developmental timing in Caenorhabditis elegans. Nature. 2000;403:901–906. doi: 10.1038/35002607. [DOI] [PubMed] [Google Scholar]
  • 59.Lagos-Quintana M, Rauhut R, Lendeckel W, Tuschl T. Identification of novel genes coding for small expressed RNAs. Science. 2001;294:853–858. doi: 10.1126/science.1064921. [DOI] [PubMed] [Google Scholar]
  • 60.Lau NC, Lim LP, Weinstein EG, Bartel DP. An abundant class of tiny RNAs with probable regulatory roles in Caenorhabditis elegans. Science. 2001;294:858–862. doi: 10.1126/science.1065062. [DOI] [PubMed] [Google Scholar]
  • 61.Lee RC, Ambros V. An extensive class of small RNAs in Caenorhabditis elegans. Science. 2001;294:862–864. doi: 10.1126/science.1065329. [DOI] [PubMed] [Google Scholar]
  • 62.Griffiths-Jones S, Saini HK, van Dongen S, Enright AJ. miRBase: tools for microRNA genomics. Nucleic Acids Res. 2008;36:D154–158. doi: 10.1093/nar/gkm952. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Reinhart BJ, Bartel DP. Small RNAs correspond to centromere heterochromatic repeats. Science. 2002;297:1831. doi: 10.1126/science.1077183. [DOI] [PubMed] [Google Scholar]
  • 64.Ambros V, Lee RC, Lavanway A, Williams PT, Jewell D. MicroRNAs and other tiny endogenous RNAs in C. elegans. Curr Biol. 2003;13:807–818. doi: 10.1016/s0960-9822(03)00287-2. [DOI] [PubMed] [Google Scholar]
  • 65.Aravin AA, Hannon GJ, Brennecke J. The Piwi-piRNA pathway provides an adaptive defense in the transposon arms race. Science. 2007;318:761–764. doi: 10.1126/science.1146484. [DOI] [PubMed] [Google Scholar]
  • 66.Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116:281–297. doi: 10.1016/s0092-8674(04)00045-5. [DOI] [PubMed] [Google Scholar]
  • 67.Lee Y, Jeon K, Lee JT, Kim S, Kim VN. MicroRNA maturation: stepwise processing and subcellular localization. EMBO J. 2002;21:4663–4670. doi: 10.1093/emboj/cdf476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Smalheiser NR. EST analyses predict the existence of a population of chimeric microRNA precursor-mRNA transcripts expressed in normal human and mouse tissues. Genome Biol. 2003;4:403. doi: 10.1186/gb-2003-4-7-403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Cai X, Hagedorn CH, Cullen BR. Human microRNAs are processed from capped, polyadenylated transcripts that can also function as mRNAs. RNA. 2004;10:1957–1966. doi: 10.1261/rna.7135204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lee Y, Ahn C, Han J, Choi H, Kim J, Yim J, Lee J, Provost P, Radmark O, Kim S, Kim VN. The nuclear RNase III Drosha initiates microRNA processing. Nature. 2003;425:415–419. doi: 10.1038/nature01957. [DOI] [PubMed] [Google Scholar]
  • 71.Yi R, Qin Y, Macara IG, Cullen BR. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes Dev. 2003;17:3011–3016. doi: 10.1101/gad.1158803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lund E, Guttinger S, Calado A, Dahlberg JE, Kutay U. Nuclear export of microRNA precursors. Science. 2004;303:95–98. doi: 10.1126/science.1090599. [DOI] [PubMed] [Google Scholar]
  • 73.Hutvagner G, Zamore PD. A microRNA in a multiple-turnover RNAi enzyme complex. Science. 2002;297:2056–2060. doi: 10.1126/science.1073827. [DOI] [PubMed] [Google Scholar]
  • 74.Schwarz DS, Hutvagner G, Du T, Xu Z, Aronin N, Zamore PD. Asymmetry in the assembly of the RNAi enzyme complex. Cell. 2003;115:199–208. doi: 10.1016/s0092-8674(03)00759-1. [DOI] [PubMed] [Google Scholar]
  • 75.Khvorova A, Reynolds A, Jayasena SD. Functional siRNAs and miRNAs exhibit strand bias. Cell. 2003;115:209–216. doi: 10.1016/s0092-8674(03)00801-8. [DOI] [PubMed] [Google Scholar]
  • 76.Mourelatos Z, Dostie J, Paushkin S, Sharma A, Charroux B, Abel L, Rappsilber J, Mann M, Dreyfuss G. miRNPs: a novel class of ribonucleoproteins containing numerous microRNAs. Genes Dev. 2002;16:720–728. doi: 10.1101/gad.974702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Song JJ, Smith SK, Hannon GJ, Joshua-Tor L. Crystal structure of Argonaute and its implications for RISC slicer activity. Science. 2004;305:1434–1437. doi: 10.1126/science.1102514. [DOI] [PubMed] [Google Scholar]
  • 78.Yekta S, Shih IH, Bartel DP. MicroRNA-directed cleavage of HOXB8 mRNA. Science. 2004;304:594–596. doi: 10.1126/science.1097434. [DOI] [PubMed] [Google Scholar]
  • 79.Filipowicz W, Bhattacharyya SN, Sonenberg N. Mechanisms of post-transcriptional regulation by microRNAs: are the answers in sight? Nat Rev Genet. 2008;9:102–114. doi: 10.1038/nrg2290. [DOI] [PubMed] [Google Scholar]
  • 80.Lim LP, Lau NC, Garrett-Engele P, Grimson A, Schelter JM, Castle J, Bartel DP, Linsley PS, Johnson JM. Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature. 2005;433:769–773. doi: 10.1038/nature03315. [DOI] [PubMed] [Google Scholar]
  • 81.van Rooij E, Sutherland LB, Qi X, Richardson JA, Hill J, Olson EN. Control of stress-dependent cardiac growth and gene expression by a microRNA. Science. 2007;316:575–579. doi: 10.1126/science.1139089. [DOI] [PubMed] [Google Scholar]
  • 82.van Rooij E, Marshall WS, Olson EN. Toward microRNA-based therapeutics for heart disease: the sense in antisense. Circ Res. 2008;103:919–928. doi: 10.1161/CIRCRESAHA.108.183426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Mocellin S, Pasquali S, Pilati P. Oncomirs: from tumor biology to molecularly targeted anticancer strategies. Mini Rev Med Chem. 2009;9:70–80. doi: 10.2174/138955709787001802. [DOI] [PubMed] [Google Scholar]
  • 84.Chan JA, Krichevsky AM, Kosik KS. MicroRNA-21 is an antiapoptotic factor in human glioblastoma cells. Cancer Res. 2005;65:6029–6033. doi: 10.1158/0008-5472.CAN-05-0137. [DOI] [PubMed] [Google Scholar]
  • 85.Cho SJ, Yin XJ, Choi E, Lee HS, Bae I, Han HS, Yee ST, Kim NH, Kong IK. DNA methylation status in somatic and placenta cells of cloned cats. Cloning Stem Cells. 2007;9:477–484. doi: 10.1089/clo.2007.0003. [DOI] [PubMed] [Google Scholar]
  • 86.Vanselow J, Selimyan R, Furbass R. DNA methylation of placenta-specific Cyp19 promoters of cattle and sheep. Exp Clin Endocrinol Diabetes. 2008;116:437–442. doi: 10.1055/s-2008-1058083. [DOI] [PubMed] [Google Scholar]
  • 87.Zhang XY, Ehrlich KC, Wang RY, Ehrlich M. Effect of site-specific DNA methylation and mutagenesis on recognition by methylated DNA-binding protein from human placenta. Nucleic Acids Res. 1986;14:8387–8397. doi: 10.1093/nar/14.21.8387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Novakovic B, Rakyan V, Ng HK, Manuelpillai U, Dewi C, Wong NC, Morley R, Down T, Beck S, Craig JM, Saffery R. Specific tumour-associated methylation in normal human term placenta and first-trimester cytotrophoblasts. Mol Hum Reprod. 2008;14:547–554. doi: 10.1093/molehr/gan046. [DOI] [PubMed] [Google Scholar]
  • 89.Zhang HJ, Siu MK, Wong ES, Wong KY, Li AS, Chan KY, Ngan HY, Cheung AN. Oct4 is epigenetically regulated by methylation in normal placenta and gestational trophoblastic disease. Placenta. 2008;29:549–554. doi: 10.1016/j.placenta.2008.03.003. [DOI] [PubMed] [Google Scholar]
  • 90.Chelbi ST, Vaiman D. Genetic and epigenetic factors contribute to the onset of preeclampsia. Mol Cell Endocrinol. 2008;282:120–129. doi: 10.1016/j.mce.2007.11.022. [DOI] [PubMed] [Google Scholar]
  • 91.Novakovic B, Sibson M, Ng HK, Manuelpillai U, Rakyan VK, Down TA, Beck S, Fournier T, Evain-Brion D, Dimitriadis E, Craig JM, Morley R, Saffery R. Placenta specific methylation of the vitamin D 24-hydroxylase gene:Implications for feedback autoregulation of active vitamin D levels at the fetomaternal interface. J Biol Chem. 2009 doi: 10.1074/jbc.M809542200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Park BH, Kim YJ, Park JS, Lee HY, Ha EH, Min JW, Park HS. [Folate and homocysteine levels during pregnancy affect DNA methylation in human placenta] J Prev Med Public Health. 2005;38:437–442. [PubMed] [Google Scholar]
  • 93.Tycko B. Imprinted genes in placental growth and obstetric disorders. Cytogenet Genome Res. 2006;113:271–278. doi: 10.1159/000090842. [DOI] [PubMed] [Google Scholar]
  • 94.Lambertini L, Diplas AI, Lee MJ, Sperling R, Chen J, Wetmur J. A sensitive functional assay reveals frequent loss of genomic imprinting in human placenta. Epigenetics. 2008;3:261–269. doi: 10.4161/epi.3.5.6755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.McMinn J, Wei M, Schupf N, Cusmai J, Johnson EB, Smith AC, Weksberg R, Thaker HM, Tycko B. Unbalanced placental expression of imprinted genes in human intrauterine growth restriction. Placenta. 2006;27:540–549. doi: 10.1016/j.placenta.2005.07.004. [DOI] [PubMed] [Google Scholar]
  • 96.Morris AC, Beresford GW, Mooney MR, Boss JM. Kinetics of a gamma interferon response: expression and assembly of CIITA promoter IV and inhibition by methylation. Mol Cell Biol. 2002;22:4781–4791. doi: 10.1128/MCB.22.13.4781-4791.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Peyman JA, Nelson PJ, Hammond GL. HLA-DR genes are silenced in human trophoblasts and stimulation of signal transduction pathways does not circumvent interferon-gamma unresponsiveness. Transplant Proc. 1992;24:470–471. [PubMed] [Google Scholar]
  • 98.Zuckermann FA, Head JR. Expression of MHC antigens on murine trophoblast and their modulation by interferon. J Immunol. 1986;137:846–853. [PubMed] [Google Scholar]
  • 99.Chuang HC, Chang CW, Chang GD, Yao TP, Chen H. Histone deacetylase 3 binds to and regulates the GCMa transcription factor. Nucleic Acids Res. 2006;34:1459–1469. doi: 10.1093/nar/gkl048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Barad O, Meiri E, Avniel A, Aharonov R, Barzilai A, Bentwich I, Einav U, Gilad S, Hurban P, Karov Y, Lobenhofer EK, Sharon E, Shiboleth YM, Shtutman M, Bentwich Z, Einat P. MicroRNA expression detected by oligonucleotide microarrays: system establishment and expression profiling in human tissues. Genome Res. 2004;14:2486–2494. doi: 10.1101/gr.2845604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Pineles BL, Romero R, Montenegro D, Tarca AL, Han YM, Kim YM, Draghici S, Espinoza J, Kusanovic JP, Mittal P, Hassan SS, Kim CJ. Distinct subsets of microRNAs are expressed differentially in the human placentas of patients with preeclampsia. Am J Obstet Gynecol. 2007;196:e261–266. doi: 10.1016/j.ajog.2007.01.008. 261. [DOI] [PubMed] [Google Scholar]
  • 102.Zhu XM, Han T, Sargent IL, Yin GW, Yao YQ. Differential expression profile of microRNAs in human placentas from preeclamptic pregnancies vs normal pregnancies. Am J Obstet Gynecol. 2009 doi: 10.1016/j.ajog.2008.12.045. [DOI] [PubMed] [Google Scholar]
  • 103.Donker RB, Mouillet JF, Nelson DM, Sadovsky Y. The expression of Argonaute2 and related microRNA biogenesis proteins in normal and hypoxic trophoblasts. Mol Hum Reprod. 2007;13:273–279. doi: 10.1093/molehr/gam006. [DOI] [PubMed] [Google Scholar]
  • 104.Chim SS, Shing TK, Hung EC, Leung TY, Lau TK, Chiu RW, Lo YM. Detection and characterization of placental microRNAs in maternal plasma. Clin Chem. 2008;54:482–490. doi: 10.1373/clinchem.2007.097972. [DOI] [PubMed] [Google Scholar]
  • 105.Gilad S, Meiri E, Yogev Y, Benjamin S, Lebanony D, Yerushalmi N, Benjamin H, Kushnir M, Cholakh H, Melamed N, Bentwich Z, Hod M, Goren Y, Chajut A. Serum microRNAs are promising novel biomarkers. PLoS ONE. 2008;3:e3148. doi: 10.1371/journal.pone.0003148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Marsit CJ, Eddy K, Kelsey KT. MicroRNA responses to cellular stress. Cancer Res. 2006;66:10843–10848. doi: 10.1158/0008-5472.CAN-06-1894. [DOI] [PubMed] [Google Scholar]
  • 107.Callis TE, Wang DZ. Taking microRNAs to heart. Trends Mol Med. 2008;14:254–260. doi: 10.1016/j.molmed.2008.03.006. [DOI] [PubMed] [Google Scholar]

RESOURCES