Skip to main content
Journal of Cellular and Molecular Medicine logoLink to Journal of Cellular and Molecular Medicine
. 2009 Jun 5;13(9b):2973–2989. doi: 10.1111/j.1582-4934.2009.00800.x

Non-classical transcriptional regulation of HLA-G: an update

Philippe Moreau 1,*, Sébastien Flajollet 1, Edgardo D Carosella 1
PMCID: PMC4516459  PMID: 19508383

Abstract

Human leucocyte antigen-G (HLA-G) plays a key role in maternal–foetal tolerance and allotransplantation acceptance and is also implicated in tumour escape from the immune system. The modulation of HLA-G expression can prove to be very important to therapeutic goals in some pregnancy complications, transplantation, cancer and possibly autoimmune diseases. In spite of substantial similarities with classical HLA-class I genes, HLA-G is characterized by a restricted tissue-specific expression in non-pathological situations. HLA-G expression is mainly controlled at the transcriptional level by a unique gene promoter when compared with classical HLA-class I genes, and at the post-transcriptional level including alternative splicing, mRNA stability, translation and protein transport to the cell surface. We focus on the characteristics of the HLA-G gene promoter and the factors which are involved in HLA-G transcriptional modulation. They take part in epigenetic mechanisms that control key functions of the HLA-G gene in the regulation of immune tolerance.

Keywords: HLA-G, transcription factors, epigenetics, gene regulation


  • Introduction

  • HLA-G gene promoter region: regulatory sites and binding factors

    • The atypical proximal promoter region of the HLA-G gene among classical HLA class I genes

    • Alternative regulatory elements within the HLA-G gene promoter

      • The locus control region

      • cAMP response element/TPA response element

      • Interferon-stimulated response element

      • Heat shock element

      • Progesterone response element

      • Leukaemia inhibitory factor target site

      • Ras response elements

    • Sequence polymorphism within the HLA-G gene promoter and 3′UT region

    • Modulation of HLA-G transcription by micro-environmental       factors with unidentified target sites

      • Cytokines, growth factors, and hormones

      • Hypoxia

  • Chromatin remodelling at the HLA-G gene locus

  • Concluding remarks

Introduction

Human leucocyte antigen-G (HLA-G) is encoded by the major histocompatibility complex with biological and structural properties associated with a specific function in immune tolerance [1]. HLA-G was first characterized as a protein associated with β2-microglobulin expressed in the BeWo choriocarcinoma cell line [2] and later as an array of five 37- to 39-kD isoforms in the cytotrophoblasts of placenta [3, 4]. During its 20-year history, HLA-G has been shown to be of crucial importance in the success of implantation and in foetal–maternal symbiosis during human pregnancy [5, 6]. Beyond this perfect example of successful physiological tolerance to semi-allografts, HLA-G has been demonstrated to contribute greatly to the protection of transplanted organs such as heart [7] and kidney/liver [8, 9] allografts against rejection. Evidence has been accumulated showing that HLA-G expression, which is very restricted in non-pathological conditions, may also be a strategy used by malignant tumours [10] and virus [11, 12] to escape the host’s immune surveillance. HLA-G exerts these major functions by inhibiting NK and T-lymphocyte-mediated cytotoxicity as well as a proliferative allogenic response [13]. This inhibition is mediated through direct binding to the inhibitory receptors ILT-2 (LILRB1/CD85j) [14], ILT-4 (LILRB2/ CD85d) [15] and KIR2DL4 (CD158d) [16]. The HLA-G protective effect can occur in the presence of a few HLA-G-expressing cells, by cell-to-cell contact-dependent uptake of HLA-G (trogocytosis) from APC and tumoral cells to T and NK cells, respectively [17, 18]. This process acts through effector cells made to act as suppressor cells locally and temporarily. Finally, ILT2, ILT3, ILT4 and KIR2DL4 expression is up-regulated by HLA-G in antigen-presenting, NK and T cells, suggesting that up-regulation of inhibitory receptors in immune cells might increase their activation thresholds and participate in immune escape mechanisms [19].

The HLA-G gene was cloned in 1987 [20] and maps on the short arm of chromosome 6 in the p21.31 region. It presents a gene structure consisting of eight exons and seven introns with approximately 86% similarities with the consensus sequence of the HLA-A, HLA-B and HLA-C genes. Exon 1 encodes the signal peptide, exons 2, 3 and 4 encode the α1, α2 and α3 extracellular domains, respectively, and exon 5 encodes the transmembrane domain. Nonetheless, a stop codon in the second codon of exon 6 results in a shorter cytoplasmic tail region in comparison with classical HLA-class I molecules. A low amount of gene polymorphism is found since only 42 alleles are listed by the WHO Nomenclature Committee for factors of the HLA System (http://www.anthonynolan.org.uk), the G*010101 group (five alleles) being the predominant one, with frequency varying from 32% to 83% in Japanese, Caucasian and African populations [21]. HLA-G alleles are essentially characterized by variations in promoter, 3′UT regions, introns and synonymous substitution in exons, low amino acid changes defining 15 protein variants only [2225].

An alternative splicing of the primary transcript generates the membrane-bound isoforms HLA-G1 (complete molecule), HLA-G2 (minus exon 3), HLA-G3 (minus exons 3–4), HLA-G4 (minus exon 4) and the soluble isoforms HLA-G5 (soluble HLA-G1 counterpart), HLA-G6 (soluble HLA-G2 counterpart) and HLA-G7 (α1 domain) (39 to 17 kD) [2628]. Recent crystallography studies have validated the HLA-G1 heterotrimeric structure with its heavy chain non-covalently associated with β2m and a nonamer peptide [29]. The loading of high-affinity peptides (KIPAQFYIL) prevents retrieval of the molecule and results in increased cell surface expression of HLA-G1 [30]. Soluble isoforms are encoded by transcripts in which intron 2 (G7) [28] or intron 4 (G5 and G6) [31, 32] are retained and are translated until a premature stop codon that prevents the translation from exon 3 and 5, respectively. Soluble HLA-G1 also may be generated by proteolytic shedding (sHLA-G1) [33], which is likely to be regulated by NF-κB activation [34]. Besides, it has been demonstrated that HLA-G1 and HLA-G5 may also be produced as β2m-free heavy chains and more importantly as disulphide-bonded homodimers [3537] and that the ILT-2 and ILT-4 binding sites of HLA-G dimers are more accessible than those of HLA-G monomers [38].

Under non-pathological conditions, HLA-G expression occurs during pregnancy primarily in the pre-implanted embryo [6, 39, 40] and at the maternal–foetal interface on extravillous invasive cytotrophoblasts [3, 4]. A number of other extrafoetal cells also express HLA-G, including amnion epithelial cells [41, 42] and endothelial cells of foetal blood vessels in the placenta [43]. Over the past few years, the HLA-G expression pattern was extended to a few healthy adult tissues in immune privileged sites, namely thymus [44, 45], cornea [46], pancreas [47] and the proximal nail matrix [48]. In addition, HLA-G protein may be produced by human decidual stromal cells [49], monocytes [50], keratinocytes [51] and erythroblasts from primitive to definitive haematopoiesis [52]. In contrast, restrictive expression of HLA-G is abrogated under pathological conditions, with up-regulation observed in grafted organs, inflammatory and autoimmune diseases, and viral aggressions [1]. In particular, more than 1000 malignant lesions have been analysed for the HLA-G expression and definitely demonstrate that HLA-G is switched on in numerous tumours [53]. Interestingly, expression frequency varies according to the tumour type, ranging from less than 30% in tumour lesions such as lung carcinoma [54, 55] and breast carcinoma [56, 57] to at least 80% in tumour lesions such as pancreatic ductal carcinoma, biliary cancer [58] and oesophageal squamous cell carcinoma [59]. Nonetheless, HLA-G1 cell surface expression and HLA-G transcripts may be lost along long-term in vitro propagation [60, 61], suggesting that not only HLA-G expression is under the control of genetics but also micro-environmental factors.

On top of the expression regulation of antigen-processing machinery components, post-translational mechanisms such as mRNA stability or protein translation are important for HLA-G expression, in particular during invasiveness of cytotrophoblasts [62]. In this situation, the recent identification of HLA-G-specific miRNAs may be assumed in these mechanisms [24]. On the other hand, a key level of regulation is undoubtedly HLA-G gene transcription, since high amounts of HLA-G transcripts are observed in cells expressing HLA-G protein, while very low amounts or absence of HLA-G transcripts are observed in cells where HLA-G protein is not detected [6366]. HLA-G transcription also exhibits temporal regulation observed during the course of gestation [67], with high levels of HLA-G mRNA in first- and second-trimester trophoblasts, whereas a reduced level of HLA-G mRNA is observed in term cells [3, 4]. This spatiotemporal pattern of HLA-G transcriptional regulation is still partially elucidated and is dependent on transcriptional factors and specific cis-regulatory elements located within a non-classical HLA-G gene promoter [68, 69]. We propose herein to present this in detail.

HLA-G gene promoter region: regulatory sites and binding factors

The HLA-G promoter is unique among the HLA genes [67] with a divergent proximal region when compared to the other HLA, a trophoblast-specific regulatory element located at −1.2 kb from exon 1 [70] and specific regulatory elements. It is worth noting that in the published data, some findings are less valid then others depending on the method used in evaluating regulatory element location and function. Indeed, the analysis could have been performed in vitro (e.g. electrophoretic mobility shift assays [EMSAs] and reporter gene assays), in situ (e.g. chromatin immunoprecipitation [ChIP] and RNA interference) or in vivo (e.g. transgenic mice), a criterion that should be taken into account in evaluating the impact of the results (Table 1). The HLA-G promoter also exhibits a pattern of variations characterized by two divergent lineages, which is consistent with balancing selection. This is probably related to highly regulated expression favouring high- and low-expressing promoters under temporally and/or spatially varying selective pressures [22]. Variations in 3′UTR that could influence RNA stability and/or translation have also been identified [24]. These observations strongly suggest that the HLA-G gene polymorphisms should now be considered a very pertinent parameter in the understanding of HLA-G gene regulation and more particularly in some HLA-G-associated diseases.

Table 1.

Promoter of HLA-G gene and associated transcriptional factors

Response element Location (bp upstream of ATG) Factors Methods used for validation References
CRE/TRE −1380/−1370; −934; −770 CREB1, ATF-1, c-jun EMSA, reporter gene, ChIP [96]
RRE −1356; −142/−133; −53 RREB-1 EMSA, reporter gene, ChIP (Flajollet et al. unpublished data)
ISRE −744 IRF-1 EMSA, reporter gene [100]
HSE −459/−454 HSF-1 EMSA [103]
P50, SP1 −187/−171; −166 P50, SP1 EMSA [72]
X1 box −124 RFX5 EMSA (negative using ChIP) [87, 88]
CAAT −71 CTF
TCTAAA −44 TFII
PRE −37 PR EMSA, reporter gene, DNA fragment-binding ELISA [105]

The atypical proximal promoter region of the HLA-G gene among classical HLA class I genes

Classical HLA-class I gene promoters contain two main regulatory modules, namely enhancer A/interferon (IFN)-stimulated response element (ISRE) and SXY boxes, located within the 220 bp 5′ of the gene initiation codon (ATG). These two distinct cis-acting elements contribute to the constitutive and inducible level of MHC class I genes, the SXY module being shared by MHC class II genes [71]. The most upstream module contains the enhancer A with κB2 and κB1, two palindromic binding sites for the NF-κB/rel family members and a Sp1 transcription factor site [72, 73]. This module is also composed of a response element localized −180 bp from the ATG (consensus sequence AGTTTCNNTTCT) that may bind factors of the interferon regulatory factor (IRF) family such as IRF-1, IFN consensus sequence binding protein (gene activation), IFN-stimulated gene factor (ISGF)-3 and IRF-2 (repressors) [73]. Additionally, E-box elements can be found in the upstream HLA-class I module and are binding sites for upstream stimulatory factor (USF)-i and USF-2 [73].

The downstream SXY module was first demonstrated to be crucial in the regulation of HLA-class II expression [74]. It comprises the X1, X2 (site α) boxes and Y box (an inverted CCAAT-binding site, also named enhancer B), bound by the multiprotein complex RFX (RFX5, -AP, -ANK/B) [7578], X2-BP/ATF/cAMP response element-binding (CREB) [79] and NF-Y [80] factors, respectively. All these factors cooperate to allow the formation of a stable multiprotein complex and the binding of the class II transactivator (CIITA), which mediates constitutive and IFN-γ-induced expression of HLA-class I molecules [8185]. S box function is not fully understood and could possibly play a role in promoter architecture [86].

Like classical HLA-class I promoters, HLA-G promoter exhibits a CCAAT box and an unusual TATA element, TCTTAA, controlling basal regulation. A transcriptional initiation site located 25 bp downstream of TATA is conserved in HLA-G. Nonetheless, a second putative initiation site of HLA-G transcription has been reported 51 bp upstream of TATA (CTCACTCCC) (http://www.ncbi.nlm.nih.gov/), but the functionality of both sites has not yet been demonstrated. On the other hand, a modified enhancer A and a deleted ISRE render the HLA-G gene promoter unresponsive to NF-κB [72] and IFN-γ[73]. The p50 homeodomain subunit of NF-κB displays a strong binding affinity to the two κB sites in vitro, but this subunit without p65/relA does not possess a transactivation function [72]. In addition, the upstream region encompassing the SXY module only contains conserved S and X1 sequences and despite its binding capacity in vitro, the intact X1 box is unresponsive to the RFX5 factor in situ[87, 88]. The absence of RFX5 binding and the presence of nucleotide variations in X2-Y boxes explain the absence of a higher-order complex and the lack of CIITA recruitment at the HLA-G promoter in situ[88] and in transfection experiments with CIITA expression plasmid alone or in co-transfection experiments with luciferase reporter constructs containing the HLA-G promoter fragment [69, 86]. In comparison with non-classical HLA-class I genes, HLA-E and HLA-F, HLA-G promoter, with its unique characteristics, is the odd one out [68] since the HLA-E promoter is induced through the SXY module even if it is not regulated by enhancer A and ISRE, and HLA-F promoter exhibits extensive similarities with those of classical HLA-class I genes [69]. Furthermore, a putative negative regulatory element located in exon 1 of HLA class I genes, whose activity can be modulated by hormones [89], is also deleted in the HLA-G gene and might, in part, explain why HLA-G is expressed in human trophoblasts, whereas HLA class I genes are not.

Finally, Monarch 1, a factor belonging to the CATERPILLER gene family, which is expressed primarily by myeloid-monocytic cells, is the only factor found to enhance both classical HLA-class I and HLA-G genes. More particularly, Monarch-1 was demonstrated to enhance the HLA-B promoter but the regulatory target elements have not been identified to date [90].

Alternative regulatory elements within the HLA-G gene promoter

To investigate alternative transactivation pathways to the conserved MHC class I regulatory routes, one of the strategies commonly used was to perform transient transfections with luciferase reporter constructs containing HLA-G promoter fragments of different lengths. This identified a negative regulatory sequence within the HLA-G promoter fragment extending 450 bp from the ATG and a slightly induced activity of intron 2-containing construct in JEG-3 choriocarcinoma cells [91]. Other major strategies were the use of HLA-G transgenic mice into which HLA-G transgenes of different lengths were introduced, in silico analysis of the HLA-G gene promoter sequence, and the identification of the target sites of specific modulatory agents present in the placental or tumoral micro-environment. From these analyses emerged regions identified as putative key components of the HLA-G gene regulation. However, precise target sites and/or transcription factors have not always been clarified, as is the case for IFN-γ, leukaemia inhibitory factor (LIF) and hypoxia.

The locus control region

A positive regulatory region of the HLA-G gene was found that contained in a 244-bp HindIII/EcoR1 fragment located −1.2 kb from the ATG, using HLA-G transgenic mice [70, 92]. This fragment was demonstrated to be critical for spatio-temporal expression of HLA-G mRNAs by using a HLA-G transgene including the entire coding region, 1 kb of the 3′ flanking region and 1.2 or 1.4 kb of the 5′ flanking region (5.7- and 6.0-kb fragment, respectively). Indeed, the presence of the 244-bp HindIII/EcoR1 region was associated with a tissue-specific pattern of HLA-G expression in spongiotrophoblasts and mesenchymal cells in placenta and in thymus. HLA-G transcription obtained with the 6.0-kb fragment parallels that seen in human extraembryonic tissues during the course of gestation, reaching the highest levels of expression in trophoblast at day 12.5 and then decreasing progressively before parturition. The distal 244-bp fragment that is required for tissue-specific expression of HLA-G presents a similarity in function with a locus control region (LCR). In agreement with this, sequence analysis of the regulatory fragment has revealed similarities to important elements for the activity of the H3S region of the β-globin LCR [93] such as the TATA symmetrically surrounded by GGGTGG and the putative AP1-binding site [94]. Moreover, DNA binding assays with nuclear extracts from HLA-G+ and HLA-G cells revealed the formation of several complexes in this region. Some of them are specific of HLA-G expression status, whereas others are shared complexes [65, 95]. One of these shared complexes has been identified in an independent study by Van den Elsen’s group and corresponds to the DNA binding of ATF1/CREB1/c-jun on the CRE/TRE (cyclic AMP-response element/TPA-response element) located at position −1380/−1370 in the putative HLA-G LCR [96]. ChIP assay demonstrated the in situ binding of CREB-1 and c-jun to this region [96].

cAMP response element/TPA response element

The computer-assisted search for alternative putative regulatory elements in the promoter had led Van den Elsen’s group to localize two additional functional CRE/TRE elements dispersed through the promoter region at positions −934 and −770 from the ATG [96]. EMSAs demonstrated that CRE/TRE−934 and CRE/TRE−770 bind CREB1 and ATF1/CREB1 factors, respectively. Promoter activity assays and mutagenesis studies revealed the crucial role of the three CREs for the basal level of HLA-G promoter activity and its transactivation, with the most important contribution of CRE−1380/−1370 within putative LCR. In accordance with this, transient transfection of the CREB repressor ICER (inducing cAMP early repressor) inhibits the CREB-induced transactivation of the HLA-G gene’s 1438-bp promoter, while enhanced transactivation occurs with the co-activators CBP/P300 (CREB binding Protein). Notably, CREB, CBP/P300 and HLA-G are co-expressed in extravillous cytotrophoblasts. Nonetheless, CREB association to the HLA-G gene promoter in situ was also observed in HLA-G cell lines, which strongly suggests that tissue-specific expression of HLA-G involves additional regulation mechanisms, including epigenetics.

Interferon-stimulated response element

The cascade of events initiated by IFNs involves the activation of JAK/STAT transduction pathways and the transactivation of the gene promoter with the ISRE and IFN-γ activation site (GAS). Despite a non-conserved HLA class I ISRE in the proximal promoter of HLA-G, several investigations have revealed that the HLA-G gene is responsive to up-regulation following treatment with IFN-α, IFN-β and IFN-γ[97, 98]. Enhancement of steady-state levels of HLA-G mRNA upon IFN treatments was observed in several cell types such as trophoblast cell lines [97], blood cells (monocytes and macrophage cell lines) [50] and glioblastoma cell lines [99]. In particular, IFN-β enhances the levels of HLA-G transcripts in trophoblast explants, amnion and thymus-derived epithelial cells [100]. However, it is a general rule that a basal HLA-G transcriptional level is required for IFN-induced up-regulation of HLA-G mRNAs [97] and then up-regulation of HLA-G cell surface expression.

Computer-assisted searches within the HLA-G promoter sequence led our group to identify an ISRE motif, which is highly homologous to the consensus ISRE. It is located at position −744 bp upstream of ATG, beside a GAS-like element (−734) previously shown to be unable to interact with a GAS-binding complex [98]. We demonstrated that the HLA-G ISRE is a binding site for IFN-response factor-1 (IRF-1) and transactivates HLA-G expression following IFN-β treatment [100]. Despite a weak induction in JEG-3 cells [96, 100], the activity of HLA-G promoter was clearly significant in the thymic epithelial cell LT-TEC2 [100].

Although IRF-1 binds to HLA-G ISRE, no transactivation effect in response to IFN-γ was observed in luciferase assays using the 1.4-kb HLA-G promoter [91, 96]. Besides, the use of a model system consisting of mouse fibroblasts transfected with a 6.0-kb fragment containing the whole HLA-G gene demonstrated the presence of elements that respond to IFN-γ[101]. Consequently, other regulatory pathways or IFN-γ responsive elements should be located outside the 1.4-kb promoter region with the HLA-G gene and/or the 3′UT region.

Heat shock element

Stress-induced proteins have been implicated in balancing immune responses during various diseases [102]. This prompted us to evaluate the effect of stress on HLA-G gene expression in the M8 (melanoma) and T98G (glioblastoma) HLA-G cell lines using heat shock at 42°C or arsenite treatment for 2 hrs. Stress induced an increase in the level of HLA-G mRNA with a specificity compared to other HLA class I transcripts. Interestingly, HLA-G6 transcript was induced prior to the other HLA-G transcripts, suggesting tight control of HLA-G alternative splicing. The study also identified a heat shock element (HSE) within the HLA-G promoter at position −459/−454 that is defined as a repetition of the pentanucleotide NGAAN arranged in alternating orientation. The HSE seems to be functional since it binds heat shock factor-1 (HSF-1) in vitro by EMSAs [103]. Nonetheless, additional functional analysis using a reporter gene under the control of HLA-G promoter with wild-type or mutated HSE would be necessary to improve HLA-G HSE functionality.

Progesterone response element

Progesterone is an essential steroid to maintain pregnancy and has been suggested to be an important immune modulator during this time. That is why Librach’s group investigated the potential effects of progesterone on HLA-G gene expression, revealing that at 10, 100 and 1000 ng/ml progesterone enhanced HLA-G mRNA expression in JEG-3 cells by 2.36-, 10.53-, and 17.58-fold, respectively, as compared to controls [104]. More recently, this group demonstrated that the HLA-G gene promoter is up-regulated by progesterone through a specific binding site for the progesterone receptor (PR) complex [105]. The identified progesterone response element (PRE)-like sequence is a 15-bp non-classical consensus core sequence that has 60% homology to the wild-type mouse mammary virus (MMTV) PRE and a weaker affinity for PR complexes than MMTV-PRE, probably because of variations in the fixed half site of steroid hormone response elements. This PRE is located −37 bp from the ATG and overlaps the HLA-G TATA box, but the authors do not exclude the presence of other PREs in the HLA-G promoter region. Indeed the chloramphenicol acetyltransferase reporter gene assay was not performed with the scrambled HLA-G PRE site and although the authors stipulated that PRE is specific of the HLA-G promoter, very similar sequences can be found at the same location in the HLA-class I promoter.

Leukaemia inhibitory factor target site

LIF is a pleiotropic cytokine that is expressed at the maternal–foetal interface and plays an essential part in embryo implantation and in mediating interactions between maternal decidual leucocytes and trophoblasts [106, 107]. Upon 72 hrs stimulation with LIF, an up to 3.6-fold elevation of HLA-G mRNA has been demonstrated with JEG-3 choriocarcinoma cells. Luciferase reporter gene assays demonstrated that the stimulation of transcription was driven by a 890-bp promoter fragment of the 5′HLA-G gene flanking region [108]. Nonetheless, precise target site(s) have not yet been identified. Moreover, endoplasmic reticulum aminopeptidase-1 (ERAP1) is also induced by LIF and plays a role in presenting antigenic peptides to HLA-G and then mediates HLA-G cell surface expression [109].

Ras response elements

To further identify factors involved in the regulation of HLA-G gene expression, our group recently developed a specific proteomic approach to characterize proteins differentially bound to the proximal and distal HLA-G gene promoter. Biotinylated, double-stranded HLA-G promoter fragments were incubated with nuclear protein extracts of HLA-G+ and HLA-G cells for transcriptional activity and isolated with streptavidin-coated magnetic beads. This DNA-affinity strategy was followed by a 2D separation and the proteins of interest were analysed using mass spectrometry. This allowed us to identify the zinc finger protein Ras responsive element binding 1 (RREB-1) [110] that is capable of binding three Ras response elements (RREs) along the HLA-G gene promoter.

We demonstrated that RREB-1 is involved in the repression of HLA-G transcriptional activity, acting through the recruitment of factors such as histone deacetylase 1 (HDAC1) and C-terminal binding protein (CtBP) [111] implicated in chromatin remodelling (Flajollet et al., unpublished data).

Sequence polymorphism within the HLA-G gene promoter and the 3′UT region

Evidence has been accumulated showing that the HLA-G gene polymorphism is involved in the regulation of the HLA-G gene transcriptional activity. In particular, some HLA-G allelic variants are associated with differences in the pattern of HLA-G mRNA isoforms and HLA-G mRNA levels [112, 113]. On the one hand, polymorphism in the 3′UT region such as the absence or presence of 14 bp of ‘exon 8’ has been studied more extensively. The presence of the 14 bp is associated with low levels of mRNA expression [113, 114] and mediates or is involved in the out-splicing of the first 92 b of exon 8 [115]. These transcripts were shown to be more stable than the complete RNA [116]. Moreover, a C/G single nucleotide polymorphism (SNP) at +3142 bp in the HLA-G mRNA has recently been demonstrated to influence the targeting of three microRNAs [24]. On the other hand, an effort has been made to extend the analysis of HLA-G gene variations to the promoter region. To date, 29 SNPs have been identified [22, 23, 117] within this region (Fig. 2) and we cannot exclude that in some cases polymorphism on the promoter may be in linkage disequilibrium with 3′UT variants and that some of them could influence alternative splicing [118]. Interestingly, many of the polymorphisms either coincide with or are closed to the known regulatory elements and thus may affect the binding of the corresponding regulatory factors (Fig. 2).

Figure 2.

Figure 2

SNPs along the 1.4 kb of the HLA-G gene promoter sequence. ATG: initiation codon; LCR: locus control region; Δ: deletion.

Figure 1.

Figure 1

HLA-class I and HLA-G gene promoter: cis-regulatory sequences and their interacting factors. Boxes with X within HLA-G promoter indicate that mutations prevent binding of classical HLA-class I transacting factors

Modulation of HLA-G transcription by micro-environmental factors with unidentified target sites

The restricted expression of HLA-G in physiological conditions and its up-regulation in pathological situations reveals a significant correlation of HLA-G transcription with biochemical environment [49]. This is strongly supported by the fact that HLA-G expression and transcription may be specifically down-regulated or even lost during long-term culture of biopsy-derived cancer cells [60, 61].

Variations of environmental factors such as growth factor, cytokines, hormones, physical conditions and stress dramatically occur during pregnancy as well as during inflammation, viral infection and cancer. Interestingly, numerous micro-environmental factors and molecular circuits are shared by placental and pathological situations such as cancer [119]; some of them have been demonstrated to be involved in HLA-G gene transcription. However, with the exception of IFNs and progesterone, the mechanisms by which most of these key regulatory molecules exert control on the HLA-G gene transcription, either directly on the HLA-G promoter or not, need further exploration.

Cytokines, growth factors and hormones

Placenta is the main HLA-G-producing site that releases and/or is in contact with a variety of cytokines, both anti-inflammatory (IL-10, IL-4, IL-5, IL-6) and pro-inflammatory (tumor necrosis factor [TNF]-α, IL-1β, IL-2), and transforming growth factors (TGF-β), granulocyte macrophage colony stimulating factor [GM-CSF], granulocyte colony stimulating factor [G-CSF], colony-stimulating factor [CSF-1], LIF and epidermal growth factor [EGF]). Autocrine and paracrine mechanisms occur through specific cytokine receptors. Cytokines are crucial for successful embryo implantation and contributes to the maternal metabolic changes necessary to accommodate the increased energy needs of the foetus. Pregnancy is associated with dynamic changes in cytokine levels and ratios and is mainly characterized by an increase in the concentration of cytokines in the second half of pregnancy, with many changes orchestrated around IL-12 [120, 121]. Pregnancy also affects hormones in the body, mostly because of the effects of hormones produced by the placenta. In particular, human chorionic gonadotrophin, produced by the developing placenta, stimulates the ovaries to produce the oestrogen and progesterone needed to sustain pregnancy. By the fourth month of pregnancy, the placenta takes over from the ovaries as the main producer of oestrogen and progesterone [122]. These hormones are involved in womb changes to make room for the growing baby. Other hormones come into play that help the womb to contract during and after labour (oxytocin) as well as stimulate the production and release of breast milk (prolactin).

Our group suggested that these molecules might influence HLA-G gene expression, so we investigated the effect of a part of the network of cytokines and hormones involved in the placental or shared tumoral micro-environment on the HLA-G transcription of cultured trophoblast explants (Fig. 3). As a result, we observed that the amounts of HLA-G transcripts increased approximately 7.2-fold following treatment with IL-10 [123] and IL-1β, and from approximately 1.6- to fivefold with glucocorticoid hormones [124], TGF-β, EGF, β-oestradiol, prolactin and IFNs in comparison with HLA-G transcript levels in untreated trophoblasts. Besides, IL-10, a major suppressor of the immune response and inflammation [125], was demonstrated to increase HLA-G mRNA and cell surface protein expression by monocytes [123] and renal cell carcinoma (RCC) cell lines [126], HLA-G mRNA and soluble HLA-G protein by mononuclear cells from patients suffering from non-Hodgkin lymphomas (T-NHL) [127], acute myeloblastic leukaemia (AML) and acute lymphoblastic leukaemia (B-ALL), and HLA-G protein cell surface expression by decidual stromal cells [49] and the FON melanoma cell line [60]. Notably, reduced placental IL-10 production occurs in human pathological pregnancies such as in pre-eclampsia [128] in which a defect in HLA-G transcription is observed [114, 129]. Likewise, IL-10 production in cancer is associated with HLA-G expression [130] and the IL-10 homolog produced by cytomegalovirus (CMV) can up-regulate HLA-G protein expression at the monocyte cell surface [131]. On the other hand, GM-CSF treatment combined with IFN-γ and/or IL-2 has been shown to enhance both HLA-G mRNA and soluble HLA-G protein in mononuclear cells from patients suffering from T-NHL [127], AML and T-ALL [132]. Moreover, cell surface expression on the U937 monohistiocyte cell line was demonstrated, but the effect on the HLA-G transcription was not investigated [133, 134] (Table 2).

Figure 3.

Figure 3

RNAse protection assay analysis of HLA-G mRNA levels in four trophoblast organ explants (namely 151, 159, 218, 172) of first-trimester gestation incubated 72 hrs without (0) or with micro-environmental factors (IL-10: 2 ng/ml; IL-1β: 17 pg/ml; HC: 10−8 mol/l hydrocortisone; DEX: 10−7 mol/l dexamethasone; TGF-β: 10 ng/ml; EGF: 20 ng/ml; β-OEST: 10−7 mol/l β-oestradiol; IFN-γ: 100 U/ml; PROL: 10−10 mol/L prolactine; IFN-β: 1000 U/ml). The HLA-G template used for the riboprobe was obtained by PCR amplification of HLA-G genomic fragment located in the 3′UT region as previously described [124]. Ratio indicates the values obtained for HLA-G signals normalized to the constitutively expressed cyclophilin (Cycloph.) signal compared with untreated trophoblasts (assigned as value of 1). M: molecular weight size marker.

Table 2.

HLA-G transcriptional effectors and effect on HLA-G expression

Effector molecules Modulation of HLA-G gene expression References
mRNA Protein
Flow cytometry Western blot ELISA/ELIspot
Growth factors/cytokines
EGF ↑. Trophoblasts nd nd nd Present paper
GM-CSF + IFN-γ ↑/=. Mononuclear cells from T-NHL, AML,T-ALL ↑. U937 (histiocytic lymphoma) ↑. Mononuclear cells from T-NHL ↑/=. Mononuclear cellsm from AML, B-ALL, T-ALL, T-NHL =. PBMC [127][132134]
IL-1 β ↑. Trophoblasts nd nd nd Present paper
IL-2 + IFN-γ ↑. Mononuclear cells from T-NHL ↑. U937 nd ↑. Mononuclear cells from T-NHL [127][133]
IL-2 + IFN-γ+ GM-CSF ↑. Mononuclear cells from T-NHL nd ↑. Mononuclear cells from T-NHL ↑ Mononuclear cellsm from B-ALL, T-NHL =. PBMC [127]
IL-10 ↑. Trophoblasts. Monocytes ↑. Monocytes. Decidual stromal cells ↑. Mononuclear cells from T-NHL. Decidual stromal cells ↑. JEG-3 ↑/=. Mononuclear cells from AML, T-NHL [49][60][123]
↑/=. Mononuclear cells from AML, B-ALL, T-NHL. RCC cell lines ↑/=. RCC cell lines=. FON (melanoma) =. PBMC [126, 127] [132] [162]
IFN-α ↑. JEG-3 ↑. U937 nd ↑. JEG-3 [50]
. U937 .THP-1 (acute monocytic leukaemia) . Serum of treated patient with melanoma [162, 163]
. Blood monocytes
=. JAR =. JEG-3
IFN-γ ↑. JEG-3 ↑. U937 ↑. Decidual stromal cells ↑. JEG-3 [49, 50]
. U937 . Blood monocytes. THP-1 =. PBMCm [60]
. Blood monocytes. THP-1 . Mononuclear [123]
cellsm from [126]
. HL-60 (acute promyelocytic leukaemia) AML, B-ALL, T-ALL [162]
[164]
. RCC cell lines . Decidual stromal cells
. Glioblastoma cell lines
. FON (melanoma)
=. JAR . RCC cell lines
=. JEG-3
LIF ↑. JEG-3 ↑. JEG-3 nd ↑. JEG-3 [60]
=. FON [108]
[109]
TGF-β ↑. Trophoblasts ↓. FON nd nd Present paper
[60]
TNF-α/PMA =. JEG-3 ↓. JEG-3 ↑. JEG-3 ↑. M8-HLA-G1 [34]
. FON . FON
. M8-HLA-G1 . M8-HLA-G1
(Melanoma transfectant)
Hormones
β-oestradiol ↑. Trophoblasts nd nd nd
Progesterone ↑. JEG-3 nd ↑. JEG-3 ↑. JEG-3 [104, 105] Present paper
.Cytotrophoblasts . Trophoblasts
Progesterone +cAMP nd ↑. Decidual stromal cells ↑. Decidual stromal cells nd [49]
Glucocorticoids ↑. Trophoblasts nd nd nd Present paper
Prolactin ↑. Trophoblasts nd nd nd Present paper
Stress
Arsenite ↑. M8a (melanoma). T98Ga (glioblastoma) =. M8 nd nd [103]
. T98G
Heat shock ↑. M8a =. M8 nd nd [103]
. T98Ga . T98G [141]
. JAR [144]
. Raji
Hypoxia/DFX/CoCl2 ↑. Extravillous cytotrophoblasts ↓. HTR-8/SVneo on matrigel (first-trimester cytotrophoblast) ↓. HTR-8/SVneo on matrigel nd [142]
[144, 145]
↑. Melanoma cells (M8a, 1074mela) . FON [165]
. Rajia (Burkitt’s B lymphoma) =. JAR . JEG-3
Epigenetic treatments
5-azacytidine/5-aza-2′deoxycytidine ↑. JARa ↑. JAR ↑. JAR [60]
. FON . FON . Raji [126]
. Melanoma cell lines (OCM-1aa, M8a, FON). Glioblastoma (U87MG, LN-229, LN-428). B,T and myelomonocytic Leukaemia cell lines.. BG-1a (ovarian cancer cells) ↑/=. RCCa cell lines ↑. Tera-2 (lung embryonic carcinoma) . Raji . LCL721.221 [144]
. LCL721.221 (lymphoblastoid B cells)=. NKL (NK cell leukaemia). KG1a (acute myelogenous leukaemia). M8 =. NKL. KG1a ↑/=. RCC cell lines [150]
[152154]
. M8 [156]
[158]
NaBu/TSA/VA ↑. M8 nd nd nd [144]
. JAR [152]
. Raji

ALL: acute lymphoblastic leukaemia; AML: acute myeloblastic leukaemia; DFX: desferrioxamine; NHL: Non-Hodgkin lymphoma; PBMC: peripheral blood mononuclear cells and VA: valproic acid.

a: Raise HLA-G gene repression.

m: Mean of several experiments.

↑: up-regulation (↑: low).

↓: down-regulation (↓: low).

=: no effect.

↑/=: up-regulation or no effect.

↓/=: down-regulation or no effect.

nd: not determined.

Yet, although preliminary data show a clear potential effect of cytokines, growth factors, and hormone on the modulation of HLA-G transcription, some of them require additional investigation and it is likely that a combination of agents might be an interesting and pertinent avenue of research. Moreover, the molecular mechanisms responsible for enhancing HLA-G transcription by almost all the modulators described, either alone or in combination, remain to be elucidated. In particular, they could act through promoters of transacting factors instead of direct interaction with the HLA-G gene promoter and they also might be involved in RNA stability. Whatever the mechanisms at work, the efficiency of IL-10 or IFN treatment requires basal transcriptional activity to enhance HLA-G gene activity. A hypoxic environment is one possible candidate factor to reverse HLA-G gene repression.

Hypoxia

Maintenance of oxygen (O2) homeostasis is critical for the maintenance of life. Hypoxia induces a series of adaptive physiological responses observed in biological processes such as maternal-foetal interactions and cancer, since it is associated with cell proliferation. Hypoxia followed by normoxia regulates the depth of the trophoblastic invasion and the vascular remodelling of the uterine tissues [135]. Tumoral hypoxia (up to 50–60% of solid tumours) arises as a result of an imbalance between the supply and consumption of O2[136]. In response to hypoxic conditions and to restore pO2 homeostasis, cells instantaneously express a key factor, the hypoxia-inducible factor (HIF), which activates transcription of over 70 genes controlling glycolysis, glucose transport, cell survival and death, cell adhesion, angiogenesis and erythropoiesis [137, 138]. HIF is a heterodimer composed of constitutively expressed HIF-1β and inducibly expressed HIF-1α subunits. Under normoxic conditions (21% O2), HIF-1α is hydroxylated and targeted by the van Hippel–Lindau tumour-suppressor protein, which upon synthesis causes its rapid degradation by the ubiquitin-26S proteasome pathways. With a low O2 level (<6%) hydroxylation in HIF-1α is inhibited, resulting in a translocation of the subunit into the nucleus, a dimerization with HIF-1β, and binding to the hypoxia responsive element (HRE) (RCGTG consensus sequence) on the promoter of target genes [139]. Besides, HIF is able to interact with the transcriptional co-activators CBP/P300, SRC1, TIF2, which potentiate transactivation [137, 140].

Stress conditions such as heat shock or arsenite treatment were demonstrated to induce HLA-G gene transcription in HLA-G cells. The effect of hypoxic stress on the HLA-G transcription was first assessed by Ferrone’s group using the iron chelator desferrioxamine (DFX) to stabilize HIF-1α in HLA-G melanoma cells. They found that DFX treatment induces HLA-G gene transcription in seven of 13 melanoma cell lines. The effect is dose- and time-dependent and approximatively 16-fold lower than the level of constitutive mRNA in the JEG-3 choriocarcinoma cell line [141]. Our group confirmed these results with the M8 melanoma cell line, thus supporting hypoxia as a candidate micro-environmental factor to reverse HLA-G gene repression [142]. The effect on the HLA-G mRNA up-regulation was further observed with undifferentiated cytotrophoblasts isolated from first-trimester placenta cultured with different concentrations (20%, 8%, 2%) of O2[143] and with Raji (Burkitt’s B lymphoma) cells [144]. The computer search analysis of the HLA-G gene promoter sequence identified a putative consensus HRE located −243 bp upstream of the ATG. Whether this HRE is functional is still unknown [141].

As for the effect of hypoxia on HLA-G cell surface expression, results differ depending on the cell type or culture conditions. It is likely that post-transcriptional mechanisms may be involved since the lack of HLA-G translation can be observed in cells with hypoxia-induced mRNA [141, 142]. On top of that, the down-regulation of HLA-G expression is reported in cells expressing HLA-G at the cell surface following exposure to low O2 concentration [145] or DFX [142].

Chromatin remodelling at the HLA-G gene locus

The epigenetic control of gene promoters is a critical mechanism in transcriptional regulation since it determines the accessibility and recruitment of regulatory factors to the DNA. Epigenetic modifications involve DNA methylation and histone tail modifications such as acetylation, phosphorylation, methylation, ubiquitylation and sumoylation [146]. Epigenetic processes control implantation, placentation, organ formation and foetal growth. In particular, there is a stepwise decline in DNA methylation from fertilization until the morula stage [147] and this correlates with the activation of HLA-G gene transcription [39, 40]. Alteration in epigenetics may contribute to pathological situations such as pre-eclampsia [148] and are widely recognized as contributing to tumorigenesis [149].

Evidence for HLA-G silencing by a DNA methylation process was reported first by Le Bouteiller’s group in the HLA-G choriocarcinoma JAR cell line [150, 151]. The use of demethylating agents such as 5-azacytidine and 5-aza-2′-deoxycitidine (5-Aza-dC) further demonstrated that the repression of HLA-G gene activity in cultured cell lines of various origins is reversed by demethylating treatment [144, 150, 152155] and is maintained at least 5 days [153]. This treatment may also enhance steady state levels of HLA-G mRNA [60, 153] and it directly induces HLA-G protein expression in JAR, Raji (Burkitt’s B lymphoma, LCL721.221 (lymphoblastoid B cell) [152], OCM-1A (melanoma) [154] and RCC cells [155] as well as human leukaemia cell lines [156]. The HLA-G inhibitory process seems to be independent of the expression of other HLA-class I and HLA-class II, which are detected differentially according to the cell line. By direct sequencing of bisulphite-treated DNA, one study found no correlation between HLA-G gene transcriptional activity and methylation of 63 CpG islands at the HLA-G locus in blood cells expressing (CD2+ lymphocytes) or not expressing (CD34+ haematopoietic cells) HLA-G mRNA [157]. Nonetheless, recent data focussing on the promoter region covering 450 bp 5′ from the ATG (19 CpG sites) strongly argue for a cis-acting CpG methylation associated with HLA-G gene silencing [153, 154, 158]. Moreover, the analysis of chromatin remodelling at the HLA-G locus performed with histone deacetylase inhibitors (HDAC), trichostatin A (TSA) and sodium butyrate (NaBu) treatments, demonstrated the HLA-G gene activation in M8 (melanoma), JAR and Raji cells, despite a lower mRNA level in comparison with 5-Aza-dC treatment [144, 152]. To back this up, the analysis of H3 and H4 histone acetylation by ChIP of the proximal and distal HLA-G gene promoter showed the presence of hyperacetylated histones in HLA-G+ cells (FON and JEG-3 cells), while hypoacetylated histones were predominant in HLA-G cells (M8 and JAR cells) [153]. Thus, the acetylation of lysine residues of H3 and H4 renders the chromatin in a permissive state, arguing in favour of HLA-G gene expression.

Concluding remarks

The present data highlight the complexity of the regulation of non-classical HLA-G transcriptional activity, which is likely to be associated with the tight control of HLA-G function participating in immune tolerance. Several regulatory elements have now been identified, but some of them need further functional validation using reporter genes and interference RNA tools. The demonstration of their presence or absence in situ using ChIP assays is also required since the chromatin environment is crucial for binding and function. Cytokines and micro-environmental agents act on HLA-G gene expression in a cell-specific manner and may have pleiotropic activities. From this point of view, the medical relevance of the current findings on the HLA-G regulation should be carefully evaluated. In this regard, the development of animal models would make it possible to validate in vivo both the concept that HLA-G is a key component in immunoregulation and the molecular mechanisms modulating its expression. Moreover, an emerging point of view is that the impact of HLA-G gene polymorphism and temporal data on regulatory processes are needed to understand modifications at the HLA-G locus, particularly epigenetic changes following micro-environmental signals during placentation and various pathological situations.

Considering all the regulatory mechanisms known to date, they support a coherent regulatory model of HLA-G gene expression based on those previously proposed by Ferrone’s group [159]: HLA-G is generally not expressed under non-pathological conditions in vivo, probably because the promoter is inactivated by DNA methylation and at least histone hypoacetylation. During in vivo proliferative processes, cells will be exposed to stress (e.g. hypoxia) and undergo epigenetic changes such as DNA demethylation and histone acetylation, leading to opened chromatin and accessibility to transcription factors. It is likely that sequence variations at specific transcription factor target sites influence the level of response. Upon gene activation, cytokines and hormones of the micro-environment will enhance the amounts of HLA-G transcripts and then protein expression. The micro-environment could also induce antigen-processing machinery components, thus contributing to the transport and stabilization of HLA-G molecules at the cell surface. TNF-α may enhance intracytoplasmic HLA-G cell content and may enhance HLA-G1 proteolytic shedding following NF-κB activation. Upon adaptation to tissue culture in vitro, cells may not be exposed to the same stressful conditions and HLA-G transcription may be changed. Thus long-term growth in vitro may subsequently lead to methylation and hypoacetylation of the HLA-G promoter and silence the gene.

Finally, the fact that demethylation treatment may activate HLA-G receptor genes [160, 161] should be considered in cancer therapy using treatments blocking HDAC and/or reversing DNA methylation to enhance tumour suppressor genes. These treatments might favour the enhancement of both HLA-G at the cell surface of tumoral cells, and KIR protein expression at the cell surface of tumour infiltrating lymphocytes might thus favour tumour escape. Therefore, in addition to the crucial need to better understand HLA-G function, extensive studies on the control of HLA-G gene expression are fundamental to developing non-deleterious therapeutic strategies.

References

  • 1.Carosella ED, Moreau P, Lemaoult J, et al. HLA-G: from biology to clinical benefits. Trends Immunol. 2008;29:125–32. doi: 10.1016/j.it.2007.11.005. [DOI] [PubMed] [Google Scholar]
  • 2.Ellis SA, Sargent IL, Redman CWG, et al. Evidence for a novel HLA antigen found on human extravillous trophoblast and a choriocarcinoma cell line. Immunology. 1986;59:595–601. [PMC free article] [PubMed] [Google Scholar]
  • 3.Kovats S, Main EK, Librach C, et al. A class I antigen, HLA-G, expressed in human trophoblasts. Science. 1990;248:220–3. doi: 10.1126/science.2326636. [DOI] [PubMed] [Google Scholar]
  • 4.McMaster MT, Librach CL, Zhou Y, et al. Human placental HLA-G expression is restricted to differentiated cytotrophoblasts. J Immunol. 1995;154:3771–8. [PubMed] [Google Scholar]
  • 5.Rouas-Freiss N, Goncalves RM, Menier C, et al. Direct evidence to support the role of HLA-G in protecting the fetus from maternal uterine natural killer cytolysis. Proc Natl Acad Sci USA. 1997;94:11520–5. doi: 10.1073/pnas.94.21.11520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Fuzzi B, Rizzo R, Criscuoli L, et al. HLA-G expression in early embryos is a fundamental prerequisite for the obtainment of pregnancy. Eur J Immunol. 2002;32:311–5. doi: 10.1002/1521-4141(200202)32:2<311::AID-IMMU311>3.0.CO;2-8. [DOI] [PubMed] [Google Scholar]
  • 7.Lila N, Amrein C, Guillemain R, et al. Human leukocyte antigen-G expression after heart transplantation is associated with a reduced incidence of rejection. Circulation. 2002;105:1949–54. doi: 10.1161/01.cir.0000015075.89984.46. [DOI] [PubMed] [Google Scholar]
  • 8.Creput C, Durrbach A, Menier C, et al. Human leukocyte antigen-G (HLA-G) expression in biliary epithelial cells is associated with allograft acceptance in liver-kidney transplantation. J Hepatol. 2003;39:587–94. doi: 10.1016/s0168-8278(03)00354-4. [DOI] [PubMed] [Google Scholar]
  • 9.Qiu J, Terasaki PI, Miller J, et al. Soluble HLA-G expression and renal graft acceptance. Am J Transplant. 2006;6:2152–6. doi: 10.1111/j.1600-6143.2006.01417.x. [DOI] [PubMed] [Google Scholar]
  • 10.Rouas-Freiss N, Moreau P, Menier C, et al. Expression of tolerogenic HLA-G molecules in cancer prevents antitumor responses. Semin Cancer Biol. 2007;17:413–21. doi: 10.1016/j.semcancer.2007.07.003. [DOI] [PubMed] [Google Scholar]
  • 11.Onno M, Pangault C, Le Friec G, et al. Modulation of HLA-G antigens expression by human cytomegalovirus: specific induction in activated macrophages harboring human cytomegalovirus infection. J Immunol. 2000;164:6426–34. doi: 10.4049/jimmunol.164.12.6426. [DOI] [PubMed] [Google Scholar]
  • 12.Lafon M, Prehaud C, Megret F, Lafage M, Mouillot G, Roa M, Moreau P, Rouas-Freiss N, Carosella ED. Modulation of HLA-G expression in human neural cells after neurotropic viral infections. J Virol. 2005;79:15226–37. doi: 10.1128/JVI.79.24.15226-15237.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Carosella ED, Favier B, Rouas-Freiss N, Moreau P, Lemaoult J. Beyond the increasing complexity of the immunomodulatory HLA-G molecule. Blood. 2008;111:4862–70. doi: 10.1182/blood-2007-12-127662. [DOI] [PubMed] [Google Scholar]
  • 14.Colonna M, Navarro F, Bellon T, et al. A common inhibitory receptor for major histocompatibility complex class I molecules on human lymphoid and myelomonocytic cells. J Exp Med. 1997;186:1809–18. doi: 10.1084/jem.186.11.1809. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Colonna M, Samaridis J, Cella M, et al. Human myelomonocytic cells express an inhibitory receptor for classical and nonclassical MHC class I molecules. J Immunol. 1998;160:3096–100. [PubMed] [Google Scholar]
  • 16.Rajagopalan S, Long EO. A human histocompatibility leucocyte antigen (HLA)-G-specific receptor expressed on all natural killer cells. J. Exp. Med. 1999;189:1093–9. doi: 10.1084/jem.189.7.1093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.LeMaoult J, Caumartin J, Daouya M, et al. Immune regulation by pretenders: cell-to-cell transfers of HLA-G make effector T cells act as regulatory cells. Blood. 2007;109:2040–8. doi: 10.1182/blood-2006-05-024547. [DOI] [PubMed] [Google Scholar]
  • 18.Caumartin J, Favier B, Daouya M, et al. Trogocytosis-based generation of suppressive NK cells. EMBO J. 2007;26:1423–33. doi: 10.1038/sj.emboj.7601570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Lemaoult J, Zafaranloo K, Le Danff C, et al. HLA-G up-regulates ILT2, ILT3, ILT4, and KIR2DL4 in antigen presenting cells, NK cells, and T cells. FASEB J. 2005 doi: 10.1096/fj.04-1617fje. [DOI] [PubMed] [Google Scholar]
  • 20.Geraghty DE, Koller BH, Orr HT. A human major histocompatibility complex class I gene that encodes a protein with shortened cytoplasmic segment. Proc Natl Acad Sci USA. 1987;84:9145–9. doi: 10.1073/pnas.84.24.9145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Hviid TV. HLA-G in human reproduction: aspects of genetics, function and pregnancy complications. Hum Reprod Update. 2006;12:209–32. doi: 10.1093/humupd/dmi048. [DOI] [PubMed] [Google Scholar]
  • 22.Tan Z, Shon AM, Ober C. Evidence of balancing selection at the HLA-G promoter region. Hum Mol Genet. 2005;14:3619–28. doi: 10.1093/hmg/ddi389. [DOI] [PubMed] [Google Scholar]
  • 23.Hviid TV, Rizzo R, Melchiorri L, et al. Polymorphism in the 5′ upstream regulatory and 3′ untranslated regions of the HLA-G gene in relation to soluble HLA-G and IL-10 expression. Hum Immunol. 2006;67:53–62. doi: 10.1016/j.humimm.2005.12.003. [DOI] [PubMed] [Google Scholar]
  • 24.Tan Z, Randall G, Fan J, et al. Allele-Specific Targeting of microRNAs to HLA-G and Risk of Asthma. Am J Hum Genet. 2007;81:829–34. doi: 10.1086/521200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Lajoie J, Jeanneau A, Faucher MC, et al. Characterisation of five novel HLA-G alleles with coding DNA base changes. Tissue Antigens. 2008;72:502–4. doi: 10.1111/j.1399-0039.2008.01130.x. [DOI] [PubMed] [Google Scholar]
  • 26.Ishitani A, Geraghty DE. Alternative splicing of HLA-G transcripts yields proteins with primary structures resembling both class I and class II antigens. Proc Natl Acad Sci USA. 1992;89:3947–51. doi: 10.1073/pnas.89.9.3947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Kirszenbaum M, Moreau P, Gluckman E, et al. An alternatively spliced form of HLA-G mRNA in human trophoblasts and evidence for the presence of HLA-G transcript in adult lymphocytes. Proc Natl Acad Sci USA. 1994;91:4209–13. doi: 10.1073/pnas.91.10.4209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Paul P, Cabestre FA, Ibrahim EC, et al. Identification of HLA-G7 as a new splice variant of the HLA-G mRNA and expression of soluble HLA-G5, -G6, and -G7 transcripts in human transfected cells. Hum. Immunol. 2000;61:1138–49. doi: 10.1016/s0198-8859(00)00197-x. [DOI] [PubMed] [Google Scholar]
  • 29.Clements CS, Kjer-Nielsen L, Kostenko L, et al. Crystal structure of HLA-G: a nonclassical MHC class I molecule expressed at the fetal-maternal interface. Proc Natl Acad Sci USA. 2005;102:3360–5. doi: 10.1073/pnas.0409676102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Park B, Lee S, Kim E, et al. The truncated cytoplasmic tail of HLA-G serves a quality-control function in post-ER compartments. Immunity. 2001;15:213–24. doi: 10.1016/s1074-7613(01)00179-0. [DOI] [PubMed] [Google Scholar]
  • 31.Moreau P, Carosella E, Teyssier M, et al. Soluble HLA-G molecule. An alternatively spliced HLA-G mRNA form candidate to encode it in peripheral blood mononuclear cells and human trophoblasts. Human Immunology. 1995;43:231–6. doi: 10.1016/0198-8859(95)00009-s. [DOI] [PubMed] [Google Scholar]
  • 32.Fujii T, Ishitani A, Geraghty DE. A soluble form of the HLA-G antigen is encoded by a messenger ribonucleic acid containing intron 4. J Immunol. 1994;153:5516–24. [PubMed] [Google Scholar]
  • 33.Park GM, Lee S, Park B, et al. Soluble HLA-G generated by proteolytic shedding inhibits NK-mediated cell lysis. Biochem Biophys Res Commun. 2004;313:606–11. doi: 10.1016/j.bbrc.2003.11.153. [DOI] [PubMed] [Google Scholar]
  • 34.Zidi I, Guillard C, Marcou C, et al. Increase in HLA-G1 proteolytic shedding by tumor cells: a regulatory pathway controlled by NF-kappaB inducers. Cell Mol Life Sci. 2006 doi: 10.1007/s00018-006-6341-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Gonen-Gross T, Achdout H, Arnon TI, et al. The CD85J/leukocyte inhibitory receptor-1 distinguishes between conformed and beta2-microglobulin-free HLA-G molecules. J Immunol. 2005;175:4866–74. doi: 10.4049/jimmunol.175.8.4866. [DOI] [PubMed] [Google Scholar]
  • 36.Apps R, Gardner L, Sharkey AM, et al. A homodimeric complex of HLA-G on normal trophoblast cells modulates antigen-presenting cells via LILRB1. Eur J Immunol. 2007;37:1924–37. doi: 10.1002/eji.200737089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Morales PJ, Pace JL, Platt JS, et al. Synthesis of beta(2)-microglobulin-free, disulphide-linked HLA-G5 homodimers in human placental villous cytotrophoblast cells. Immunology. 2007;122:179–88. doi: 10.1111/j.1365-2567.2007.02623.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Shiroishi M, Tsumoto K, Amano K, et al. Human inhibitory receptors Ig-like transcript 2 (ILT2) and ILT4 compete with CD8 for MHC class I binding and bind preferentially to HLA-G. Proc Natl Acad Sci USA. 2003;100:8856–61. doi: 10.1073/pnas.1431057100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Jurisicova A, Casper RF, MacLusky NJ, et al. HLA-G expression during preimplantation human embryo development. Proc Natl Acad Sci USA. 1996;93:161–5. doi: 10.1073/pnas.93.1.161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Menicucci A, Noci I, Fuzzi B, et al. Non-classic sHLA class I in human oocyte culture medium. Hum. Immunol. 1999;60:1054–7. doi: 10.1016/s0198-8859(99)00108-1. [DOI] [PubMed] [Google Scholar]
  • 41.Houlihan JM, Biro PA, Harper HM, et al. The human amnion is a site of MHC class Ib expression: evidence for the expression of HLA-E and HLA-G. J Immunol. 1995;154:5665–74. [PubMed] [Google Scholar]
  • 42.Hammer A, Hutter H, Blaschitz A, et al. Amnion epithelial cells, in contrast to trophoblast cells, express all classical HLA class I molecules together with HLA-G. Am J Reprod Immunol. 1997;37:161–71. doi: 10.1111/j.1600-0897.1997.tb00208.x. [DOI] [PubMed] [Google Scholar]
  • 43.Blaschitz A, Lenfant F, Mallet V, et al. Endothelial cells in chorionic fetal vessels of first trimester placenta express HLA-G. Eur J Immunol. 1997;27:3380–8. doi: 10.1002/eji.1830271237. [DOI] [PubMed] [Google Scholar]
  • 44.Crisa L, Mc Master MT, Ishii JK, et al. Identification of a thymic epithelial cell subset sharing expression of the class Ib HLA-G molecule with fetal trophoblasts. J Exp Med. 1997;186:289–98. doi: 10.1084/jem.186.2.289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Mallet V, Blaschitz A, Crisa L, et al. HLA-G in the human thymus: a subpopulation of medullary epithelial but not CD83(+) dendritic cells expresses HLA-G as a membrane-bound and soluble protein. Int Immunol. 1999;11:889–98. doi: 10.1093/intimm/11.6.889. [DOI] [PubMed] [Google Scholar]
  • 46.Le Discorde M, Moreau P, Sabatier P, et al. Expression of HLA-G in human cornea, an immune-privileged tissue. Hum Immunol. 2003;64:1039–44. doi: 10.1016/j.humimm.2003.08.346. [DOI] [PubMed] [Google Scholar]
  • 47.Cirulli V, Zalatan J, McMaster M, et al. The class I HLA repertoire of pancreatic islets comprises the nonclassical class Ib antigen HLA-G. Diabetes. 2006;55:1214–22. doi: 10.2337/db05-0731. [DOI] [PubMed] [Google Scholar]
  • 48.Ito T, Ito N, Saathoff M, et al. Immunology of the human nail apparatus: the nail matrix is a site of relative immune privilege. J Invest Dermatol. 2005;125:1139–48. doi: 10.1111/j.0022-202X.2005.23927.x. [DOI] [PubMed] [Google Scholar]
  • 49.Blanco O, Tirado I, Munoz-Fernandez R, et al. Human decidual stromal cells express HLA-G: effects of cytokines and decidualization. Hum Reprod. 2008;23:144–52. doi: 10.1093/humrep/dem326. [DOI] [PubMed] [Google Scholar]
  • 50.Yang Y, Chu W, Geraghty DE, et al. Expression of HLA-G in human mononuclear phagocytes and selective induction by IFN-gamma. J Immunol. 1996;156:4224–31. [PubMed] [Google Scholar]
  • 51.Ulbrecht M, Rehberger B, Strobel I, et al. HLA-G: expression in human keratinocytes in vitro and in human skin in vivo. Eur J Immunol. 1994;24:176–80. doi: 10.1002/eji.1830240127. [DOI] [PubMed] [Google Scholar]
  • 52.Le Rond S, Le Maoult J, Creput C, et al. Alloreactive CD4+ and CD8+ T cells express the immunotolerant HLA-G molecule in mixed lymphocyte reactions: in vivo implications in transplanted patients. Eur J Immunol. 2004;34:649–60. doi: 10.1002/eji.200324266. [DOI] [PubMed] [Google Scholar]
  • 53.Fanchin R, Gallot V, Rouas-Freiss N, et al. Implication of HLA-G in human embryo implantation. Hum Immunol. 2007;68:259–63. doi: 10.1016/j.humimm.2006.11.002. [DOI] [PubMed] [Google Scholar]
  • 54.Urosevic M, Kurrer MO, Kamarashev J, et al. Human leukocyte antigen G up-regulation in lung cancer associates with high-grade histology, human leukocyte antigen class I loss and interleukin-10 production. Am J Pathol. 2001;159:817–24. doi: 10.1016/S0002-9440(10)61756-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Pangault C, Le Friec G, Caulet-Maugendre S, et al. Lung macrophages and dendritic cells express HLA-G molecules in pulmonary diseases. Hum Immunol. 2002;63:83–90. doi: 10.1016/s0198-8859(01)00373-1. [DOI] [PubMed] [Google Scholar]
  • 56.Lefebvre S, Antoine M, Uzan S, et al. Specific activation of the non-classical class I histocompatibility HLA-G antigen and expression of the ILT2 inhibitory receptor in human breast cancer. J Pathol. 2002;196:266–74. doi: 10.1002/path.1039. [DOI] [PubMed] [Google Scholar]
  • 57.Singer G, Rebmann V, Chen YC, et al. HLA-G is a potential tumor marker in malignant ascites. Clin Cancer Res. 2003;9:4460–4. [PubMed] [Google Scholar]
  • 58.Hansel DE, Rahman A, Wilentz RE, et al. HLA-G upregulation in pre-malignant and malignant lesions of the gastrointestinal tract. Int J Gastrointest Cancer. 2005;35:15–24. doi: 10.1385/ijgc:35:1:015. [DOI] [PubMed] [Google Scholar]
  • 59.Yie SM, Yang H, Ye SR, et al. Expression of HLA-G is associated with prognosis in esophageal squamous cell carcinoma. Am J Clin Pathol. 2007;128:1002–9. doi: 10.1309/JNCW1QLDFB6AM9WE. [DOI] [PubMed] [Google Scholar]
  • 60.Rouas-Freiss N, Bruel S, Menier C, et al. Switch of HLA-G alternative splicing in a melanoma cell line causes loss of HLA-G1 expression and sensitivity to NK lysis. Int J Cancer. 2005;117:114–22. doi: 10.1002/ijc.21151. [DOI] [PubMed] [Google Scholar]
  • 61.Bukur J, Malenica B, Huber C, et al. Altered expression of nonclassical HLA class Ib antigens in human renal cell carcinoma and its association with impaired immune response. Hum Immunol. 2003;64:1081–92. doi: 10.1016/j.humimm.2003.08.350. [DOI] [PubMed] [Google Scholar]
  • 62.Copeman J, Han RN, Caniggia I, et al. Posttranscriptional regulation of human leukocyte antigen G during human extravillous cytotrophoblast differentiation. Biol Reprod. 2000;62:1543–50. doi: 10.1095/biolreprod62.6.1543. [DOI] [PubMed] [Google Scholar]
  • 63.Onno M, Guillaudeux T, Amiot L, et al. The HLA-G gene is expressed at a low mRNA level in different human cells and tissues. Hum Immunol. 1994;41:79–86. doi: 10.1016/0198-8859(94)90089-2. [DOI] [PubMed] [Google Scholar]
  • 64.Teyssier M, Bensussan A, Kirszenbaum M, et al. Natural killer cells are the unique lymphocyte cell subset which do not express HLA-G. Nat Immun. 1995;14:262–70. [PubMed] [Google Scholar]
  • 65.Moreau P, Lefebvre S, Gourand L, et al. Specific binding of nuclear factors to the HLA-G gene promoter correlates with a lack of HLA-G transcripts in first trimester human fetal liver. Hum Immunol. 1998;59:751–7. doi: 10.1016/s0198-8859(98)00081-0. [DOI] [PubMed] [Google Scholar]
  • 66.Paul P, Cabestre FA, Le Gal FA, et al. Heterogeneity of HLA-G gene transcription and protein expression in malignant melanoma biopsies. Cancer Res. 1999;59:1954–60. [PubMed] [Google Scholar]
  • 67.Solier C, Mallet V, Lenfant F, et al. HLA-G unique promoter region: functional implications. Immunogenetics. 2001;53:617–25. doi: 10.1007/s00251-001-0373-0. [DOI] [PubMed] [Google Scholar]
  • 68.Gobin SJ, van den Elsen PJ. The regulation of HLA class I expression: is HLA-G the odd one out? Semin Cancer Biol. 1999;9:55–9. doi: 10.1006/scbi.1998.0109. [DOI] [PubMed] [Google Scholar]
  • 69.Gobin SJ, van den Elsen PJ. Transcriptional regulation of the MHC class Ib genes HLA-E, HLA-F, and HLA-G. Hum Immunol. 2000;61:1102–7. doi: 10.1016/s0198-8859(00)00198-1. [DOI] [PubMed] [Google Scholar]
  • 70.Schmidt CM, Ehlenfeldt RG, Athanasiou MC, et al. Extraembryonic expression of the human MHC class I gene HLA-G in transgenic mice. Evidence for a positive regulatory region located 1 kilobase 5′ to the start site of transcription. J Immunol. 1993;151:2633–45. [PubMed] [Google Scholar]
  • 71.van den Elsen PJ, Gobin SJ, van Eggermond MC, et al. Regulation of MHC class I and II gene transcription: differences and similarities. Immunogenetics. 1998;48:208–21. doi: 10.1007/s002510050425. [DOI] [PubMed] [Google Scholar]
  • 72.Gobin SJ, Keijsers V, van Zutphen M, et al. The role of enhancer A in the locus-specific transactivation of classical and nonclassical HLA class I genes by nuclear factor kappa B. J Immunol. 1998;161:2276–83. [PubMed] [Google Scholar]
  • 73.Gobin SJ, van Zutphen M, Woltman AM, et al. Transactivation of classical and nonclassical HLA class I genes through the IFN-stimulated response element. J Immunol. 1999;163:1428–34. [PubMed] [Google Scholar]
  • 74.Reith W, Mach B. The bare lymphocyte syndrome and the regulation of MHC expression. Annu Rev Immunol. 2001;19:331–73. doi: 10.1146/annurev.immunol.19.1.331. [DOI] [PubMed] [Google Scholar]
  • 75.Steimle V, Durand B, Barras E, et al. A novel DNA-binding regulatory factor is mutated in primary MHC class II deficiency (bare lymphocyte syndrome) Genes Dev. 1995;9:1021–32. doi: 10.1101/gad.9.9.1021. [DOI] [PubMed] [Google Scholar]
  • 76.Durand B, Sperisen P, Emery P, et al. RFXAP, a novel subunit of the RFX DNA binding complex is mutated in MHC class II deficiency. EMBO J. 1997;16:1045–55. doi: 10.1093/emboj/16.5.1045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Masternak K, Barras E, Zufferey M, et al. A gene encoding a novel RFX-associated transactivator is mutated in the majority of MHC class II deficiency patients. Nat Genet. 1998;20:273–7. doi: 10.1038/3081. [DOI] [PubMed] [Google Scholar]
  • 78.Nagarajan UM, Louis-Plence P, DeSandro A, et al. RFX-B is the gene responsible for the most common cause of the bare lymphocyte syndrome, an MHC class II immunodeficiency. Immunity. 1999;10:153–62. doi: 10.1016/s1074-7613(00)80016-3. [DOI] [PubMed] [Google Scholar]
  • 79.Moreno CS, Beresford GW, Louis-Plence P, et al. CREB regulates MHC class II expression in a CIITA-dependent manner. Immunity. 1999;10:143–51. doi: 10.1016/s1074-7613(00)80015-1. [DOI] [PubMed] [Google Scholar]
  • 80.Mantovani R. The molecular biology of the CCAAT-binding factor NF-Y. Gene. 1999;239:15–27. doi: 10.1016/s0378-1119(99)00368-6. [DOI] [PubMed] [Google Scholar]
  • 81.Gobin SJ, Peijnenburg A, Keijsers V, et al. Site alpha is crucial for two routes of IFN gamma-induced MHC class I transactivation: the ISRE-mediated route and a novel pathway involving CIITA. Immunity. 1997;6:601–11. doi: 10.1016/s1074-7613(00)80348-9. [DOI] [PubMed] [Google Scholar]
  • 82.Gobin SJ, Peijnenburg A, van Eggermond M, et al. The RFX complex is crucial for the constitutive and CIITA-mediated transactivation of MHC class I and beta2-microglobulin genes. Immunity. 1998;9:531–41. doi: 10.1016/s1074-7613(00)80636-6. [DOI] [PubMed] [Google Scholar]
  • 83.van den Elsen PJ, Peijnenburg A, van Eggermond MC, et al. Shared regulatory elements in the promoters of MHC class I and class II genes. Immunol Today. 1998;19:308–12. doi: 10.1016/s0167-5699(98)01287-0. [DOI] [PubMed] [Google Scholar]
  • 84.Martin BK, Chin KC, Olsen JC, et al. Induction of MHC class I expression by the MHC class II transactivator CIITA. Immunity. 1997;6:591–600. doi: 10.1016/s1074-7613(00)80347-7. [DOI] [PubMed] [Google Scholar]
  • 85.Lefebvre S, Moreau P, Dausset J, et al. Downregulation of HLA class I gene transcription in choriocarcinoma cells is controlled by the proximal promoter element and can be reversed by CIITA. Placenta. 1999;20:293–301. doi: 10.1053/plac.1998.0380. [DOI] [PubMed] [Google Scholar]
  • 86.Gobin SJ, van Zutphen M, Westerheide SD, et al. The MHC-specific enhanceosome and its role in MHC class I and beta(2)-microglobulin gene transactivation. J Immunol. 2001;167:5175–84. doi: 10.4049/jimmunol.167.9.5175. [DOI] [PubMed] [Google Scholar]
  • 87.Rousseau P, Paul P, O’Brien M, et al. The X1 box of HLA-G promoter is a target site for RFX and Sp1 factors. Hum Immunol. 2000;61:1132–7. doi: 10.1016/s0198-8859(00)00199-3. [DOI] [PubMed] [Google Scholar]
  • 88.Rousseau P, Masternak K, Krawczyk M, et al. In vivo, RFX5 binds differently to the human leucocyte antigen-E, -F, and -G gene promoters and participates in HLA class I protein expression in a cell type-dependent manner. Immunology. 2004;111:53–65. doi: 10.1111/j.1365-2567.2003.01783.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Sim BC, Hui KM. A HLA class I cis-regulatory element whose activity can be modulated by hormones. Int J Cancer. 1994;59:646–56. doi: 10.1002/ijc.2910590512. [DOI] [PubMed] [Google Scholar]
  • 90.Williams KL, Taxman DJ, Linhoff MW, et al. Cutting edge: monarch-1: a pyrin/nucleotide-binding domain/leucine-rich repeat protein that controls classical and nonclassical MHC class I genes. J Immunol. 2003;170:5354–8. doi: 10.4049/jimmunol.170.11.5354. [DOI] [PubMed] [Google Scholar]
  • 91.Gobin SJ, Keijsers V, Cheong C, et al. Transcriptional regulation of HLA-G. Transplant Proc. 1999;31:1857–9. doi: 10.1016/s0041-1345(99)00188-8. [DOI] [PubMed] [Google Scholar]
  • 92.Yelavarthi KK, Schmidt CM, Ehlenfeldt RG, et al. Cellular distribution of HLA-G mRNA in transgenic mouse placentas. Journal of Immunology. 1993;151:3638–45. [PubMed] [Google Scholar]
  • 93.Strauss EC, Orkin SH. In vivo protein-DNA interactions at hypersensitive site 3 of the human beta-globin locus control region. Proc Natl Acad Sci USA. 1992;89:5809–13. doi: 10.1073/pnas.89.13.5809. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Schmidt CM, Orr HT. HLA-G transgenic mice: a model for studying expression and function at the maternal/fetal interface. Immunol. Rev. 1995;147:53–65. doi: 10.1111/j.1600-065x.1995.tb00087.x. [DOI] [PubMed] [Google Scholar]
  • 95.Moreau P, Paul P, Gourand L, et al. HLA-G gene transcriptional regulation in trophoblasts and blood cells: differential binding of nuclear factors to a regulatory element located 1.1 kb from exon 1. Hum Immunol. 1997;52:41–6. doi: 10.1016/S0198-8859(96)00242-X. [DOI] [PubMed] [Google Scholar]
  • 96.Gobin SJ, Biesta P, de Steenwinkel JE, et al. HLA-G transactivation by cAMP-response element-binding protein (CREB). An alternative transactivation pathway to the conserved major histocompatibility complex (MHC) class I regulatory routes. J Biol Chem. 2002;277:39525–31. doi: 10.1074/jbc.M112273200. [DOI] [PubMed] [Google Scholar]
  • 97.Yang Y, Geraghty DE, Hunt JS. Cytokine regulation of HLA-G expression in human trophoblast cell lines. J. Reprod. Immunol. 1995;29:179–95. doi: 10.1016/0165-0378(95)00942-e. [DOI] [PubMed] [Google Scholar]
  • 98.Chu W, Gao J, Murphy WJ, et al. A candidate interferon-gamma activated site (GAS element) in the HLA-G promoter does not bind nuclear proteins. Hum Immunol. 1999;60:1113–8. doi: 10.1016/s0198-8859(99)00091-9. [DOI] [PubMed] [Google Scholar]
  • 99.Wiendl H, Mitsdoerffer M, Hofmeister V, et al. A functional role of HLA-G expression in human gliomas: an alternative strategy of immune escape. J Immunol. 2002;168:4772–80. doi: 10.4049/jimmunol.168.9.4772. [DOI] [PubMed] [Google Scholar]
  • 100.Lefebvre S, Berrih-Aknin S, Adrian F, et al. A specific interferon (IFN)-stimulated response element of the distal HLA-G promoter binds IFN-regulatory factor 1 and mediates enhancement of this nonclassical class I gene by IFN-beta. J Biol Chem. 2001;276:6133–9. doi: 10.1074/jbc.M008496200. [DOI] [PubMed] [Google Scholar]
  • 101.Chu W, Yang Y, Geraghty DE, et al. Interferons enhance HLA-G mRNA and protein in transfected mouse fibroblasts. J Reprod Immunol. 1999;42:1–15. doi: 10.1016/s0165-0378(98)00077-1. [DOI] [PubMed] [Google Scholar]
  • 102.Morimoto RI, Santoro MG. Stress-inducible responses and heat shock proteins: new pharmacologic targets for cytoprotection. Nat Biotechnol. 1998;16:833–8. doi: 10.1038/nbt0998-833. [DOI] [PubMed] [Google Scholar]
  • 103.Ibrahim EC, Morange M, Dausset J, et al. Heat shock and arsenite induce expression of the nonclassical class I histocompatibility HLA-G gene in tumor cell lines. Cell Stress Chaperones. 2000;5:207–18. doi: 10.1379/1466-1268(2000)005<0207:hsaaie>2.0.co;2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Yie SM, Li LH, Li GM, et al. Progesterone enhances HLA-G gene expression in JEG-3 choriocarcinoma cells and human cytotrophoblasts in vitro. Hum Reprod. 2006;21:46–51. doi: 10.1093/humrep/dei305. [DOI] [PubMed] [Google Scholar]
  • 105.Yie SM, Xiao R, Librach CL. Progesterone regulates HLA-G gene expression through a novel progesterone response element. Hum Reprod. 2006;21:2538–44. doi: 10.1093/humrep/del126. [DOI] [PubMed] [Google Scholar]
  • 106.Stewart CL, Kaspar P, Brunet LJ, et al. Blastocyst implantation depends on maternal expression of leukaemia inhibitory factor. Nature. 1992;359:76–9. doi: 10.1038/359076a0. [DOI] [PubMed] [Google Scholar]
  • 107.Nachtigall MJ, Kliman HJ, Feinberg RF, et al. The effect of leukemia inhibitory factor (LIF) on trophoblast differentiation: a potential role in human implantation. J Clin Endocrinol Metab. 1996;81:801–6. doi: 10.1210/jcem.81.2.8636307. [DOI] [PubMed] [Google Scholar]
  • 108.Bamberger AM, Jenatschke S, Schulte HM, et al. Leukemia inhibitory factor (LIF) stimulates the human HLA-G promoter in JEG3 choriocarcinoma cells. J Clin Endocrinol Metab. 2000;85:3932–6. doi: 10.1210/jcem.85.10.6849. [DOI] [PubMed] [Google Scholar]
  • 109.Shido F, Ito T, Nomura S, et al. Endoplasmic reticulum aminopeptidase-1 mediates leukemia inhibitory factor-induced cell surface human leukocyte antigen-G expression in JEG-3 choriocarcinoma cells. Endocrinology. 2006;147:1780–8. doi: 10.1210/en.2005-1449. [DOI] [PubMed] [Google Scholar]
  • 110.Thiagalingam A, De Bustros A, Borges M, et al. RREB-1, a novel zinc finger protein, is involved in the differentiation response to Ras in human medullary thyroid carcinomas. Mol Cell Biol. 1996;16:5335–45. doi: 10.1128/mcb.16.10.5335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Shi Y, Sawada J, Sui G, et al. Coordinated histone modifications mediated by a CtBP co-repressor complex. Nature. 2003;422:735–8. doi: 10.1038/nature01550. [DOI] [PubMed] [Google Scholar]
  • 112.Rebmann V, van der Ven K, Passler M, et al. Association of soluble HLA-G plasma levels with HLA-G alleles. Tissue Antigens. 2001;57:15–21. doi: 10.1034/j.1399-0039.2001.057001015.x. [DOI] [PubMed] [Google Scholar]
  • 113.Hviid TV, Hylenius S, Rorbye C, et al. HLA-G allelic variants are associated with differences in the HLA-G mRNA isoform profile and HLA-G mRNA levels. Immunogenetics. 2003;55:63–79. doi: 10.1007/s00251-003-0547-z. [DOI] [PubMed] [Google Scholar]
  • 114.O’Brien M, McCarthy T, Jenkins D, et al. Altered HLA-G transcription in pre-eclampsia is associated with allele specific inheritance: possible role of the HLA-G gene in susceptibility to the disease. Cell Mol Life Sci. 2001;58:1943–9. doi: 10.1007/PL00000828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Hiby S, King A, Sharkey A, et al. Molecular studies of trophoblast HLA-G: polymorphism, isoforms, imprinting and expression in preimplantation embryo. Tissue Antigens. 1999;53:1–13. doi: 10.1034/j.1399-0039.1999.530101.x. [DOI] [PubMed] [Google Scholar]
  • 116.Rousseau P, Le Discorde M, Mouillot G, et al. The 14 bp deletion-insertion polymorphism in the 3′ UT region of the HLA-G gene influences HLA-G mRNA stability. Hum Immunol. 2003;64:1005–10. doi: 10.1016/j.humimm.2003.08.347. [DOI] [PubMed] [Google Scholar]
  • 117.Hviid TV, Rizzo R, Christiansen OB, et al. HLA-G and IL-10 in serum in relation to HLA-G genotype and polymorphisms. Immunogenetics. 2004;56:135–41. doi: 10.1007/s00251-004-0673-2. [DOI] [PubMed] [Google Scholar]
  • 118.Auboeuf D, Honig A, Berget SM, et al. Coordinate regulation of transcription and splicing by steroid receptor coregulators. Science. 2002;298:416–9. doi: 10.1126/science.1073734. [DOI] [PubMed] [Google Scholar]
  • 119.Ferretti C, Bruni L, Dangles-Marie V, et al. Molecular circuits shared by placental and cancer cells, and their implications in the proliferative, invasive and migratory capacities of trophoblasts. Hum Reprod Update. 2007;13:121–41. doi: 10.1093/humupd/dml048. [DOI] [PubMed] [Google Scholar]
  • 120.Orsi NM, Gopichandran N, Ekbote UV, et al. Murine serum cytokines throughout the estrous cycle, pregnancy and post partum period. Anim Reprod Sci. 2006;96:54–65. doi: 10.1016/j.anireprosci.2005.11.010. [DOI] [PubMed] [Google Scholar]
  • 121.Orsi NM, Tribe RM. Cytokine networks and the regulation of uterine function in pregnancy and parturition. J Neuroendocrinol. 2008;20:462–9. doi: 10.1111/j.1365-2826.2008.01668.x. [DOI] [PubMed] [Google Scholar]
  • 122.Staun-Ram E, Shalev E. Human trophoblast function during the implantation process. Reprod Biol Endocrinol. 2005;3:56–68. doi: 10.1186/1477-7827-3-56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Moreau P, Adrian-Cabestre F, Menier C, et al. IL-10 selectively induces HLA-G expression in human trophoblasts and monocytes. Int Immunol. 1999;11:803–11. doi: 10.1093/intimm/11.5.803. [DOI] [PubMed] [Google Scholar]
  • 124.Moreau P, Faure O, Lefebvre S, et al. Glucocorticoid hormones upregulate levels of HLA-G transcripts in trophoblasts. Transplant Proc. 2001;33:2277–80. doi: 10.1016/s0041-1345(01)01990-x. [DOI] [PubMed] [Google Scholar]
  • 125.Cadet P, Rady PL, Tyring SK, et al. Interleukin-10 messenger ribonucleic acid in human placenta: implications of a role for interleukin-10 in fetal allograft protection. Am J Obstet Gynecol. 1995;173:25–9. doi: 10.1016/0002-9378(95)90164-7. [DOI] [PubMed] [Google Scholar]
  • 126.Dunker K, Schlaf G, Bukur J, et al. Expression and regulation of non-classical HLA-G in renal cell carcinoma. Tissue Antigens. 2008;72:137–48. doi: 10.1111/j.1399-0039.2008.01090.x. [DOI] [PubMed] [Google Scholar]
  • 127.Sebti Y, Le Friec G, Pangault C, et al. Soluble HLA-G molecules are increased in lymphoproliferative disorders. Hum Immunol. 2003;64:1093–101. doi: 10.1016/j.humimm.2003.08.345. [DOI] [PubMed] [Google Scholar]
  • 128.Makris A, Xu B, Yu B, et al. Placental deficiency of interleukin-10 (IL-10) in preeclampsia and its relationship to an IL10 promoter polymorphism. Placenta. 2006;27:445–51. doi: 10.1016/j.placenta.2005.05.003. [DOI] [PubMed] [Google Scholar]
  • 129.Goldman-Wohl DS, Ariel I, Greenfield C, et al. HLA-G expression in extravillous trophoblasts is an intrinsic property of cell differentiation: a lesson learned from ectopic pregnancies. Mol Hum Reprod. 2000;6:535–40. doi: 10.1093/molehr/6.6.535. [DOI] [PubMed] [Google Scholar]
  • 130.Urosevic M, Dummer R. HLA-g and IL-10 expression in human cancer–different stories with the same message. Semin Cancer Biol. 2003;13:337–42. doi: 10.1016/s1044-579x(03)00024-5. [DOI] [PubMed] [Google Scholar]
  • 131.Spencer JV, Lockridge KM, Barry PA, et al. Potent immunosuppressive activities of cytomegalovirus-encoded interleukin-10. J Virol. 2002;76:1285–92. doi: 10.1128/JVI.76.3.1285-1292.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Gros F, Sebti Y, de Guibert S, et al. Soluble HLA-G molecules increase during acute leukemia, especially in subtypes affecting monocytic and lymphoid lineages. Neoplasia. 2006;8:223–30. doi: 10.1593/neo.05703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Amiot L, Onno M, Drenou B, et al. HLA-G class I gene expression in normal and malignant hematopoietic cells. Hum Immunol. 1998;59:524–8. doi: 10.1016/s0198-8859(98)00041-x. [DOI] [PubMed] [Google Scholar]
  • 134.Onno M, Le Friec G, Pangault C, et al. Modulation of HLA-G antigens expression in myelomonocytic cells. Hum Immunol. 2000;61:1086–94. doi: 10.1016/s0198-8859(00)00191-9. [DOI] [PubMed] [Google Scholar]
  • 135.Genbacev O, Zhou Y, Ludlow JW, et al. Regulation of human placental development by oxygen tension. Science. 1997;277:1669–72. doi: 10.1126/science.277.5332.1669. [DOI] [PubMed] [Google Scholar]
  • 136.Pouyssegur J, Dayan F, Mazure NM. Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature. 2006;441:437–43. doi: 10.1038/nature04871. [DOI] [PubMed] [Google Scholar]
  • 137.Mazure NM, Brahimi-Horn MC, Berta MA, et al. HIF-1: master and commander of the hypoxic world. A pharmacological approach to its regulation by siRNAs. Biochem Pharmacol. 2004;68:971–80. doi: 10.1016/j.bcp.2004.04.022. [DOI] [PubMed] [Google Scholar]
  • 138.Brahimi-Horn MC, Chiche J, Pouyssegur J. Hypoxia and cancer. J Mol Med. 2007;85:1301–7. doi: 10.1007/s00109-007-0281-3. [DOI] [PubMed] [Google Scholar]
  • 139.Semenza GL. Regulation of mammalian O2 homeostasis by hypoxia-inducible factor 1. Annu Rev Cell Dev Biol. 1999;15:551–78. doi: 10.1146/annurev.cellbio.15.1.551. [DOI] [PubMed] [Google Scholar]
  • 140.Gray MJ, Zhang J, Ellis LM, et al. HIF-1alpha, STAT3, CBP/p300 and Ref-1/APE are components of a transcriptional complex that regulates Src-dependent hypoxia-induced expression of VEGF in pancreatic and prostate carcinomas. Oncogene. 2005;24:3110–20. doi: 10.1038/sj.onc.1208513. [DOI] [PubMed] [Google Scholar]
  • 141.Chang CC, Ferrone S. HLA-G in melanoma: can the current controversies be solved? Semin Cancer Biol. 2003;13:361–9. doi: 10.1016/s1044-579x(03)00027-0. [DOI] [PubMed] [Google Scholar]
  • 142.Mouillot G, Marcou C, Zidi I, et al. Hypoxia modulates HLA-G gene expression in tumor cells. Hum Immunol. 2007;68:277–85. doi: 10.1016/j.humimm.2006.10.016. [DOI] [PubMed] [Google Scholar]
  • 143.Nagamatsu T, Fujii T, Yamashita T, et al. Hypoxia does not reduce HLA-G expression on extravillous cytotrophoblasts. J Reprod Immunol. 2004;63:85–95. doi: 10.1016/j.jri.2004.07.001. [DOI] [PubMed] [Google Scholar]
  • 144.Polakova K, Bandzuchova E, Tirpakova J, et al. Modulation of HLA-G expression. Neoplasma. 2007;54:455–62. [PubMed] [Google Scholar]
  • 145.Kilburn BA, Wang J, Duniec-Dmuchowski ZM, et al. Extracellular matrix composition and hypoxia regulate the expression of HLA-G and integrins in a human trophoblast cell line. Biol Reprod. 2000;62:739–47. doi: 10.1095/biolreprod62.3.739. [DOI] [PubMed] [Google Scholar]
  • 146.Berger SL. The complex language of chromatin regulation during transcription. Nature. 2007;447:407–12. doi: 10.1038/nature05915. [DOI] [PubMed] [Google Scholar]
  • 147.Nafee TM, Farrell WE, Carroll WD, et al. Epigenetic control of fetal gene expression. BJOG. 2008;115:158–68. doi: 10.1111/j.1471-0528.2007.01528.x. [DOI] [PubMed] [Google Scholar]
  • 148.Chelbi ST, Vaiman D. Genetic and epigenetic factors contribute to the onset of preeclampsia. Mol Cell Endocrinol. 2008;282:120–9. doi: 10.1016/j.mce.2007.11.022. [DOI] [PubMed] [Google Scholar]
  • 149.Esteller M. Epigenetics in cancer. N Engl J Med. 2008;358:1148–59. doi: 10.1056/NEJMra072067. [DOI] [PubMed] [Google Scholar]
  • 150.Boucraut J, Guillaudeux T, Alizadeh M, et al. HLA-E is the only class I gene that escapes CpG methylation and is transcriptionally active in the trophoblast-derived human cell line JAR. Immunogenetics. 1993;38:117–30. doi: 10.1007/BF00190899. [DOI] [PubMed] [Google Scholar]
  • 151.Guillaudeux T, Rodriguez AM, Girr M, et al. Methylation status and transcriptional expression of the MHC class I loci in human trophoblast cells from term placenta. J Immunol. 1995;154:3283–99. [PubMed] [Google Scholar]
  • 152.Moreau P, Mouillot G, Rousseau P, et al. HLA-G gene repression is reversed by demethylation. Proc Natl Acad Sci USA. 2003;100:1191–6. doi: 10.1073/pnas.0337539100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Mouillot G, Marcou C, Rousseau P, et al. HLA-G gene activation in tumor cells involves cis-acting epigenetic changes. Int J Cancer. 2005;113:928–36. doi: 10.1002/ijc.20682. [DOI] [PubMed] [Google Scholar]
  • 154.Yan WH, Lin AF, Chang CC, et al. Induction of HLA-G expression in a melanoma cell line OCM-1A following the treatment with 5-aza-2′-deoxycytidine. Cell Res. 2005;15:523–31. doi: 10.1038/sj.cr.7290376. [DOI] [PubMed] [Google Scholar]
  • 155.Seliger B, Schlaf G. Structure, expression and function of HLA-G in renal cell carcinoma. Semin Cancer Biol. 2007;17:444–50. doi: 10.1016/j.semcancer.2007.07.001. [DOI] [PubMed] [Google Scholar]
  • 156.Polakova K, Bandzuchova E, Kuba D, et al. Demethylating agent 5-aza-2′-deoxycytidine activates HLA-G expression in human leukemia cell lines. Leuk Res. 2009;33:518–24. doi: 10.1016/j.leukres.2008.08.003. [DOI] [PubMed] [Google Scholar]
  • 157.Onno M, Amiot L, Bertho N, et al. CpG methylation patterns in the 5′ part of the nonclassical HLA-G gene in peripheral blood CD34+ cells and CD2+ lymphocytes. Tissue Antigens. 1997;49:356–64. doi: 10.1111/j.1399-0039.1997.tb02763.x. [DOI] [PubMed] [Google Scholar]
  • 158.Menendez L, Walker LD, Matyunina LV, et al. Epigenetic changes within the promoter region of the HLA-G gene in ovarian tumors. Mol Cancer. 2008;7:43–54. doi: 10.1186/1476-4598-7-43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Chang CC, Murphy SP, Ferrone S. Differential in vivo and in vitro HLA-G expression in melanoma cells: potential mechanisms. Hum Immunol. 2003;64:1057–63. doi: 10.1016/j.humimm.2003.08.357. [DOI] [PubMed] [Google Scholar]
  • 160.Chan HW, Kurago ZB, Stewart CA, et al. DNA methylation maintains allele-specific KIR gene expression in human natural killer cells. J Exp Med. 2003;197:245–55. doi: 10.1084/jem.20021127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Li G, Weyand CM, Goronzy JJ. Epigenetic mechanisms of age-dependent KIR2DL4 expression in T cells. J Leukoc Biol. 2008;824:824–34. doi: 10.1189/jlb.0807583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Rebmann V, Regel J, Stolke D, et al. Secretion of sHLA-G molecules in malignancies. Semin Cancer Biol. 2003;13:371–7. doi: 10.1016/s1044-579x(03)00028-2. [DOI] [PubMed] [Google Scholar]
  • 163.Ugurel S, Rebmann V, Ferrone S, et al. Soluble human leukocyte antigen–G serum level is elevated in melanoma patients and is further increased by interferon-alpha immunotherapy. Cancer. 2001;92:369–76. doi: 10.1002/1097-0142(20010715)92:2<369::aid-cncr1332>3.0.co;2-u. [DOI] [PubMed] [Google Scholar]
  • 164.Bukur J, Rebmann V, Grosse-Wilde H, et al. Functional role of human leukocyte antigen-g up-regulation in renal cell carcinoma. Cancer Res. 2003;63:4107–11. [PubMed] [Google Scholar]
  • 165.Nagamatsu T, Fujii T, Ishikawa T, et al. A primary cell culture system for human cytotrophoblasts of proximal cytotrophoblast cell columns enabling in vitro acquisition of the extra-villous phenotype. Placenta. 2004;25:153–65. doi: 10.1016/j.placenta.2003.08.015. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Cellular and Molecular Medicine are provided here courtesy of Blackwell Publishing

RESOURCES