Skip to main content
Erschienen in: Acta Neuropathologica Communications 1/2014

Open Access 01.12.2014 | Research

Increased 5-hydroxymethylcytosine and decreased 5-methylcytosine are indicators of global epigenetic dysregulation in diffuse intrinsic pontine glioma

verfasst von: Sama Ahsan, Eric H Raabe, Michael C Haffner, Ajay Vaghasia, Katherine E Warren, Martha Quezado, Leomar Y Ballester, Javad Nazarian, Charles G Eberhart, Fausto J Rodriguez

Erschienen in: Acta Neuropathologica Communications | Ausgabe 1/2014

Abstract

Introduction

Diffuse intrinsic pontine glioma (DIPG) is a malignant pediatric brain tumor associated with dismal outcome. Recent high-throughput molecular studies have shown a high frequency of mutations in histone-encoding genes (H3F3A and HIST1B) and distinctive epigenetic alterations in these tumors. Epigenetic alterations described in DIPG include global DNA hypomethylation. In addition to the generally repressive methylcytosine DNA alteration, 5-hydroxymethylation of cytosine (5hmC) is recognized as an epigenetic mark associated with active chromatin. We hypothesized that in addition to alterations in DNA methylation, that there would be changes in 5hmC. To test this hypothesis, we performed immunohistochemical studies to compare epigenetic alterations in DIPG to extrapontine adult and pediatric glioblastoma (GBM) and normal brain. A total of 124 tumors were scored for histone 3 lysine 27 trimethylation (H3K27me3) and histone 3 lysine 9 trimethylation (H3K9me3) and 104 for 5hmC and 5-methylcytosine (5mC). An H-score was derived by multiplying intensity (0–2) by percentage of positive tumor nuclei (0-100%).

Results

We identified decreased H3K27me3 in the DIPG cohort compared to pediatric GBM (p < 0.01), adult GBM (p < 0.0001) and normal brain (p < 0.0001). H3K9me3 was not significantly different between tumor types. Global DNA methylation as measured by 5mC levels were significantly lower in DIPG compared to pediatric GBM (p < 0.001), adult GBM (p < 0.01), and normal brain (p < 0.01). Conversely, 5hmC levels were significantly higher in DIPG compared to pediatric GBM (p < 0.0001) and adult GBM (p < 0.0001). Additionally, in an independent set of DIPG tumor samples, TET1 and TET3 mRNAs were found to be overexpressed relative to matched normal brain.

Conclusions

Our findings extend the immunohistochemical study of epigenetic alterations in archival tissue to DIPG specimens. Low H3K27me3, decreased 5mC and increased 5hmC are characteristic of DIPG in comparison with extrapontine GBM. In DIPG, the relative imbalance of 5mC compared to 5hmC may represent an opportunity for therapeutic intervention.
Begleitmaterial
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​2051-5960-2-59) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

SA participated in design and coordination of study, scored and collected data, performed the statistical analysis, and drafted the manuscript. ER participated in design and coordination of the study and manuscript writing. Also provided materials necessary for the study. MH optimized and performed immunohistochemistry and participated in the design and coordination of the study. AV optimized and performed immunohistochemistry. KW participated in design and coordination of the study. MQ participated in design and coordination of the study. LB participated in design and coordination of the study. JN participated in design and coordination of the study, collected data, and performed statistical analysis. CE conceived of the study and participated in design and coordination of the study. FR conceived of the study and participated in design and coordination of the study. All authors critically read and approved the final manuscript.
Abkürzungen
5hmC
5-hydroxymethylcytosine
5mC
5-methylcytosine
aGBM
Adult glioblastoma
α-KG
Alpha-ketoglutarate
BER
Base excision repair
DIPG
Diffuse intrinsic pontine glioma
DNMT
DNA methyltransferase
EZH2
Enhancer of zeste homologue 2
GBM
Glioblastoma
H
H-Score
H3K9me3
Histone 3 Lysine 9 trimethylation
H3K27
Histone 3 Lysine at position 27
H3K27me3
Histone 3 Lysine 27 trimethylation
K27M
Lysine at position 27 of histone 3 changed to methionine
PcG
Polycomb group
pGBM
Pediatric glioblastoma
PRC2
Polycomb repressive complex 2
SHH
Hedgehog
TET
Ten eEleven translocation
TDG
Thymine DNA glycosylase.

Introduction

Tumors of the central nervous system are the second most common malignancy in pediatric patients. Diffuse intrinsic pontine gliomas (DIPGs) comprise approximately 10% of pediatric brain tumors and are universally fatal[1]. Due to their location, DIPGs have rarely been biopsied at diagnosis outside of recent clinical trials. No improvements in DIPG outcome have been noted in the past 20 years, perhaps due to the paucity of patient-derived samples and lack of molecular insights leading to novel treatments. Encouragingly, recent molecular and proteomic analyses of autopsy specimens have identified key genetic alterations in DIPG, including amplifications in genes encoding receptor tyrosine kinases (PDGFRA, MET)[2], PDGFRA mutations[3], as well as distinct DIPG subgroups based on Hedgehog (SHH) and MYCN pathway activation[4]. Mutations in the ACVR1 gene, encoding the transforming growth factor-beta (TGF-beta) superfamily member activin, have been reported in approximately 20% of DIPGs[5, 6].
Epigenetic research has added to our understanding of how chromatin remodeling by methylation and acetylation of histones affects gene expression in tumors[7]. Among brain tumors, glioblastomas (GBMs) can be subdivided into 6 groups based on DNA methylation patterns. DIPGs are classified within one group that has relatively hypomethylated DNA and are associated with mutations in genes encoding for histone proteins[811]. Approximately 60% of DIPGs have a mutation in the H3F3A gene which encodes the variant histone 3.3 proteins[11], which is also associated with a worse prognosis[12]. Less frequently, DIPGs have mutations in HIST1H3B, encoding for histone H3.1, while other extrapontine pediatric high-grade gliomas may have alternative mutations in H3F3A (G34R or G34V)[13, 8, 9]. As a result of the common missense mutation (H3F3A), lysine at position 27 (H3K27) is changed to methionine (K27M) in the amino terminal tail of histone 3.3[10]. Post-translational modification of this critical histone regulates gene expression, DNA repair, and maintenance of centromeres/telomeres[8]. Methylation of the H3K27 is mediated by polycomb repressive complex 2 (PRC2), which binds to polycomb group (PcG) target genes and induces the PRC2 component enhancer of zeste homologue 2 (EZH2) methyltransferase to methylate H3K27. Histone methylation creates a repressed state by inhibiting transcription, altering chromatin compaction, and affecting recruitment of DNA methyltransferases (DNMTs)[8, 14]. The H3K27M mutant histone inhibits PRC2 by interacting with EZH2 and suppressing its function[13]. Therefore, as a result of the K27M mutation, histone methylation is decreased at this site[8, 9, 13, 10]. H3.3 is primarily found at transcription sites and telomeres, and the H3K27M mutant H3.3 is associated with transcriptionally active (open) chromatin[8]. Inhibition of histone K27 tri-methylation (H3K27me3) leads to global activation of transcription[8, 13].
One of the most studied epigenetic alterations is methylation of DNA. Mammalian DNA methylation occurs primarily at the 5-position of cytosines (5mC) in CpG dinucleotides. Methylated cytosines at gene promoters with high GC content (CpG islands) are usually associated with transcriptional silencing. Loss of 5mC leads to a redistribution of PRC2 complexes which suggests that 5mC could affect interactions between PRC2 and chromatin[15]. Aberrant recruitment of PRC2 to DNA not usually associated with H3K27me3 may shift PRC2 away from original targets and again promote an active transcriptional state. H3K27me3 is not uniformly increased in these hypomethylated DNA regions, which provides evidence that DNA methylation is one of many epigenetic factors affecting PRC2 function[15, 10]. A hypomethylated genome, in addition to the H3F3A mutation, likely accentuates the transcriptionally active state of DIPGs by disrupting histone methylation at H3K27me3.
The methylation of cytosine in CpG islands is a modification produced by DNMTs. Reversal of 5mC methylation is accomplished in a multistep enzymatic process using Ten Eleven Translocation (TET) enzymes, thymine DNA glycosylase (TDG), and base excision repair (BER)[16, 17]. TET enzymes can convert 5mC in a reaction dependent on alpha-ketoglutarate (α-KG) to 5-hydroxymethylcytosine (5hmC)[16, 18]. 5hmC can then either be further processed by TDG and BER or persist as 5hmC in mammalian genomes[19]. Decreased 5hmC levels have been described in a variety of cancers[20], as well as in high-grade gliomas[21]. 5hmC is often associated with the gene bodies of actively transcribed genes and is considered an epigenetic mark in its own right[17].
These studies raise the possibility that unregulated loss of H3K27me3 through H3K27M mutation and elevated 5hmC could shift normal development into a pathologic state. The association between loss of H3K27me3 and elevated 5hmC in neural development also suggests a regulatory cross talk between these two pathways. Histone 3 lysine 9 methylation (H3K9me3) is another important histone methylation mark implicated in the development of gliomas. Methylation at this site affects global DNA methylation, chromatin compaction, and transcription[22].
In this study we compared epigenetic alterations between DIPG and GBM with respect to patient age and tumor geographical location using archival formalin-fixed paraffin-embedded material to gain a better understanding of epigenetic alterations specific to DIPG.

Materials and methods

Ethics statement

Human brain tumor samples and normal control tissue were obtained at biopsy or autopsy at Johns Hopkins Hospital Department of Pathology, Children’s National Medical Center and National Institutes of Health Center for Cancer Research after Institutional Review Board approval or exemption. The research ethics committee waived the requirement for informed consent for retrospective samples and no informed consent was obtained. The patient data was de-identified prior to inclusion in this study.

Human tissue microarray

DIPG tumor samples were obtained from tissue microarrays created at the National Institutes of Health from rapid autopsy tissue for a total of 24 patients (3 to 15 years of age, with a median age of 7) with each patient having 1–3 representative cores on the array and scored. Tumor samples varied in size with a maximum width of 0.4 cm and maximum length of 1.8 cm. Clinical and pathologic features of this group have been previously published (Additional file1: Table S1)[1]. The subjects were anonymized as part of the study approval and thus therapeutic data was not collected. Tissue microarrays containing 64 adult GBM (22 to 86 years of age, with median age of 55) and 36 pediatric GBM (less than 1year old to 21 years of age, with median age of 13) were used as a comparison group. Adult and pediatric GBM arrays were created by the Johns Hopkins microarray core facility and have been previously characterized (core diameter 0.6 mm)[23]. Samples with two or more scorable cores were included in our dataset. Cores/samples were excluded from scoring and data analysis if the sample was absent, degraded, or no tumor present. Eight single cores from the GBM arrays and one sample from the DIPG arrays were normal brain and used as controls (i.e. the same normal brain samples from patients were used for comparison between the different IHC stains). Only one normal brain was confirmed to be from the pediatric age group.

Immunohistochemistry and scoring

The mouse monoclonal anti-Histone H3 (tri methyl K27) antibody ChIP grade (Abcam, Cat# ab6002, Cambridge, MA) was used at 1:1600 dilution. The antibody was incubated overnight at 4°C. For H3K9me3 detection, rabbit polyclonal anti-Histone H3 (tri methyl K9) antibody ChIP Grade (Abcam, Cat# ab8898, Cambridge, MA) was used at 1:8000 and incubated overnight at 4°C. Secondary antibodies used were anti-rabbit (PV6119) or anti-mouse (PV6114) from Novocastra PowerVision Poly-HR IHC Detection System. Prostate tissue with characteristic staining patterns for H3K27me3 and H3K9me3 were used as positive/negative control, and characteristic vessel wall staining was used as an internal control for H3K27me3 staining. Once antibody concentration had been optimized with prostate tissue, tissue microarrays were stained. 5hmC detection was completed using the rabbit polyclonal 5-hmC specific antibody (Active Motif, Cat# 39769, Carlsbad, CA) at a 1:20,000 dilution for 1 hour at room temperature. The mouse monoclonal 5-methylcytosine specific antibody (Calbiochem, EMD Chemicals Inc., Cat# NA81-50UG, San Diego, CA) at a 1:2000 dilution incubated for 1 hour at room temperature was used for 5mC detection. After immunohistochemical staining, all tissue arrays were counterstained with hematoxylin.
HEK293 cells transfected with expression vectors coding for TET2 or control vectors were used as a standard internal control for all immunolabeling experiments for 5hmC as described previously[20]. HEK293 cells show very low baseline levels of 5hmC. Overexpression of TET2 increases 5hmC levels resulting in detectable immunoreactivity. The consistency of immunolabeling is documented in Additional file2: Figure S1 showing 5hmC staining from 3 independent experiments. Control HEK293 cells never show immunoreactivity, whereas TET2 transfected cells consistently show 5hmC labeling. For 5mC staining the same control slides were used.
Immunoreactivity was assessed by a neuropathologist (FJR) using H-scores (H) (0–200) which were obtained by multiplying intensity of stain (0: no stain, 1: weak stain, 2: strong stain) by percentage (0–100) of neoplastic cells showing the staining intensity[21]. Since most normal terminally differentiated cells show high 5hmC levels, all H-scores were internally normalized to the staining intensities in the non-neoplastic cells. Every case therefore served as its own internal control which greatly reduced inconsistencies due to inadequate tissue fixation.

mRNA expression profiling

Tissue mRNA expression profiling of 6 DIPG tissue samples and adjacent normal sections[4] were reanalyzed for relative TET expression using Partek Genomics Suite v6.6 (Partek Incorporated, St. Louis, MO). Two of these samples are also represented in the TMA. Expression values were extracted from Partek and graphed using Prism Graph-Pad software (San Diego CA). All of the six specimens used for this study had the H3.3K27M mutation.

Statistical analysis

Statistical analysis was completed using Graph-Pad Prism 6 software (Graph-Pad Software, La Jolla, CA). The non-parametric Kruskal-Wallis one way analysis of variance (ANOVA) test was used to identify differences between all tumor subtype’s H-score medians. The Mann–Whitney test was used to analyze differences by rank sum of H scores between two tumor subtypes.

Results

H3K27me3 levels are lower in DIPG compared to adult and pediatric extrapontine GBM

H3K27me3 nuclear immunoreactivity was present in all samples, and was particularly strong in non-neoplastic brain, including neurons, glial and endothelial cells (Figure 1). Decreased H3K27me3, defined as a H-score < 100, was present in 37/116 (32%) scorable tumors, particularly in DIPG 17/24 (71%) followed by pediatric 8/30(27%) and adult GBM 12/62(19%). H3K27me3 median H-score was significantly lower for DIPG (H = 80) compared to pediatric GBM (H = 138) (p <0.01, Mann–Whitney Test), adult GBM (H = 160) (p < 0.0001), and normal brain (H-score 200) (p < 0.0001) tissue (Figure 2). A significant difference was also noted in immunoreactivity between control and adult GBM (p < 0.0001) and also control and pediatric GBM (p < 0.0001). No difference was noted between adult and pediatric GBM immunoreactivity for H3K27me3 (p > 0.05). Conversely, H3K9me3 immunoreactivity was relatively preserved in DIPG (H = 180) and not significantly different from extrapontine GBM (adult GBM H = 140, pediatric GBM H = 170, p > 0.05, Kruskal-Wallis test of multiple comparisons). A significant difference was noted between tumor groups and normal brain (H = 200, p < 0.05, Kruskal-Wallis test of multiple comparisons) explaining the Kruskal-Wallis ANOVA p < 0.0001 (Figure 3).

DIPGs have increased 5hmC and decreased 5mC compared with extrapontine GBM

Because alterations in DNA methylation levels represent an important feature of DIPG, we next studied 5mC and 5hmC levels in a total of 94 scorable tumors from 3 tissue microarrays (Figure 1). 5hmC immunoreactivity was significantly higher for DIPG (H = 190) compared to pediatric GBM (H = 110) (p < 0.0001) and adult GBM (H = 100) (p < 0.0001). No significant difference in H-scores was noted between DIPG tissue and normal brain (H = 200) (p = 0.23). There was no significant difference in 5hmC H-score between adult and pediatric GBM (p = 0.18) tissue (Figure 4). Conversely, 5mC was significantly lower in DIPGs (H = 120) compared to pediatric GBM (H = 170) (p < 0.001), adult GBM (H = 170) (p < 0.01), and normal brain (H = 190) (p < 0.01). No significant difference in H-scores was noted between adult and pediatric GBM (p = 0.61) tissue for 5mC (Figure 5).

DIPGs have increased mRNA expression of TET1 and TET3 relative to normal brain

Comparative analysis of DIPG tumor tissue and non-neoplastic brain sections showed increased TET1 and TET3 mRNA expression in tumor relative to patient matched brain (as expressed in fold change). TET2 expression was not increased relative to normal brain (Figure 6).

Discussion

We determined by IHC relative levels of H3K27me3, 5mC, and 5hmC in DIPG, pediatric GBM, and adult GBM tissue. Our data are consistent with previously reported low levels of H3K27me3 due to the H3K27M mutation in the majority of DIPG. H3K9me3 in DIPG was not significantly different from H3K9me3 immunoreactivity in extrapontine GBM and unlikely to be playing a role in the tumorigenicity of DIPG. Surprisingly, 5hmC immunoreactivity was significantly elevated in DIPG, in contrast to 5mC, suggesting that imbalance between 5hmC and 5mC plays a global role in the biology of DIPG. In general, high levels of 5hmC have been shown to be a feature of terminally differentiated cells[20]. In numerous solid tumors including carcinoma of the breast, prostate, colon, melanoma, and gliomas, 5hmC levels appear to be greatly reduced in neoplastic cells[20]. Our group has previously shown that in adult GBM and anaplastic astrocytoma high levels of 5hmC are associated with a less aggressive phenotype[21]. However, the global DNA hypomethylation that occurs in DIPG in conjunction with relatively increased 5hmC may represent signs of a novel global epigenetic dysregulation state distinct from that of adult high grade glioma. The elevation of 5hmC in DIPG must be considered in the context of global loss of 5mC in these tumors. The striking imbalance between 5hmC and 5mC in DIPG may be another sign of the marked epigenetic dysregulation which underlies these aggressive tumors.
A limitation of our study is the relative lack of clinical data available for the patient samples. We do not know which, if any, chemotherapeutic agents patients have been treated with and whether they have received radiation. Our findings may be affected by prior treatment but the internal consistency of our findings within tumor groups supports their significance. The underlying cause for the shift in balance towards increased 5hmC expression relative to 5mC in DIPG compared to normal brain and extrapontine GBM is unclear. However, one possibility is altered activity of TETs. Preliminary evaluation of TET mRNA expression shows elevation of TET1 and TET3 in DIPG compared to control brain. In other cancer types, TET is mutated and inactivated, leading to globally decreased 5hmC and increased DNA methylation[17]. In contrast, a possible result of the loss of global DNA methylation could be deposition of the 5hmC intermediary and epigenetic imbalance.
During differentiation, global levels of 5hmC and H3K27me3 are tightly co-regulated and in most solid tumors both marks are concordantly reduced[24]. The data presented here suggests that DIPGs have a unique epigenetic fingerprint featuring high 5hmC and loss of H3K27me3. This epigenetic constellation may contribute to DIPG tumor initiation and maintenance. The alteration in H3.3 which inactivates EZH2 and global DNA hypomethylation suggests that two regulatory pathways that communicate to maintain normal brain development are altered in this tumor type, leading to an unopposed active transcriptional state (Figure 7). Elevated 5hmC and depressed H3K27me3 in normal brain development are associated with neural differentiation and affect neural migration through inhibition of EZH2 (part of the PRC2 complex)[20, 18]. We hypothesize that migration is aberrant in DIPG since EZH2 is no longer able to function at Lysine 27 in Histone 3.3. Elevated 5hmC is found in differentiating cells and usually in conjunction with H3K27me3 loss. There is a possibility that the Histone 3 and cytosine methylation pathways cross-talk and co-regulate each other. However, in DIPG, aberrant TET enzyme expression may result in loss of 5mC and increased 5hmC. This may lead to aberrant PRC2 sequestration from primary targets, a transcriptionally active state, and tumorigenicity.
Histone and cytosine methylation dysregulation is unique to DIPG. When neither pathway is able to function in its regulatory role in neural development, DIPG develops. This epigenetic imbalance may underlie DIPG tumor formation and resistance to treatment. Our work supports recent findings indicating that H3K27me3 loss, through the dominant negative H3K27M mutation, is associated with DNA hypomethylation and an increase in transcriptionally upregulated genes. This suggests that there is crosstalk between histone methylation pathways and DNA methylation pathways leading to changes in transcriptional activity[10].

Conclusions

The H3.3 Lysine 27 is a critical site for the deposition of an inhibitory epigenetic mark and is mutated in the majority of DIPG, suggesting that targeting epigenetics could be one therapeutic approach for this highly aggressive pediatric tumor. Understanding the epigenetic landscape of DIPG opens up the opportunity for epigenetic modifiers, which could potentially shift the active genome of this deadly tumor into a silent and regulated state. The role of epigenetic modifiers such as methylation inhibitors and histone deacetylase inhibitors[7] in treating DIPG needs to be reevaluated as the epigenetic mechanisms of DIPG are better understood.

Acknowledgements

This work was supported in part by the T32 CA60441 NIH/NCI Grant: Lab Research Training in Pediatric Oncology-Hematology (SA), the Children’s Cancer Foundation Inc. (CGE, FJR), Childhood Brain Tumor Foundation (FJR), St. Baldrick’s Foundation (EHR), and a CureSearch Brain Tumor Young Investigator Grant (EHR). The authors also thank Jessica Hicks of the Immunohistochemical Core for technical assistance.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made.
The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
The Creative Commons Public Domain Dedication waiver (https://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

SA participated in design and coordination of study, scored and collected data, performed the statistical analysis, and drafted the manuscript. ER participated in design and coordination of the study and manuscript writing. Also provided materials necessary for the study. MH optimized and performed immunohistochemistry and participated in the design and coordination of the study. AV optimized and performed immunohistochemistry. KW participated in design and coordination of the study. MQ participated in design and coordination of the study. LB participated in design and coordination of the study. JN participated in design and coordination of the study, collected data, and performed statistical analysis. CE conceived of the study and participated in design and coordination of the study. FR conceived of the study and participated in design and coordination of the study. All authors critically read and approved the final manuscript.
Anhänge

Electronic supplementary material

Additional file 2: Figure S1: HEK293 cells are positive controls for 5mC IHC and negative controls for 5hmC. TET overexpressed HEK293 cells function as positive controls for 5hmC IHC. Figure 1a is a representative image of IHC staining of HEK293 cells for 5mC (positive control). Figure 1b shows representative images from three different rounds of staining for 5hmC with HEK293 cells overexpressed for TET2 stained for 5hmC on the right (positive control) and control transfected HEK293 cells not showing any immunoreactivity for 5hmC on the left (negative control). (PPTX 464 KB)
Literatur
1.
Zurück zum Zitat Ballester L, Wang Z, Shandilya S, Miettinen M, Burger P, Eberhart C, Rodriguez F, Raabe E, Nazarian J, Warren K, Quezado M: Morphologic characteristics and immunohistochemical profile of diffuse intrinsic pontine gliomas. Am J Surg Pathol 2013, 37(9):1357–1364. doi:10.1097/PAS.0b013e318294e817 10.1097/PAS.0b013e318294e817CrossRefPubMedPubMedCentral Ballester L, Wang Z, Shandilya S, Miettinen M, Burger P, Eberhart C, Rodriguez F, Raabe E, Nazarian J, Warren K, Quezado M: Morphologic characteristics and immunohistochemical profile of diffuse intrinsic pontine gliomas. Am J Surg Pathol 2013, 37(9):1357–1364. doi:10.1097/PAS.0b013e318294e817 10.1097/PAS.0b013e318294e817CrossRefPubMedPubMedCentral
2.
Zurück zum Zitat Paugh BS, Broniscer A, Qu C, Miller CP, Zhang J, Tatevossian RG, Olson JM, Geyer JR, Chi SN, Da Silva NS, Onar-Thomas A, Baker JN, Gajjar A, Ellison DW, Baker SJ: Genome-wide analyses identify recurrent amplifications of receptor tyrosine kinases and cell-cycle regulatory genes in diffuse intrinsic pontine glioma. J Clin Oncol 2011, 29(30):3999–4006. doi:10.1200/JCO.2011.35.5677 10.1200/JCO.2011.35.5677CrossRefPubMedPubMedCentral Paugh BS, Broniscer A, Qu C, Miller CP, Zhang J, Tatevossian RG, Olson JM, Geyer JR, Chi SN, Da Silva NS, Onar-Thomas A, Baker JN, Gajjar A, Ellison DW, Baker SJ: Genome-wide analyses identify recurrent amplifications of receptor tyrosine kinases and cell-cycle regulatory genes in diffuse intrinsic pontine glioma. J Clin Oncol 2011, 29(30):3999–4006. doi:10.1200/JCO.2011.35.5677 10.1200/JCO.2011.35.5677CrossRefPubMedPubMedCentral
3.
Zurück zum Zitat Paugh BS, Zhu X, Qu C, Endersby R, Diaz AK, Zhang J, Bax DA, Carvalho D, Reis RM, Onar-Thomas A, Broniscer A, Wetmore C, Zhang J, Jones C, Ellison DW, Baker SJ: Novel oncogenic PDGFRA mutations in pediatric high-grade gliomas. Cancer Res 2013, 73(20):6219–6229. doi:10.1158/0008–5472. CAN-13–1491 10.1158/0008-5472.CAN-13-1491CrossRefPubMedPubMedCentral Paugh BS, Zhu X, Qu C, Endersby R, Diaz AK, Zhang J, Bax DA, Carvalho D, Reis RM, Onar-Thomas A, Broniscer A, Wetmore C, Zhang J, Jones C, Ellison DW, Baker SJ: Novel oncogenic PDGFRA mutations in pediatric high-grade gliomas. Cancer Res 2013, 73(20):6219–6229. doi:10.1158/0008–5472. CAN-13–1491 10.1158/0008-5472.CAN-13-1491CrossRefPubMedPubMedCentral
4.
Zurück zum Zitat Saratsis AM, Kambhampati M, Snyder K, Yadavilli S, Devaney JM, Harmon B, Hall J, Raabe EH, An P, Weingart M, Rood BR, Magge SN, Macdonald TJ, Packer RJ, Nazarian J: Comparative multidimensional molecular analyses of pediatric diffuse intrinsic pontine glioma reveals distinct molecular subtypes. Acta Neuropathol 2013. doi:10.1007/s00401–013–1218–2 Saratsis AM, Kambhampati M, Snyder K, Yadavilli S, Devaney JM, Harmon B, Hall J, Raabe EH, An P, Weingart M, Rood BR, Magge SN, Macdonald TJ, Packer RJ, Nazarian J: Comparative multidimensional molecular analyses of pediatric diffuse intrinsic pontine glioma reveals distinct molecular subtypes. Acta Neuropathol 2013. doi:10.1007/s00401–013–1218–2
5.
Zurück zum Zitat Buczkowicz P, Hoeman C, Rakopoulos P, Pajovic S, Letourneau L, Dzamba M, Morrison A, Lewis P, Bouffet E, Bartels U, Zuccaro J, Agnihotri S, Ryall S, Barszczyk M, Chornenkyy Y, Bourgey M, Bourque G, Montpetit A, Cordero F, Castelo-Branco P, Mangerel J, Tabori U, Ho KC, Huang A, Taylor KR, Mackay A, Bendel AE, Nazarian J, Fangusaro JR, Karajannis MA, et al.: Genomic analysis of diffuse intrinsic pontine gliomas identifies three molecular subgroups and recurrent activating ACVR1 mutations. Nat Genet 2014. doi:10.1038/ng.2936 Buczkowicz P, Hoeman C, Rakopoulos P, Pajovic S, Letourneau L, Dzamba M, Morrison A, Lewis P, Bouffet E, Bartels U, Zuccaro J, Agnihotri S, Ryall S, Barszczyk M, Chornenkyy Y, Bourgey M, Bourque G, Montpetit A, Cordero F, Castelo-Branco P, Mangerel J, Tabori U, Ho KC, Huang A, Taylor KR, Mackay A, Bendel AE, Nazarian J, Fangusaro JR, Karajannis MA, et al.: Genomic analysis of diffuse intrinsic pontine gliomas identifies three molecular subgroups and recurrent activating ACVR1 mutations. Nat Genet 2014. doi:10.1038/ng.2936
6.
Zurück zum Zitat Taylor KR, Mackay A, Truffaux N, Butterfield YS, Morozova O, Philippe C, Castel D, Grasso CS, Vinci M, Carvalho D, Carcaboso AM, De Torres C, Cruz O, Mora J, Entz-Werle N, Ingram WJ, Monje M, Hargrave D, Bullock AN, Puget S, Yip S, Jones C, Grill J: Recurrent activating ACVR1 mutations in diffuse intrinsic pontine glioma. Nat Genet 2014. doi:10.1038/ng.2925 Taylor KR, Mackay A, Truffaux N, Butterfield YS, Morozova O, Philippe C, Castel D, Grasso CS, Vinci M, Carvalho D, Carcaboso AM, De Torres C, Cruz O, Mora J, Entz-Werle N, Ingram WJ, Monje M, Hargrave D, Bullock AN, Puget S, Yip S, Jones C, Grill J: Recurrent activating ACVR1 mutations in diffuse intrinsic pontine glioma. Nat Genet 2014. doi:10.1038/ng.2925
7.
Zurück zum Zitat Dewoskin V, Million R: The epigenetics pipeline. Nat Rev Drug Discov 2013, 12(9):661–662. doi:10.1038/nrd4091 10.1038/nrd4091CrossRefPubMed Dewoskin V, Million R: The epigenetics pipeline. Nat Rev Drug Discov 2013, 12(9):661–662. doi:10.1038/nrd4091 10.1038/nrd4091CrossRefPubMed
8.
Zurück zum Zitat Schwartzentruber J, Korshunov A, Liu X-Y, Jones D, Pfaff E, Jacob K, Sturm D, Fontebasso A, Quang D-AK, Tönjes M, Hovestadt V, Albrecht S, Kool M, Nantel A, Konermann C, Lindroth A, Jäger N, Rausch T, Ryzhova M, Korbel J, Hielscher T, Hauser P, Garami M, Klekner A, Bognar L, Ebinger M, Schuhmann M, Scheurlen W, Pekrun A, Frühwald M, et al.: Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 2012, 482(7384):226–231. doi:10.1038/nature10833 10.1038/nature10833CrossRefPubMed Schwartzentruber J, Korshunov A, Liu X-Y, Jones D, Pfaff E, Jacob K, Sturm D, Fontebasso A, Quang D-AK, Tönjes M, Hovestadt V, Albrecht S, Kool M, Nantel A, Konermann C, Lindroth A, Jäger N, Rausch T, Ryzhova M, Korbel J, Hielscher T, Hauser P, Garami M, Klekner A, Bognar L, Ebinger M, Schuhmann M, Scheurlen W, Pekrun A, Frühwald M, et al.: Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 2012, 482(7384):226–231. doi:10.1038/nature10833 10.1038/nature10833CrossRefPubMed
9.
Zurück zum Zitat Sturm D, Witt H, Hovestadt V, Khuong-Quang D-A, Jones D, Konermann C, Pfaff E, Tönjes M, Sill M, Bender S, Kool M, Zapatka M, Becker N, Zucknick M, Hielscher T, Liu X-Y, Fontebasso A, Ryzhova M, Albrecht S, Jacob K, Wolter M, Ebinger M, Schuhmann M, van Meter T, Frühwald M, Hauch H, Pekrun A, Radlwimmer B, Niehues T, von Komorowski G, et al.: Hotspot mutations in H3F3A and IDH1 define distinct epigenetic and biological subgroups of glioblastoma. Cancer Cell 2012, 22(4):425–437. doi:10.1016/j.ccr.2012.08.024 10.1016/j.ccr.2012.08.024CrossRefPubMed Sturm D, Witt H, Hovestadt V, Khuong-Quang D-A, Jones D, Konermann C, Pfaff E, Tönjes M, Sill M, Bender S, Kool M, Zapatka M, Becker N, Zucknick M, Hielscher T, Liu X-Y, Fontebasso A, Ryzhova M, Albrecht S, Jacob K, Wolter M, Ebinger M, Schuhmann M, van Meter T, Frühwald M, Hauch H, Pekrun A, Radlwimmer B, Niehues T, von Komorowski G, et al.: Hotspot mutations in H3F3A and IDH1 define distinct epigenetic and biological subgroups of glioblastoma. Cancer Cell 2012, 22(4):425–437. doi:10.1016/j.ccr.2012.08.024 10.1016/j.ccr.2012.08.024CrossRefPubMed
10.
Zurück zum Zitat Bender S, Tang Y, Lindroth A, Hovestadt V, Jones D, Kool M, Zapatka M, Northcott P, Sturm D, Wang W, Radlwimmer B, Højfeldt J, Truffaux N, Castel D, Schubert S, Ryzhova M, Seker-Cin H, Gronych J, Johann P, Stark S, Meyer J, Milde T, Schuhmann M, Ebinger M, Monoranu C-M, Ponnuswami A, Chen S, Jones C, Witt O, Collins V, et al.: Reduced H3K27me3 and DNA Hypomethylation Are Major Drivers of Gene Expression in K27M Mutant Pediatric High-Grade Gliomas. Cancer Cell 2013, 24(5):660–672. doi:10.1016/j.ccr.2013.10.006 10.1016/j.ccr.2013.10.006CrossRefPubMed Bender S, Tang Y, Lindroth A, Hovestadt V, Jones D, Kool M, Zapatka M, Northcott P, Sturm D, Wang W, Radlwimmer B, Højfeldt J, Truffaux N, Castel D, Schubert S, Ryzhova M, Seker-Cin H, Gronych J, Johann P, Stark S, Meyer J, Milde T, Schuhmann M, Ebinger M, Monoranu C-M, Ponnuswami A, Chen S, Jones C, Witt O, Collins V, et al.: Reduced H3K27me3 and DNA Hypomethylation Are Major Drivers of Gene Expression in K27M Mutant Pediatric High-Grade Gliomas. Cancer Cell 2013, 24(5):660–672. doi:10.1016/j.ccr.2013.10.006 10.1016/j.ccr.2013.10.006CrossRefPubMed
11.
Zurück zum Zitat Wu G, Broniscer A, McEachron TA, Lu C, Paugh BS, Becksfort J, Qu C, Ding L, Huether R, Parker M, Zhang J, Gajjar A, Dyer MA, Mullighan CG, Gilbertson RJ, Mardis ER, Wilson RK, Downing JR, Ellison DW, Zhang J, Baker SJ, St. Jude Children’s Research Hospital-Washington University Pediatric Cancer Genome P: Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat Genet 2012, 44(3):251–253. doi:10.1038/ng.1102 10.1038/ng.1102CrossRefPubMedPubMedCentral Wu G, Broniscer A, McEachron TA, Lu C, Paugh BS, Becksfort J, Qu C, Ding L, Huether R, Parker M, Zhang J, Gajjar A, Dyer MA, Mullighan CG, Gilbertson RJ, Mardis ER, Wilson RK, Downing JR, Ellison DW, Zhang J, Baker SJ, St. Jude Children’s Research Hospital-Washington University Pediatric Cancer Genome P: Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat Genet 2012, 44(3):251–253. doi:10.1038/ng.1102 10.1038/ng.1102CrossRefPubMedPubMedCentral
12.
Zurück zum Zitat Khuong-Quang DA, Buczkowicz P, Rakopoulos P, Liu XY, Fontebasso AM, Bouffet E, Bartels U, Albrecht S, Schwartzentruber J, Letourneau L, Bourgey M, Bourque G, Montpetit A, Bourret G, Lepage P, Fleming A, Lichter P, Kool M, von Deimling A, Sturm D, Korshunov A, Faury D, Jones DT, Majewski J, Pfister SM, Jabado N, Hawkins C: K27M mutation in histone H3.3 defines clinically and biologically distinct subgroups of pediatric diffuse intrinsic pontine gliomas. Acta Neuropathol 2012, 124(3):439–447. doi:10.1007/s00401–012–0998–0 10.1007/s00401-012-0998-0CrossRefPubMedPubMedCentral Khuong-Quang DA, Buczkowicz P, Rakopoulos P, Liu XY, Fontebasso AM, Bouffet E, Bartels U, Albrecht S, Schwartzentruber J, Letourneau L, Bourgey M, Bourque G, Montpetit A, Bourret G, Lepage P, Fleming A, Lichter P, Kool M, von Deimling A, Sturm D, Korshunov A, Faury D, Jones DT, Majewski J, Pfister SM, Jabado N, Hawkins C: K27M mutation in histone H3.3 defines clinically and biologically distinct subgroups of pediatric diffuse intrinsic pontine gliomas. Acta Neuropathol 2012, 124(3):439–447. doi:10.1007/s00401–012–0998–0 10.1007/s00401-012-0998-0CrossRefPubMedPubMedCentral
13.
Zurück zum Zitat Lewis P, Müller M, Koletsky M, Cordero F, Lin S, Banaszynski L, Garcia B, Muir T, Becher O, Allis C: Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science (New York, NY) 2013, 340(6134):857–861. doi:10.1126/science.1232245 10.1126/science.1232245CrossRef Lewis P, Müller M, Koletsky M, Cordero F, Lin S, Banaszynski L, Garcia B, Muir T, Becher O, Allis C: Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science (New York, NY) 2013, 340(6134):857–861. doi:10.1126/science.1232245 10.1126/science.1232245CrossRef
14.
Zurück zum Zitat Sparmann A, van Lohuizen M: Polycomb silencers control cell fate, development and cancer. Nat Rev Cancer 2006, 6(11):846–856. doi:10.1038/nrc1991 10.1038/nrc1991CrossRefPubMed Sparmann A, van Lohuizen M: Polycomb silencers control cell fate, development and cancer. Nat Rev Cancer 2006, 6(11):846–856. doi:10.1038/nrc1991 10.1038/nrc1991CrossRefPubMed
15.
Zurück zum Zitat Reddington J, Perricone S, Nestor C, Reichmann J, Youngson N, Suzuki M, Reinhardt D, Dunican D, Prendergast J, Mjoseng H, Ramsahoye B, Whitelaw E, Greally J, Adams I, Bickmore W, Meehan R: Redistribution of H3K27me3 upon DNA hypomethylation results in de-repression of Polycomb target genes. Genome Biol 2013., 14(3): doi:10.1186/gb-2013–14–3-r25 Reddington J, Perricone S, Nestor C, Reichmann J, Youngson N, Suzuki M, Reinhardt D, Dunican D, Prendergast J, Mjoseng H, Ramsahoye B, Whitelaw E, Greally J, Adams I, Bickmore W, Meehan R: Redistribution of H3K27me3 upon DNA hypomethylation results in de-repression of Polycomb target genes. Genome Biol 2013., 14(3): doi:10.1186/gb-2013–14–3-r25
16.
Zurück zum Zitat Pfeifer G, Kadam S, Jin S-G: 5-hydroxymethylcytosine and its potential roles in development and cancer. Epigenetics Chromatin 2013, 6(1):10. doi:10.1186/1756–8935–6-10 10.1186/1756-8935-6-10CrossRefPubMedPubMedCentral Pfeifer G, Kadam S, Jin S-G: 5-hydroxymethylcytosine and its potential roles in development and cancer. Epigenetics Chromatin 2013, 6(1):10. doi:10.1186/1756–8935–6-10 10.1186/1756-8935-6-10CrossRefPubMedPubMedCentral
17.
Zurück zum Zitat Sun W, Guan M, Li X: 5-hydroxymethylcytosine-mediated DNA demethylation in stem cells and development. Stem Cells Dev 2014. doi:10.1089/scd.2013.0428 Sun W, Guan M, Li X: 5-hydroxymethylcytosine-mediated DNA demethylation in stem cells and development. Stem Cells Dev 2014. doi:10.1089/scd.2013.0428
18.
Zurück zum Zitat Hahn M, Qiu R, Wu X, Li A, Zhang H, Wang J, Jui J, Jin S-G, Jiang Y, Pfeifer G, Lu Q: Dynamics of 5-hydroxymethylcytosine and chromatin marks in Mammalian neurogenesis. Cell Rep 2013, 3(2):291–300. doi:10.1016/j.celrep.2013.01.011 10.1016/j.celrep.2013.01.011CrossRefPubMedPubMedCentral Hahn M, Qiu R, Wu X, Li A, Zhang H, Wang J, Jui J, Jin S-G, Jiang Y, Pfeifer G, Lu Q: Dynamics of 5-hydroxymethylcytosine and chromatin marks in Mammalian neurogenesis. Cell Rep 2013, 3(2):291–300. doi:10.1016/j.celrep.2013.01.011 10.1016/j.celrep.2013.01.011CrossRefPubMedPubMedCentral
19.
Zurück zum Zitat Ito S, Shen L, Dai Q, Wu SC, Collins LB, Swenberg JA, He C, Zhang Y: Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 2011, 333(6047):1300–1303. doi:10.1126/science.1210597 10.1126/science.1210597CrossRefPubMedPubMedCentral Ito S, Shen L, Dai Q, Wu SC, Collins LB, Swenberg JA, He C, Zhang Y: Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 2011, 333(6047):1300–1303. doi:10.1126/science.1210597 10.1126/science.1210597CrossRefPubMedPubMedCentral
20.
Zurück zum Zitat Haffner MC, Chaux A, Meeker AK, Esopi DM, Gerber J, Pellakuru LG, Toubaji A, Argani P, Iacobuzio-Donahue C, Nelson WG, Netto GJ, De Marzo AM, Yegnasubramanian S: Global 5-hydroxymethylcytosine content is significantly reduced in tissue stem/progenitor cell compartments and in human cancers. Oncotarget 2011, 2(8):627–637.CrossRefPubMedPubMedCentral Haffner MC, Chaux A, Meeker AK, Esopi DM, Gerber J, Pellakuru LG, Toubaji A, Argani P, Iacobuzio-Donahue C, Nelson WG, Netto GJ, De Marzo AM, Yegnasubramanian S: Global 5-hydroxymethylcytosine content is significantly reduced in tissue stem/progenitor cell compartments and in human cancers. Oncotarget 2011, 2(8):627–637.CrossRefPubMedPubMedCentral
21.
Zurück zum Zitat Orr B, Haffner M, Nelson W, Yegnasubramanian S, Eberhart C: Decreased 5-hydroxymethylcytosine is associated with neural progenitor phenotype in normal brain and shorter survival in malignant glioma. PLoS One 2012., 7(7): doi:10.1371/journal.pone.0041036 Orr B, Haffner M, Nelson W, Yegnasubramanian S, Eberhart C: Decreased 5-hydroxymethylcytosine is associated with neural progenitor phenotype in normal brain and shorter survival in malignant glioma. PLoS One 2012., 7(7): doi:10.1371/journal.pone.0041036
22.
Zurück zum Zitat Venneti S, Garimella MT, Sullivan LM, Martinez D, Huse JT, Heguy A, Santi M, Thompson CB, Judkins AR: Evaluation of histone 3 lysine 27 trimethylation (H3K27me3) and enhancer of Zest 2 (EZH2) in pediatric glial and glioneuronal tumors shows decreased H3K27me3 in H3F3A K27M mutant glioblastomas. Brain Pathol 2013, 23(5):558–564. doi:10.1111/bpa.12042 10.1111/bpa.12042CrossRefPubMed Venneti S, Garimella MT, Sullivan LM, Martinez D, Huse JT, Heguy A, Santi M, Thompson CB, Judkins AR: Evaluation of histone 3 lysine 27 trimethylation (H3K27me3) and enhancer of Zest 2 (EZH2) in pediatric glial and glioneuronal tumors shows decreased H3K27me3 in H3F3A K27M mutant glioblastomas. Brain Pathol 2013, 23(5):558–564. doi:10.1111/bpa.12042 10.1111/bpa.12042CrossRefPubMed
23.
Zurück zum Zitat Odia Y, Orr BA, Bell WR, Eberhart CG, Rodriguez FJ: cMYC expression in infiltrating gliomas: associations with IDH1 mutations, clinicopathologic features and outcome. J Neurooncol 2013, 115(2):249–259. doi:10.1007/s11060–013–1221–4 10.1007/s11060-013-1221-4CrossRefPubMed Odia Y, Orr BA, Bell WR, Eberhart CG, Rodriguez FJ: cMYC expression in infiltrating gliomas: associations with IDH1 mutations, clinicopathologic features and outcome. J Neurooncol 2013, 115(2):249–259. doi:10.1007/s11060–013–1221–4 10.1007/s11060-013-1221-4CrossRefPubMed
24.
Zurück zum Zitat Haffner M, Pellakuru L, Ghosh S, Lotan T, Nelson W, De Marzo A, Yegnasubramanian S: Tight correlation of 5-hydroxymethylcytosine and Polycomb marks in health and disease. Cell Cycle 2013, 12(12):1835–1841. doi:10.4161/cc.25010 10.4161/cc.25010CrossRefPubMedPubMedCentral Haffner M, Pellakuru L, Ghosh S, Lotan T, Nelson W, De Marzo A, Yegnasubramanian S: Tight correlation of 5-hydroxymethylcytosine and Polycomb marks in health and disease. Cell Cycle 2013, 12(12):1835–1841. doi:10.4161/cc.25010 10.4161/cc.25010CrossRefPubMedPubMedCentral
Metadaten
Titel
Increased 5-hydroxymethylcytosine and decreased 5-methylcytosine are indicators of global epigenetic dysregulation in diffuse intrinsic pontine glioma
verfasst von
Sama Ahsan
Eric H Raabe
Michael C Haffner
Ajay Vaghasia
Katherine E Warren
Martha Quezado
Leomar Y Ballester
Javad Nazarian
Charles G Eberhart
Fausto J Rodriguez
Publikationsdatum
01.12.2014
Verlag
BioMed Central
Erschienen in
Acta Neuropathologica Communications / Ausgabe 1/2014
Elektronische ISSN: 2051-5960
DOI
https://doi.org/10.1186/2051-5960-2-59

Weitere Artikel der Ausgabe 1/2014

Acta Neuropathologica Communications 1/2014 Zur Ausgabe

Leitlinien kompakt für die Neurologie

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Sozialer Aufstieg verringert Demenzgefahr

24.05.2024 Demenz Nachrichten

Ein hohes soziales Niveau ist mit die beste Versicherung gegen eine Demenz. Noch geringer ist das Demenzrisiko für Menschen, die sozial aufsteigen: Sie gewinnen fast zwei demenzfreie Lebensjahre. Umgekehrt steigt die Demenzgefahr beim sozialen Abstieg.

Hirnblutung unter DOAK und VKA ähnlich bedrohlich

17.05.2024 Direkte orale Antikoagulanzien Nachrichten

Kommt es zu einer nichttraumatischen Hirnblutung, spielt es keine große Rolle, ob die Betroffenen zuvor direkt wirksame orale Antikoagulanzien oder Marcumar bekommen haben: Die Prognose ist ähnlich schlecht.

Was nützt die Kraniektomie bei schwerer tiefer Hirnblutung?

17.05.2024 Hirnblutung Nachrichten

Eine Studie zum Nutzen der druckentlastenden Kraniektomie nach schwerer tiefer supratentorieller Hirnblutung deutet einen Nutzen der Operation an. Für überlebende Patienten ist das dennoch nur eine bedingt gute Nachricht.

Thrombektomie auch bei großen Infarkten von Vorteil

16.05.2024 Ischämischer Schlaganfall Nachrichten

Auch ein sehr ausgedehnter ischämischer Schlaganfall scheint an sich kein Grund zu sein, von einer mechanischen Thrombektomie abzusehen. Dafür spricht die LASTE-Studie, an der Patienten und Patientinnen mit einem ASPECTS von maximal 5 beteiligt waren.

Update Neurologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.