Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jun 1.
Published in final edited form as: Curr Allergy Asthma Rep. 2012 Jun;12(3):211–220. doi: 10.1007/s11882-012-0257-4

Epigenetics, Asthma, and Allergic Diseases: A Review of the Latest Advancements

Stephanie Lovinsky-Desir 1, Rachel L Miller 2
PMCID: PMC3358775  NIHMSID: NIHMS376795  PMID: 22451193

Abstract

Environmental epigenetic regulation in asthma and allergic disease is an exciting area that has gained a great deal of scientific momentum in recent years. Environmental exposures, including prenatal maternal smoking, have been associated with asthma-related outcomes that may be explained by epigenetic regulation. In addition, several known allergy and asthma genes have been found to be susceptible to epigenetic regulation. We review the latest experimental and translational studies that have been published this past year in several areas, including 1) characterization of environmental asthma triggers that induce epigenetic changes, 2) characterization of allergic immune and regulatory pathways important to asthma that undergo epigenetic regulation, 3) evidence of active epigenetic regulation in asthma experimental models and the production of asthma biomarkers, 4) evidence of transmission of an asthma-related phenotype across multiple generations, and 5) “pharmaco-epigenetics.” The field has certainly advanced significantly in the past year.

Keywords: Epigenetics, Asthma, Allergy, Environment, DNA methylation, Histone modification, microRNA, T-helper cell, Regulatory T cell, Transgenerational, Pharmaco-epigenetics

Introduction

Recently, there have been numerous publications addressing the role of epigenetic regulation in complex diseases, and how environmental exposures may induce these molecular events. The most common epigenetic mechanisms include DNA methylation, histone modifications, and noncoding RNAs, all of which can affect gene transcription through effects on DNA structure and induction of gene silencing (Table 1). They may explain immune modulation induced by the environment even when we cannot measure the inciting environmental exposure itself. Epigenetic regulation also provides an attractive mechanism for explaining known associations of several prenatal exposures on the later development of various disorders in children and adults.

Table 1.

Mechanisms of epigenetic regulation

Site(s) of action Mechanism(s) of action Effect(s)
DNA methylation CpG islands (stretches of DNA >200 bp with >50% GC content) and other CpG regions (<200 bp) Addition of methyl group to the 5′ position of cytosines by DNMTs Turn off gene transcription
Histone modification Histone proteins (H2A, H2B, H3, and H4) that form the nucleosome core Chromatin remodeling by methylation, acetylation (via HAT, HDAC), phosphorylation, or ubiquitylation of histone tails Regulate accessibility of DNA to RNA polymerase II and transcription factors
Noncoding RNAs miRNAs (21–23 nucleotide long regulatory RNAs) Regulatory transcripts: binding of miRNA to the 3′ untranslated regions of mRNA Induce degradation of target mRNA

CpG, cytosine-phosphate-guanine, DNMT, DNA methyltransferase; HAT, histone acetyltransferases, HDAC, histone deacetylases; miRNA, microRNA

All these explanations fit well with why epigenetic regulation may be an important mechanism in asthma and allergic disease. Several prenatal exposures, with maternal smoking being the most described, have been associated with asthma-related outcomes in children [1, 2]. Allergies and asthma are more likely to develop in children if the mother has allergies or asthma compared with the father, suggesting the possibility of intrauterine programming [3, 4]. Moreover, the atopic phenotype characteristically is heterogeneous, in part related to varying individual environmental exposures, and can remit or be induced at various times in a person’s lifetime [57]. Furthermore, several known atopy and asthma genes have been found to be susceptible to epigenetic regulation, including genes important to T-effector pathways (interferon (IFN)-γ, interleukin (IL)4, IL13, IL17) [813], T-regulatory pathways (forkhead box P3 [FoxP3]) [14], and airway inflammation (arginase [ARG], inducible nitric oxide synthase [iNOS]) [15, 16]. Hence, the investigation into epigenetic regulation and allergic disease continues.

The scientific literature in recent years has benefited from the publication of numerous comprehensive review articles and call for more studies on epigenetic regulation in allergic disease and asthma [1719]. Nonetheless, it has been apparent that these raised many more questions than provided answers. The published experimental work, with few exceptions [15, 16, 20], mostly has comprised small observational studies and models in cell systems and animals [2123]. However, in the past year, exciting and elegant experimental studies and novel translational research works were published. Advances were made in many areas, including 1) better characterization of how several environmental asthma triggers induce epigenetic changes, 2) better characterization of how allergic immune pathways and regulatory pathways important to asthma undergo epigenetic regulation, 3) growing evidence of active epigenetic regulation in asthma experimental models and the production of asthma biomarkers, 4) experimental evidence of transmission of an asthma-related phenotype across multiple generations, and 5) “pharmaco-epigenetics.: This review discusses such key papers.

Environmental Asthma Triggers

Among the airborne pollutants, the one most associated with the induction of epigenetic changes is tobacco smoke. Exposure to cigarette smoke was shown to inhibit the expression of histone deacetylase (HDAC) in alveolar macrophages. Blocking deacetylation or removal of acetyl groups on the histone presumably permitted greater access of the DNA to RNA polymerases and transcription factors, augmented the inflammatory cytokine gene transcription, and thereby reduced the efficacy of anti-inflammatory therapy such as glucocorticoids in chronic obstructive lung disease [24]. More recently, experimental models and epidemiologic studies have demonstrated the ability of exposure to cigarette smoke to change patterns of DNA methylation and microRNA (miRNA) expression important to the transcription of genes implicated in apoptosis, proliferation, and cancer [25, 26]. Liu and colleagues [27•], in an informative experimental model using primary and immortalized human airway epithelial cells chronically exposed to cigarette smoke condensate, found that smoke condensate induced dose- and time-dependent histone alterations that were associated with decreased expression of DNA methyltransferase (DNMT)1 and increased expression of DNMT3b, enzymes that catalyze the transfer of methyl groups to DNA. They also found time-dependent demethylation of several specific genes implicated in carcinogenesis (eg, D4Z4, NBL2) and globally that were variably associated with changes in gene expression, suggesting a large array of possible epigenetic targets of cigarette smoke exposure [27•]. In addition, prenatal exposure to cigarette smoke in a pediatric cohort was associated with decreased DNA methylation globally, as indicated by reduced methylation of Alu sequences (the short interspersed nucleotide elements) and in CpG regions in the promoter of receptor tyrosine kinase (AXL), protein tyrosine phosphatase, receptor type (PTPRO), and other genes [28•].

Exposure to traffic-related pollutants, known triggers of allergies and asthma [29, 30], also has emerged as a potentially important inducer of epigenetic changes since Baccarelli and colleagues [20] found that measured levels of fine particulate matter and black carbon/soot were associated with time-dependent reductions in global methylation in a cohort of older adult participants in the Boston area Normative Aging Study. Ambient levels of another combustion traffic emission, namely polycyclic aromatic hydrocarbons (PAH), measured from backpacks worn by pregnant women were associated with asthma and allergy candidate gene-specific changes in DNA methylation [31]. In even more recent studies of PAH exposure, administration of benzo(a)pyrene (BaP), one of the PAHs measured in this last study, to cultures of immortalized HeLa cells, induced the early recruitment of several histone activation markers, followed by decreased DNMT1 protein levels, and globally reduced DNA methylation at CpG islands [32]. Nadeau and colleagues [33•] found while studying asthmatic children living in either the relatively polluted Fresno area (ie, airborne fine particulate matter concentrations that exceeded federal annual standards) versus cleaner Stanford areas of California that children with greater exposure to air pollution were more likely to have evidence of increased DNA methylation of the forkhead box P3 (FoxP3) transcription factor and impaired T-regulatory (Treg) function. Annual average PAH levels, estimated based on land use regression analyses that used actual PAH measures from a subset of homes, were correlated with the number of methylated CpG islands in the FoxP3 gene [33•]. Moreover, changes in the expression of several miRNAs have been reported in association with exposure to particulate matter. In one such study, Bollati and colleagues examined individual exposure to fine particulate matter and metals including chromium, lead, cadmium, arsenic, nickel, manganese, many of which have been implicated in allergies and asthma [34, 35]. They found in foundry workers from an electrical furnace steel plant that estimated individual levels of exposure to fine particulate matter and metals were associated with both upregulation and downregulation of an array of miRNAs important in inflammation and oxidative stress [36••].

Questions about the effect of diet during pregnancy on the subsequent risk of allergies or asthma in the offspring have been raised for years, with some data supporting the ability of Mediterranean diets, and supplemental fish oil and vitamin E, to provide some protection from risk [37, 38]. Given the presumption that folic acid supplementation is a source of methyl donors that could alter DNA methylation and thereby gene expression, its use has become the focus of several studies. This question, in part, was fueled by elegant murine models that demonstrated that a diet high in methyl donors during pregnancy was associated with a more pronounced allergic phenotype in the offspring. The allergic phenotype was associated with altered DNA methylation of several genes, including the T-cell suppressor runt related transcription factor 3 (RUNX3) [39]. Of course, this issue is of great public health concern given the established importance of folic acid supplementation in the prevention of neural tube defects in infants [40]. However, data on folic acid supplementation in humans and associated allergic disease have been mixed. While some found that prenatal folic acid supplementation was associated with more asthma, wheeze, and other respiratory problems in early-childhood [41, 42], these findings have been difficult to replicate [43]. This debate arguably continued this past year, when Haberg and colleagues [44], in follow-up studies now looking at nonfasting plasma folate levels in their cohort of pregnant mothers, reported a trend of increasing risk of asthma in children across quintiles of plasma folate levels during pregnancy. In contrast, Magdelijns et al. [45••] found that folic acid use during pregnancy was not associated with a greater risk of wheeze, asthma, or eczema. Intracellular folic acid measured in later pregnancy was associated with a reduced risk of childhood asthma. Therefore, the weight of current evidence is insufficient to recommend any change from the current practice of periconceptional folic acid supplementation to protect children from neural tube defects.

The role of vitamin D as an immunoregulatory agent has gained wide recognition in recent years. Indeed, higher intake of vitamin D during pregnancy has been associated with some protection from wheeze in children at age 5 years [46]. Modification of T-helper effector cell cytokine production and Treg cell function are some of the suggested mechanisms [47]. Evidence of epigenetic regulation of these processes thus far is limited, though in one recent study, the active metabolite, 1,25-dihydroxyvitamin D3, upregulated histone H4 acetylation and participated in other chromatin remodeling events important to the expression of proinflammatory genes [48].

Food allergens themselves have been found to be susceptible to epigenetic regulation. In one paper, the promoter of the peanut allergen Ara h 3 underwent changes in histone acetylation influencing its gene expression in early- and late-maturation embryos [49]. Clearly the implication of this research in relation to susceptibility to allergy is unclear but does suggest the potential wide scope of some of these processes.

Allergic Immune and Regulatory Pathways Important to Asthma

Growing fundamental cellular and experimental work continues to provide strong biological support for the premise that key pathways important to the allergic immune response, such as T-cell differentiation and Treg function, are susceptible to distinct epigenetic modifications. This includes DNA methylation and histone modifications that direct immune programming related to pro-allergic IL-4, IL-13 [9, 22, 50, 51], and counterregulatory IFN-γ production [8, 1012]. For example, we now are learning that pro-allergic T-helper type 2 (Th2) or counterregulatory Th1 cytokine production such as IFN-γ by human effector T cells can be flexible or plastic [52]. In many experimental settings, the induction of IFN-γ transcription is regulated primarily by DNA demethylation of CpG sites within the IFN-γ gene [8, 1012, 53]. New findings also have emerged detailing the induction of FoxP3, a Treg-specific transcription factor, and its dependence on DNA methylation and histone modifications [14, 54, 55]. As described below and in Table 2, major advances in the past year have been made in the further refinement/description of the role of DNA methylation, histone changes, and miRNA in these pathways, as well as better localization of the sites involved.

Table 2.

Epigenetic T-cell regulation and asthma- and allergy-related outcomes

Th pathway responses
Authors Gene(s) Epigenetic finding(s)
Koh et al. [58] Th2 Chromatin remodeling of the Th2 cytokine locus critical to systemic and airway inflammation
Grausenburger et al. [59] Th2 HDAC1 important to early Th2 cell differentiation and development of allergic airway inflammation
Sonkoly et al. [60] CTLA-4 miR-155 upregulated in atopic dermatitis and its overexpression associated with decreased CTLA-4
Lim et al. [61] Eotaxin Inhibition of HDAC increased IL-13–induced and CBP-mediated eotaxin expression
Kumar et al. [70] IL-13 Transfection of T cells with let-7 reduced IL-13 levels; intranasal administration of let-7 reduced airway allergic inflammation in mice
Polikepahad et al. [71•] Th2 Inhibition of let-7 miRNAs reduced allergic inflammation in mouse lung
Kim et al. [62••] STAT6 Inhibition of DNMT increased STAT6 expression
Tanaka et al. [63••] IL4, IL13 Targeted deletion of HS sites in the IL-4–IL-13 locus inhibited GATA-3 activation
Mucasa et al. [65] IL17 Th17 cell lineage is not stable and mediated epigenetically through chromatin remodeling
Treg pathway responses
Authors Gene(s) Epigenetic finding(s)
Lal and Bromberg [67•] FoxP3 Differential FoxP3 methylation in thymus derived nTreg versus peripheral TGF-β–induced Treg
Th and Treg pathway responses
Authors Gene(s) Epigenetic finding(s)
Janson et al. [57] Th1, Th2, Th17, Treg Epigenetic immune lineage analysis profiled CD4+ cells in rheumatoid arthritis, multiple sclerosis
Kim et al. [68] Th2, Treg Memory antigen-specific Th2 cells can redifferentiate into functional Tregs

CBP, cAMP-responsive element binding protein; CTLA, cytotoxic T-lymphocyte antigen; DNMT, DNA methyltransferases; FoxP3, forkhead box P3; GATA-3, trans-acting T-cell–specific transcription factor 3; HDAC, histone deacetylases; HS, DNase I-hypersensitive site; IL, interleukin; miRNA, microRNA; nTreg, natural Treg; STAT6, signal transducer and activator of transcription 6; Th, T helper; Treg, regulatory T cell

For example, Janson and colleagues [54] used the high-resolution Epigenetic Immune Lineage Analysis (EILA) in combination with several probes as a method to evaluate simultaneously Th1, Th2, Th17, and Treg commitment. The method involves profiling each lineage in an isolated population of human CD4+ T cells based on the specific CpG methylation status of signature cytokine and transcription factor loci. In examples described in the paper, they were able to apply the method to CD4+ cells from patients with two autoimmune diseases associated with imbalances of the Th1 and the Treg cell responses [56••], rheumatoid arthritis, and multiple sclerosis, and demonstrate specific CpG methylation patterns compatible with their Th effector populations. Their technique holds some interesting promise for future studies in the ability to define epigenomes relevant to a cell population and possibly many clinical diseases, including allergy and asthma [57].

One informative paper by Koh and colleagues [58] demonstrated the importance of chromatin remodeling of the Th2 cytokine locus (clustered on human chromosome 5, mouse chromosome 11) in murine models deficient for the conserved regulatory regions within the locus. These cLCR knockout mice experienced a loss of general H3 acetylation and histone H3-K4 methylation and demethylation of the DNA in the Th2 cytokine locus. This led to a marked reduction in the recruitment of eosinophils and lymphocytes in the airways, as well as decreased IgE levels and airway hyperreactivity [58]. Grausenburger et al. [59] showed in conditional T-cell–specific HDAC1 knockout mice that HDAC was important to the development of Th2 cytokine production and allergic airway inflammation. Using ChIP assays in nonactivated naïve CD4+ T cells, they also showed that HDAC was bound to the IL-4 gene locus, suggesting HDAC may be particularly important during early T-cell differentiation [59].

Interesting work was done on the function of miRNA expression on lesional and healthy skin of patients with atopic dermatitis. miR-155 was one of the most upregulated miRNA in atopic dermatitis lesions and was most expressed in the infiltrating immune cells. Additional experiments by the group showed that the miRNA may induce the dermatitis via inhibition of the negative T-cell regulator cytotoxic T lymphocyte–associated antigen (CTLA)-4 [60].

Major advances describing epigenetic regulation of the pro-allergic cytokine IL-13 were made this year as well. For example, Lim and colleagues [61] examined cAMP-responsive element binding protein (CBP) and chromatin remodeling in IL-13 pro-allergic chemokine regulation. Inhibition of HDAC (using culture of primary esophageal epithelial cells with trichostatin A) and induction of acetylated histone 3 increased IL-13–induced, and CBP-mediated, eotaxin-3 gene expression. Interestingly, in a sample of patients with eosinophilic esophagitis, they were able to document a greater level of acetylated histone 3 staining in the nuclei of the epithelial cells in their esophageal tissue [61]. The critical IL-4 signaling transcription factor STAT6 also appears susceptible to DNA methylation. This work was delineated in experiments by Kim and colleagues [62••], who treated activated human primary T cells isolated from healthy donors with a DNMT inhibitor (5-Aza-2-DC) and found that expression of STAT6 mRNA and protein were increased. In a recent study by Tanaka and colleagues [63••], a greater understanding was gained regarding chromosome modification of the IL-4 locus during Th2 differentiation. Their study demonstrated that GATA protein binding-3 (GATA-3) chromosomal modification of IL-4’s second intron is regulated by the DNase I-hypersensitive site (HS) 2 element. Targeted deletion of a series of HS sites in the IL4–IL13 locus led to functional impairment of epigenetic regulation of the IL4 locus in mice. Specifically, the deletion impaired acetylation and the onset of methylation by histone H3, resulting in failure to activate GATA-3 and abrogation of the allergic response [63••].

Th17 cells comprise a relatively recently discovered subset of T-helper cells producing IL-17 that provide host defense against extracellular bacteria, especially at the mucosa. Histone H3 acetylation and Lys-4 tri-methylation have been associated specifically with IL-17 and IL-17f gene promoters in Th17 cells [64]. More recent work this year by Mukasa and colleagues [65] focused on the stability versus plasticity of IL-17 epigenetic changes. They performed experiments of the DNase I HS and histone modifications of the clustered IL-17a–IL-17f and IFN-γ loci in naive Th1 and Th17 cells, as well as in Th17 cell precursors both prior to and following IL-12 versus transforming growth factor (TGF)-β stimulation. The experiments were intended to shift toward a Th1 phenotype (after IL-12) versus maintain the Th17 phenotype (after TGF-β). They observed permissive modifications across the IL-17a–IL-17f locus when cells were stimulated with TGF-β. However, STAT4- and Th1-specific T box transcription factor (T-bet) silenced the pro-IL17 RAR-related orphan receptor gamma (RORC) gene when the cells were stimulated with IL-12. The experiments suggest that Th17 cell lineage is plastic, and this instability is mediated epigenetically through reversal and remodeling of the chromatin structure [65].

In recent studies, Treg cell function and regulation of FoxP3 transcription have emerged as targets of epigenetic modifications. Some of the experimental work may have been inspired by the reported role of Tregs in the protection from atopy found in association with being born to a mother who farms [66]. Since then, the FoxP3 promoter has been shown to undergo both histone acetylation and DNA methylation affecting gene transcription (reviewed in [14]). The expression of the FoxP3 gene in the Treg promoter, and differentiation of naive T cells into Tregs, has been shown to be dependent on epigenetic regulation. FoxP3 methylation also distinguished natural Treg (nTregs), found in the thymus, from TGF-β–induced Tregs, found in the periphery [67−]. Most recently, Kim and colleagues [68], in elegant experiments, found that antigen-specific memory Th2 cells were able to redifferentiate into functional FoxP3+ T cells by TGF-β in the presence of all-trans retinoic acid and rapamycin. These new Tregs functioned as usual Tregs and were able to downregulate key transcription factors such as GATA-3 and IRF-4, suppress T-cell proliferation, and suppress pro-allergic cytokine production. Also, Th2 memory-derived converted Tregs adoptively transferred into Th2 memory-bearing recipient mice suppressed ongoing airway hyperreactivity, airway eosinophilia, and antigen-specific IgE responses. Also, they were able to home very efficiently to the airways [68].

In addition to advances in epigenetic regulation of T-cell lineage, new studies have examined dendritic cell (DC) immune responses. In attempt to identify epigenome-wide DNA methylation, Fedulov and Kobzik [69] isolated splenic DCs from their neonatal asthma–susceptible murine model. The splenic DCs were isolated from 14-day-old naïve neonatal mice that were from mothers who previously were sensitized to ovalbumin, and transferred these cells to healthy, nonsensitized mice. They found an increase in the overall methylation and increased allergen presentation in the sensitized DCs compared with DCs isolated from control mice that were not sensitized but were otherwise genetically and environmentally identical. These findings suggest that allergen induction in mothers can be passed along to offspring by epigenetic mechanisms [69].

Asthma Experimental Models and Asthma Biomarkers

The recent literature, using animal models for asthma and biomarkers as surrogates of the human disease processes, has brought us exciting new support for epigenetic regulation in asthma. Kumar and colleagues [70] first found that transfection of primary cultured T cells with let-7 (a highly conserved group of at least nine miRNAs known to be involved in Toll-like receptor 4 signaling) reduced IL-13 levels. Intranasal administration of let-7 reduced the allergic phenotype (airway inflammation, airway hyperresponsiveness, mucus metaplasia, and subepithelial brosis) in sensitized mice [70]. Let-7 miRNAs were tested further in a murine model for asthma by Polikepahad and colleagues [71]. Short RNAs and miRNAs were very enriched in both naïve and allergen-sensitized and challenged mouse lungs, with the most abundant ones belonging to the let-7 family. However, targeted inhibition of let-7 miRNA suppressed Th2 cytokine production, eosinophil recruitment to the lung, and airway hyperresponsiveness, suggesting that under some circumstances, and in this in vivo model, let-7 miRNAs may exert a proinflammatory role. The authors attributed these unexpected findings to the large (>800) repertoire of targets of let-7 miRNAs and possible secondary effects [71•].

As another example, Collision and colleagues [72] applied a murine model of allergic airway disease induced by house dust mite (HDM) allergen to new work on the contribution of several miRNAs. Interestingly, the authors identified significantly elevated levels of miR-145, miR-21, and let-7b in the HDM-induced allergic airways. All these specific miRNAs previously had been implicated in airway smooth muscle function, allergic inflammation, and airway epithelial cell function. Of the miRNAs identified, they found that the most significant proinflammatory role was played by miR-145 that, when specifically inhibited, suppressed allergic inflammation and airway remodeling. In additional experiments, both the suppression of miR-145 as well as treatment with the known anti-inflammatory dexamethasone significantly reduced the production of IL-5 and IL-13 from Th2 cells, though the effect of dexamethasone appeared to be more potent. These informative studies are among the few to identify altered regulation of specific miRNAs in allergic-mediated inflamed airways and point to potential future therapeutic targets directed at airway inflammation [72].

The first paper to look at DNA methylation of specific asthma genes and biomarkers of airway inflammation was published by Breton and colleagues [16] in 2011. Under the auspices of the Children’s Health Study, methylation levels of several CpG loci located in the promoter regions of the iNOS and ARG genes were compared with levels of fractional exhaled nitric oxide (FENO), an established asthma biomarker associated with airway inflammation and lung function decrements, among their asthmatic children. DNA methylation in the ARG, but not the iNOS gene, was associated inversely with FENO, suggesting a possible role of DNA methylation in the regulation of nitric oxide production [16].

In comparison, Isidoro-Garcia and colleagues [73] compared CpG methylation levels in a small cohort of allergic asthmatic patients with that of controls for the D prostanoid receptor (PTGDR) gene that mediates the production of PGD2. Indeed, PTGDR methylation levels were decreased at multiple CpG sites among patients compared with controls. One site of differential methylation was located at the same site as a recognized 613C>T single nucleotide polymorphism, possibly implicating genetic and epigenetic factors in the association with allergic asthma [73].

Transmission of an Asthma-Related Phenotype Across Multiple Generations

The idea that environmental exposures early in life may impact clinical outcomes is a fascinating concept. However, also intriguing is the notion that these epigenetically regulated genes may be passed from parent to offspring. Previously it was thought that epigenetic marks were erased upon passage through the germline. However, there was a huge paradigm shift more than 10 years ago that disproved that previously held belief in pivotal work involving agouti mice performed by Morgan et al. [74] and Waterland and Jirtle [75]. Since their discoveries, there is now emerging evidence that environmental factors may alter the germline of the epigenome, leading to transgenerational epigenetic inheritance [76]. A recent meta-analysis illustrated the stronger inheritance pattern of asthma in children of mothers versus fathers with history of asthma [4]. This work suggests that intrauterine events yield sustained effects on the child’s phenotype, possibly via passage of epigenetic marks through the germline. Recent studies have sought to delineate further these mechanisms.

Brand and colleagues demonstrated a microbial protective effect from asthma that appears to be inherited transgenerationally in a murine model. They studied Acinetobacter lwoffii, a farm-derived gram negative bacteria, to test whether this particular microbial exposure may be responsible for some of the asthma-protective effects of early farm exposure. Female BALB/c mice sensitized to ovalbumin (vs negative controls) were exposed to A. lwoffii (5 exposures during the 1.5 weeks before mating and every other day during pregnancy) and studies were performed on their offspring (who were not exposed to A. lwoffii). They found that stimulated T cells from the splenic mononuclear cells of these offspring had significantly decreased ability to produce Th2 cytokines such as IL-4, IL-5, and IL-13, while production of the counterregulatory IFN-γ remained intact, suggesting a blunted pro-allergic T-helper cell response. However, when the IFN-γ pathway was blocked selectively by treating the offspring with neutralizing anti–IFN-γ monoclonal antibody, Th2 cytokine production was restored [77]. These experiments suggest that prenatal A. lwoffii exposure upregulated Th1 activity, suppressed Th2 activity, and protected against the allergic inflammatory phenotype in the offspring.

Furthermore, our group hypothesized that combined prenatal in vivo exposure to Aspergillus fumigatus allergen and diesel exhaust particles was associated with changes in the asthma phenotype in the mouse offspring as well. Following sensitization via the airway route to A. fumigatus and mating, pregnant BALB/c mice were exposed to additional A. fumigatus and/or diesel exhaust particles. At age 9 to 10 weeks, the offspring were sensitized and challenged with A. fumigatus. Adult offspring from mice that were exposed to A. fumigatus or diesel exhaust particles during pregnancy experienced decreased instead of the anticipated increased IgE production. Adult offspring of mice that were exposed to both A. fumigatus and diesel exhaust particles during pregnancy experienced decreases in airway eosinophilia [78•].

Transgenerational passages of phenotypes and differential DNA methylation also recently were published in Arabidopsis thaliana plants. Schmitz and colleagues [79] identified 114,289 CpG single methylation polymorphisms and 2,485 CpG differentially methylated regions in the plant’s progeny that diverged over 30 generations from the ancestor to the descendent. Their findings suggest transgenerational methylation variation over multiple generations may affect gene transcription without changing the sequence of the genome [79]. Although we are closer in understanding that epigenetically modified genes may be differentially expressed by the offsprings, further studies are necessary to explore these relationships in human models.

Pharmaco-epigenetics

Epigenetic regulation also has emerged as a potential mechanism for the action of asthma-related pharmacologic therapies. For example, corticosteroids that have been used for decades to treat the inflammation present in acute and chronic asthma as well as chronic obstructive pulmonary disease (COPD) are now thought to act, at least in part, through histone acetylation. Corticosteroids bind intracellularly to glucocorticoid receptors, which become activated and bind to glucocorticoid response elements in the promoter regions of glucocorticoid-responsive genes. Corticosteroids exert their anti-inflammatory action by inducing histone acetylation of anti-inflammatory genes (eg, mitogen-activated protein kinase phosphatase-1 [MKP-1]), and by recruiting HDAC2 and inducing deacetylation of proinflammatory genes (eg, IL-8, NF-κB, activator protein-1 [AP-1]) [80]. Yet another proposed mechanism of corticosteroid action through induction of miRNA expression was studied in a recent paper by Williams and colleagues [81•] and found to be noncontributory. Five young adults with mild asthma underwent flexible bronchoscopy with bronchial biopsy before and after a 28-day course of inhaled corticosteroids. Although there was significant improvement in methacholine challenge following treatment, there was no change in the miRNA pattern following treatment with inhaled corticosteroids [81•]. Moreover, low concentrations of theophylline, another drug used for asthma and COPD, has been shown to reverse the effects of corticosteroid resistance by restoring HDAC2 activity, possibly via selective inhibition of phosphoinositide-3-kinase (PI3K)-δ and the phosphylation of downstream kinases [80].

Finally, PGI2 analogues have emerged as potential new asthma therapies because of their anti-inflammatory effects. Kou and colleagues [82], in a recent study, found that iloprost, a PGI2 analogue, enhanced H3 acetylation in the pro-allergic Th2 macrophage-derived chemokine promoter area and suppressed H3 acetylation, H3K4, and H3K36 trimethylation (addition of three methyl groups) in the counterregulatory Th1 I prostanoid-10 chemokine promoter area. These findings suggest that the PGI2 analogues in fact may increase Th2 pro-allergic inflammation via histone modifications of chemokines [82]. They also raise the possibility that epigenetic regulation of airway inflammation via histone modification of PGI2 analogues may have benefit in patients with steroid-nonresponsive diseases such as COPD or refractory asthma.

Conclusions

Recent publications shed a great deal of light on several important epigenetic responses to environmental exposures, epigenetically regulated molecular pathways, and experimental models of allergy and asthma. These new insights open doors for new approaches toward prevention of disease by targeted improvements in the environment and by development of novel drugs. Questions remain regarding the time course of these epigenetic mechanisms in vivo as well as the sustainability of their effects. The emerging evidence that some of these epigenetic effects may be transferred across multiple generations highlights how far-reaching these effects and novel therapeutic approaches can be.

Acknowledgments

Dr. Lovinsky-Desir has received grant support from the National Institutes of Health and the T32 Training Grant.

Dr. Miller has received grant support from the National Institutes of Health, the Environmental Protection Agency, and the Thrasher Foundation.

Footnotes

Disclosure No potential conflicts of interest relevant to this article were reported.

Contributor Information

Stephanie Lovinsky-Desir, Email: stl2003@nyp.org, Division of Pediatric Pulmonary, Columbia University College of Physicians and Surgeons, 630 West 168th Street, New York, NY 10032, USA.

Rachel L. Miller, Email: rlm14@columbia.edu, Columbia University College of Physicians and Surgeons, PH8E-101; 630 West 168th Street, New York, NY, 10032 USA

References

  • 1.Alati R, Al-Mamun A, O’Callaghan M, Najman JM, Williams GM. In utero and postnatal maternal smoking and asthma in adolescence. Epidemiology. 2006;17(2):138–44. doi: 10.1097/01.ede.0000198148.02347.33. [DOI] [PubMed] [Google Scholar]
  • 2.Eisner MD. Environmental tobacco smoke exposure and pulmonary function among adults in NHANES III: impact on the general population and adults with current asthma. Environ Health Perspect. 2002;110(8):765–70. doi: 10.1289/ehp.02110765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Litonjua A, Carey VJ, Burge HA, Weiss ST, Gold DR. Parental history and the risk for childhood asthma. Does mother confer more risk than father? Am J Respir Crit Care Med. 1998;158(1):176–81. doi: 10.1164/ajrccm.158.1.9710014. [DOI] [PubMed] [Google Scholar]
  • 4.Lim RH, Kobzik L, Dahl M. Risk for asthma in offspring of asthmatic mothers versus fathers: A Meta-Analysis. PLoS One. 2010;5(4):e10134. doi: 10.1371/journal.pone.0010134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Mandhane PJ, Greene JM, Cowan JO, Taylor DR, Sears MR. Sex differences in factors associated with childhood- and adolescent-onset wheeze. Am J Respir Crit Care Med. 2005;172:45–54. doi: 10.1164/rccm.200412-1738OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.deMarco R, Marcon A, Jarvis D. Prognostic factors of asthma severity: A 9-year international prospective cohort study. J Allergy and Clin Immunol. 2006;117:1249–56. doi: 10.1016/j.jaci.2006.03.019. [DOI] [PubMed] [Google Scholar]
  • 7.Burgess J, Matheson M, Gurrin L, Byrnes G, Adams K, Wharton C, et al. Factors influencing asthma remission: a longitudinal study from childhood to middle age. Thorax. 2011;66(6):508–13. doi: 10.1136/thx.2010.146845. [DOI] [PubMed] [Google Scholar]
  • 8.Liu J, Ballaney M, Al-alem U, Quan C, Jin X, Perera F, et al. Combined inhaled diesel exhaust particles and allergen exposure alter methylation of T helper genes and IgE production in vivo. Toxicological Sciences. 2008;102(1):76–81. doi: 10.1093/toxsci/kfm290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Webster RB, Rodriguez Y, Klimecki W, Vercelli D. The human IL-13 locus in neonatal CD4+ T cells is refractory to the acquisition of a repressive chromatin architecture. J Biol Chem. 2007;282(1):700–9. doi: 10.1074/jbc.M609501200. [DOI] [PubMed] [Google Scholar]
  • 10.White GP, Hollams EM, Yerkovich ST, Bosco A, Holt BJ, Bassami MR, et al. CpG methylation patterns in the IFNg promoter in naive T cells: Variations during Th1 and Th2 differentiation and between atopics and non-atopics. Pediatr Allergy Immunol. 2006;17:557–64. doi: 10.1111/j.1399-3038.2006.00465.x. [DOI] [PubMed] [Google Scholar]
  • 11.Jones B, Chen J. Inhibition of IFN-g transcription by site-specific methylation during T helper cell development. The EMBO Journal. 2006;25:2443–52. doi: 10.1038/sj.emboj.7601148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Schoenborn J, Dorschner M, Sekimata M, Santer D, Shnyreva M, Fitzpatrick D, et al. Comprehensive epigenetic profiling identifies multiple distal regulatory elements directing transcription of the gene encoding interferon-gamma. Nat Immunol. 2007;8:732–42. doi: 10.1038/ni1474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Akimzhanov A, Yang X, Dong C. Chromatin remodeling of interleukin-17 (IL-17)-IL-17F cytokine gene locus during inflammatory helper T cell differentiation. J Biol Chem. 2007;282(9):5969–72. doi: 10.1074/jbc.C600322200. [DOI] [PubMed] [Google Scholar]
  • 14.Huehn J, Polansky JK, Hamann A. Epigenetic control of FOXP3 expression: the key to a stable regulatory T-cell lineage? Nat Rev Immunol. 2009;9(2):83–9. doi: 10.1038/nri2474. [DOI] [PubMed] [Google Scholar]
  • 15.Tarantini L, Bonzini M, Apostoli P, Pegoraro V, Bollati V, Marinelli B, et al. Effects of particulate matter on genomic DNA methylation content and iNOS promoter methylation. Environ Health Perspect. 2009;117:217–22. doi: 10.1289/ehp.11898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Breton CV, Byun H-M, Wang X, Salam MT, Siegmund K, Gilliland FD. DNA methylation in the ARG-NOS pathway is associated with exhaled nitric oxide in asthmatic children. American J of Respiratory and Critical Care Medicine. 2011;184(2):191–97. doi: 10.1164/rccm.201012-2029OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Miller RL, Ho S. Environmental epigenetics and asthma: current concepts and call for studies. Am J Respir Crit Care Med. 2008;177:567–73. doi: 10.1164/rccm.200710-1511PP. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Kuriakose JS, Miller RL. Environmental epigenetics and allergic diseases: recent advances. Clinical and Experimental Allergy. 2010;40(11):1602–10. doi: 10.1111/j.1365-2222.2010.03599.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Martino D, Prescott S. Silent mysteries: Epigenetic paradigms could hold the key to conquering the epidemic of allergy and immune disease. Allergy. 2010;65(1):7–15. doi: 10.1111/j.1398-9995.2009.02186.x. [DOI] [PubMed] [Google Scholar]
  • 20.Baccarelli A, Wright RO, Bollati V, Tarantini L, Litonjua AA, Suh HH, et al. Rapid DNA methylation changes after exposure to traffic particles. American J Respiratory and Critical Care Medicine. 2009;179:572–8. doi: 10.1164/rccm.200807-1097OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Cao D, Bromberg PA, Samet JM. COX-2 expression induced by diesel particles involves chromatin modification and degradation of HDAC1. Am J Respir Cell Mol Biol. 2007;37(2):232–9. doi: 10.1165/rcmb.2006-0449OC. [DOI] [PubMed] [Google Scholar]
  • 22.Kwon N-H, Kim J-S, Lee J-Y, Oh M-J, Choi D-C. DNA methylation and the expression of IL-4 and IFN-γ promoter genes in patients with bronchial asthma. J Clin Immunol. 2008;28:139–46. doi: 10.1007/s10875-007-9148-1. [DOI] [PubMed] [Google Scholar]
  • 23.Su R, Becker A, Kozyrskyj A, Hayglass K. Epigenetic regulation of established human type 1 versus type 2 cytokine responses. J Allergy and Clin Immunol. 2008;121(1):57–63. doi: 10.1016/j.jaci.2007.09.004. [DOI] [PubMed] [Google Scholar]
  • 24.Ito K, Lim S, Caramori G, Chung KF, Barnes PJ, Adcock IM. Cigarette smoking reduces histone deacetylase 2 expression, enhances cytokine expression, and inhibits glucocorticoid actions in alveolar macrophages. FASEB J. 2001;15(6):1110–2. [PubMed] [Google Scholar]
  • 25.Izzotti A, Calin G, Arrigo P, Steele V, Croce C, Flora SD. Downregulation of microRNA expression in the lungs of rats exposed to cigarette smoke. FASEB J. 2009;23(3):806–12. doi: 10.1096/fj.08-121384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Wilhelm-Benartzi C, Christensen B, Koestler D, Houseman EA, Schned A, Karagas M, et al. Association of secondhand smoke exposures with DNA methylation in bladder carcinomas. Cancer Causes Control. 2011;22(8):1205–13. doi: 10.1007/s10552-011-9788-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27•.Liu F, Killian J, Yang M, Walker R, Hong J, Zhang M, et al. Epigenomic alterations and gene expression profiles in respiratory epithelia exposed to cigarette smoke condensate. Oncogene. 2010;29:3650–64. doi: 10.1038/onc.2010.129. This is one of the few studies of airway cells following environmental exposures. Several epigenetic alterations induced by exposure to cigarette smoke condensate were identified. There was also time-dependent demethylation at several sites implicated in carcinogenesis in the lung. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28•.Breton CV, Byun H-M, Wenten M, Pan F, Yang A, Gilliland FD. Prenatal tobacco smoke exposure affects global and gene-specific DNA methylation. American J Respiratory and Critical Care Medicine. 2009;180:462–7. doi: 10.1164/rccm.200901-0135OC. This was the first paper to look at DNA methylation of specific asthma genes and biomarkers of airway inflammation. The authors found that DNA methylation in the ARG gene was associated inversely with FENO, a marker of airway inflammation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Patel MM, Miller RL. Air pollution and childhood asthma: recent advances and future directions. Current Opinions in Pediatrics. 2009;21:235–42. doi: 10.1097/MOP.0b013e3283267726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Morgenstern V, Zutavern A, Cyrys J, Brockow I, Koletzko S, Krämer U, et al. Atopic diseases, allergic sensitization, and exposure to traffic-related air pollution in children. Amer J Respir Crit Care Med. 2008;177(12):1331–37. doi: 10.1164/rccm.200701-036OC. [DOI] [PubMed] [Google Scholar]
  • 31.Perera F, Tang WY, Herbstman J, Tang D, Levin L, Miller R, et al. Relation of DNA methylation of 5′-CpG island of ACSL3 to transplacental exposure to airborne polycyclic aromatic hydrocarbons and childhood asthma. PLoS One. 2009;4(2):e4488. doi: 10.1371/journal.pone.0004488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Teneng I, Montoya-Durango DE, Quertermous JL, Lacy ME, Ramos KS. Reactivation of L1 retrotransposon by benzo(a)pyrene involves complex genetic and epigenetic regulation. Epigenetics. 2011;6(3):355–67. doi: 10.4161/epi.6.3.14282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33•.Nadeau K, McDonald-Hyman C, Noth E, Pratt B, Hammond S, Balmes J, et al. Ambient air pollution impairs regulatory T-cell function in asthma. J Allergy and Clin Immunol. 2010;126(4):845–52. doi: 10.1016/j.jaci.2010.08.008. This study found decreased Treg function in asthmatics living in high-pollution neighborhoods. There was also increased methylation and decreased expression of the FoxP3 gene locus (important in Treg expression) in the exposed asthmatics, suggesting an important link between Treg, FoxP3, and asthma severity in high-pollution neighborhoods. [DOI] [PubMed] [Google Scholar]
  • 34.Patel MM, Hoepner L, Garfinkel R, Chillrud S, Reyes A, Quinn JW, et al. Ambient metals, elemental carbon, and wheeze and cough in New York City children through 24 months of age. Am J Respir Crit Care Med. 2009;180(11):1107–13. doi: 10.1164/rccm.200901-0122OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Hamada K, Suzaki Y, Leme A, Ito T, Miyamoto K, Kobzik L, et al. Exposure of pregnant mice to an air pollutant aerosol increases asthma susceptibility in offspring. J Toxicol Environ Health. 2007;70(8):688–95. doi: 10.1080/15287390600974692. [DOI] [PubMed] [Google Scholar]
  • 36••.Bollati V, Marinelli B, Apostoli P, Bonzini M, Nordio F, Hoxha M, et al. Exposure to metal-rich particulate matter modifies the expression of candidate microRNAs in peripheral blood leukocytes. Environ Health Perspect. 2010;118:763–8. doi: 10.1289/ehp.0901300. The authors measured specific miRNAs implicated in inflammation and oxidative stress in electric furnace steel plant workers at baseline and after 3 days of work-related exposure to particulate matter. Their important work found that levels of exposure to particulate matter were associated with upregulation and downregulation of these miRNAs. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Devereux G, Turner SW, Craig LC, McNeill G, Martindale S, Harbour PJ, et al. Low maternal vitamin E intake during pregnancy is associated with asthma in 5-year-old children. Am J Respir Crit Care Med. 2006;174(5):499–507. doi: 10.1164/rccm.200512-1946OC. [DOI] [PubMed] [Google Scholar]
  • 38.Salem M, Li Y, Langholz B, Gilliland F. Maternal fish consumption during pregnancy and risk of early childhood asthma. J Asthma. 2005;42:513–18. doi: 10.1081/JAS-67619. [DOI] [PubMed] [Google Scholar]
  • 39.Hollingsworth JW, Maruoka S, Boon K, Garantziotis S, Li Z, Tomfohr J, et al. In utero supplementation with methyl-donors enhances allergic airway disease in mice. Journal of Clinical Investigation. 2008;118(10):3462–9. doi: 10.1172/JCI34378. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 40.Miller RL. Prenatal maternal diet and asthma risk. The Journal of Clinical Investigation. 2008;118(10):3265–68. doi: 10.1172/JCI37171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Håberg SE, London SJ, Stigum H, Nafstad P, Nystad W. Folic acid supplements in pregnancy and early childhood respiratory health. Arch Dis Child. 2009;94:180–4. doi: 10.1136/adc.2008.142448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Whitrow M, Moore V, Rumbold A, Davies M. Effect of supplemental folic acid in pregnancy on childhood asthma: a prospective birth cohort study. Am J Epidemiol. 2009;170:1486–93. doi: 10.1093/aje/kwp315. [DOI] [PubMed] [Google Scholar]
  • 43.Matsui E, Matsui W. Higher serum folate levels are associated with a lower risk of atopy and wheeze. J Allergy Clin Immunol. 2009;123(6):1253–9. doi: 10.1016/j.jaci.2009.03.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Håberg SE, London S, Nafstad P, Nilsen R, Ueland P, Vollset S, et al. Maternal folate levels in pregnancy and asthma in children at age 3 years. J Allergy and Clin Immunol. 2010;127(1):262–4. doi: 10.1016/j.jaci.2010.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45••.Magdelijns FH, Mommers M, Penders J, Smits L, Thijs C. Folic acid use in pregnancy and the development of atopy, asthma, and lung function in childhood. Pediatrics. 2011;128:e135–e44. doi: 10.1542/peds.2010-1690. This study of prenatal folic acid supplementation and the risk of asthma-related outcomes (wheeze, atopy) found no association between the two. In fact, higher maternal serum folic acid levels were associated with decreased risk of asthma. [DOI] [PubMed] [Google Scholar]
  • 46.Devereux G, Litonjua A, Turner SW, Craig LC, McNeill G, Martindale S, et al. Maternal vitamin D intake during pregnancy and early childhood wheezing. Am J Clin Nutr. 2007;85(3):853–59. doi: 10.1093/ajcn/85.3.853. [DOI] [PubMed] [Google Scholar]
  • 47.Paul G, Brehm J, Alcorn J, Holguin F, Aujla S, Celedón J. Vitamin D and asthma. American Journal of Resipiratory and Critical Care Medicine. 2012 doi: 10.1164/rccm.201108-1502CI. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Sundar IK, Rahman J. Vitamin D and suscpetibility of chronic lung diseases: role of epiegentics. Frontiers in Pharmacology. 2011;2:1–10. doi: 10.3389/fphar.2011.00050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Fu G, Zhong Y, Li C, Li Y, Lin X, Liao B, et al. Epigenetic regulation of peanut allergen gene Ara h 3 in developing embryos. Planta. 2010;231(5):1049–60. doi: 10.1007/s00425-010-1111-3. [DOI] [PubMed] [Google Scholar]
  • 50.Lee DU, Agarwal S, Rao A. Th2 lineage commitment and efficient IL-4 production involves extended demethylation of the IL-4 gene. Immunity. 2002;16:649–60. doi: 10.1016/s1074-7613(02)00314-x. [DOI] [PubMed] [Google Scholar]
  • 51.Lee G, Fields P, Griffin T, Flavell R. Regulation of the Th2 cytokine locus by a locus control region. Immunity. 2003;19:145–53. doi: 10.1016/s1074-7613(03)00179-1. [DOI] [PubMed] [Google Scholar]
  • 52.Bluestone JA, Mackay CR, O’Shea JJ, Stockinger B. The functional plasticity of T cell subsets. Nature Reviews/Immunology. 2009;9:811–6. doi: 10.1038/nri2654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Vuillermin PJ, Ponsonby A-L, Saffery R, Tang ML, Ellis JA, Sly P, et al. Microbial exposure, interferon gamma gene demethylation in naïve T-cells, and the risk of allergic disease. Allergy. 2009;64:348–53. doi: 10.1111/j.1398-9995.2009.01970.x. [DOI] [PubMed] [Google Scholar]
  • 54.Janson P, Winerdal M, Winqvist O. At the crossroads of T helper lineage commitment-epigenetics points the way. BIochim Biophys Acta. 2009;1790(9):906–19. doi: 10.1016/j.bbagen.2008.12.003. [DOI] [PubMed] [Google Scholar]
  • 55.Toker A, Huehn J. To be or not to be a T reg cell: Lineage decisions controlled by epigenetic mechanisms. Science Signaling. 2011;4(158):e4. doi: 10.1126/scisignal.2001783. [DOI] [PubMed] [Google Scholar]
  • 56••.Leung S, Liu X, Fang L, Chen X, Guo T, Zhang J. The cytokine milieu in the interplay of pathogenic Th1/Th17 cells and regulatory T cells in autoimmune disease. Cell Mo Immunol. 2010;7(3):182–89. doi: 10.1038/cmi.2010.22. Following ovalbumin challenge in the cLCR knockout mice, there was reduction in systemic and airway inflammation, hyperresponsiveness, and mucus production. This study sheds light on the significance of chromatin remodeling in the Th2 LCR and the development of allergic asthma. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Janson PCJ, Linton LB, Bergman EA, Marits P, Eberhardson M, Piehl F, et al. Profiling of CD4+ T cells with epigenetic immune lineage analysis. J Immunol. 2011;186:92–102. doi: 10.4049/jimmunol.1000960. [DOI] [PubMed] [Google Scholar]
  • 58.Koh BH, Hwang SS, Kim JY, Lee W, Kang M-J, Lee CG, et al. Th2 LCR is essential for regulation of Th2 cytokine genes and for pathogenesis of allergic asthma. PNAS. 2010;107(23):10614–19. doi: 10.1073/pnas.1005383107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Grausenburger R, Bilic I, Boucheron N, et al. Conditional deletion of histone deacetylase 1 in T cells leads to enhanced airway inflammation and increased Th2 cytokine production. J Immunol. 2010;185:3489–97. doi: 10.4049/jimmunol.0903610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Sonkoly E, Janson P, Majuri M-L, Savinko T, Fyhrquist N, Eidsmo L, et al. MiR-155 is overexpressed in patients with atopic dermatitis and modulates T-cell proliferative responses by targeting cytotoxic T lymphocyte–associated antigen 4. J Allergy and Clin Immunol. 2010;126:581–89. doi: 10.1016/j.jaci.2010.05.045. [DOI] [PubMed] [Google Scholar]
  • 61.Lim EJ, Lu TX, Blanchard C, Rothenberg ME. Epigenetic regulation of the IL-13-induced human eotaxin-3 gene by CREB-binding protein-mediated histone 3 acetylation. J of Biological Chemistry. 2011;286(15):13193–204. doi: 10.1074/jbc.M110.210724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62••.Kim E-G, Shin H-J, Lee CG, Park H-Y, Kim Y-K, Park H-W, et al. DNA methylation and not allelic variation regulates STAT6 expression in human T cells. Clin Exp Med. 2010;10:143–52. doi: 10.1007/s10238-009-0083-8. This is one of the first papers to elucidate the mechanisms of epigenetic regulation of STAT6, a transcription factor thought to be important in Th2 cell differentiation. [DOI] [PubMed] [Google Scholar]
  • 63••.Tanaka S, Motomura Y, Suzuki Y, Yagi R, Inoue H, Miyatake S, et al. The enhancer HS2 critically regulates GATA-3-mediated Il4 transcription in TH2 cells. Nature Immunol. 2010;12(1):77–86. doi: 10.1038/ni.1966. The authors studied chromosomal regulation of IL-4 production by the transcription factor GATA-3. These findings suggest that DNase I-hypersensitive site 2 is an important target in the regulation of chromosomal modification of the IL-4 locus. [DOI] [PubMed] [Google Scholar]
  • 64.Akimzhanov A, Yang X, Dong C. Chromatin remodeling of interleukin-17 (IL-17)-IL-17F cytokine gene locus during inflammatory helper T cell differentiation. J Biol Chem. 2007;282(9):5969–72. doi: 10.1074/jbc.C600322200. [DOI] [PubMed] [Google Scholar]
  • 65.Mukasa R, Balasubramani A, Lee YK, Whitley SK, Weaver BT, Shibata Y, et al. Epigenetic instability of cytokine and transcription factor gene loci underlies plasticity of the T helper 17 cell lineage. Immunity. 2010;32:616–27. doi: 10.1016/j.immuni.2010.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Schaub B, Liu J, Hoppler S, Schleich I, Huehn J, Olek S, et al. Maternal farm exposure modulates neonatal immune mechanisms through regulatory T cells. J Allergy and Clin Immunol. 2009;123:774–82. doi: 10.1016/j.jaci.2009.01.056. [DOI] [PubMed] [Google Scholar]
  • 67•.Lal G, Bromberg JS. Epigenetic mechanisms of regulation of Foxp3 expression. Blood. 2009;114:3727–35. doi: 10.1182/blood-2009-05-219584. The authors offer an excellent summary of the different epgenetic markers of FoxP3 and the importance in Treg expression and suppression. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Kim B-S, Kim I-K, Park Y-J, Kim Y-S, Kim Y-J, Chang W-S, et al. Conversion of Th2 memory cells into Foxp3+ regulatory T cells suppressing Th2-mediated allergic asthma. PNAS. 2010;107(19):8742–47. doi: 10.1073/pnas.0911756107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Fedulov AV, Kobzik L. Allergy risk Is mediated by dendritic cells with congenital epigenetic ehanges. Am J Respir Cell Mol Biol. 2011;44:285–92. doi: 10.1165/rcmb.2009-0400OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Kumar M, Ahmad T, Sharma A, Mabalirajan U, Kulshreshtha A, Agrawal A, et al. Let-7 microRNA–mediated regulation of IL-13 and allergic airway inflammation. J Allergy Clin Immunol. 2011;128(5):1077–85. doi: 10.1016/j.jaci.2011.04.034. [DOI] [PubMed] [Google Scholar]
  • 71•.Polikepahad S, Knight JM, Naghavi AO, Oplt T, et al. Proinflammatory role for let-7 microRNAS in experimental asthma J of Biological Chemistry 20102853930139–49.Contrary to the authors’ hypothesis, inhibition of let-7 led to reduction in allergic cytokine production, suggesting significant complexities in these pathways. Nonetheless, miRNAs, namely the abdundant let-7, appear to play an important role in allergic lung disease. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Collison A, Mattes J, Plank M, Foster PS. Inhibition of house dust mite–induced allergic airways disease by antagonism of microRNA-145 is comparable to glucocorticoid treatment. J Allergy and Clin Immunol. 2011;128:160–67. doi: 10.1016/j.jaci.2011.04.005. [DOI] [PubMed] [Google Scholar]
  • 73.Isidoro-Garcıa M, Sanz C, Garcıa-Solaesa V, Pascual M, Pescador DB, Lorente F, et al. PTGDR gene in asthma: a functional, genetic, and epigenetic study. Allergy. 2011 doi: 10.1111/j.1398-9995.2011.02685.x. [DOI] [PubMed] [Google Scholar]
  • 74.Morgan H, Sutherland H, Martin D, Whitelaw E. Epigenetic inheritance at the agouti locus in the mouse. Nat Genetics. 1999;23(3):314–8. doi: 10.1038/15490. [DOI] [PubMed] [Google Scholar]
  • 75.Waterland RA, Jirtle RL. Transposable elements: Targets for early nutritional effects on epigenetic gene regulation. Molecular and Cellular Biology. 2003;23(15):5293–300. doi: 10.1128/MCB.23.15.5293-5300.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Skinner M, Guerrero-Bosagna C. Environmental signals and transgenerationa; epigenetics. Epigenomics. 2009;1(1):111–17. doi: 10.2217/epi.09.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Douwes J, Cheng S, Travier N, et al. Farm exposure in utero may protect against asthma, hay fever, and eczema. Eur Respir J. 2008;32(3):603–11. doi: 10.1183/09031936.00033707. [DOI] [PubMed] [Google Scholar]
  • 78•.Lin L, Zhu H, Quan C, Grunig G, Ballaney M, Jin X, et al. Prenatal allergen and diesel exhaust exposure and their effects on allergy in adult offspring mice. Allergy, Asthma and Clinical Immunology. 2010;6(1):7. doi: 10.1186/1710-1492-6-7. This is one of the few studies to look at the long-term effects of prenatal allergen and environmental pollution exposure on adult offspring mice. The study found a decrease in IgE production in mice exposed prenatally to A. fumigatus or diesel exhaust particles. Mice that were exposed to both allergen and pollutant were found to have decreased airway eosinophilia. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Schmitz RJ, Schultz MD, Lewsey MG. Transgenerational epigenetic instability is a source of novel methylation variants. ScienceExpress. 2011:1. doi: 10.1126/science.1212959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Barnes P. Targeting the epigenome in the treatment of asthma and chronic obstructive pulmonary disease. Proc Am Thorac Soc. 2009;6:693–96. doi: 10.1513/pats.200907-071DP. [DOI] [PubMed] [Google Scholar]
  • 81•.Williams A, Larner-Svensson H, Perry M, Campbell G, Herrick S, Adcock I, et al. MicroRNA expression profiling in mild asthmatic human airways and effect of corticosteroid. PLoS One. 2009;4(6):e5889. doi: 10.1371/journal.pone.0005889. The authors studied the effects of the PGI2 analogue iloprost on histone modifications in the Th1 and Th2 pathways. They found that iloprost had enhancing effects of histones in the Th2 pathway (histone H3 acetylation) versus suppressive effects of the Th1 pathway. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Kuo C-H, Ko Y-C, Yang S-N, Chu Y-T, Wang W-L, Huang S-K, et al. Effects of PGI2 analogues on Th1- and Th2-related chemokines in monocytes via epigenetic regulation. J Mol Med. 2011;89:29–41. doi: 10.1007/s00109-010-0694-2. [DOI] [PubMed] [Google Scholar]

RESOURCES