Skip to main content
Erschienen in: Journal of Neuroinflammation 1/2020

Open Access 01.12.2020 | Review

Aberrant DNA methylation profile exacerbates inflammation and neurodegeneration in multiple sclerosis patients

verfasst von: Naiara Celarain, Jordi Tomas-Roig

Erschienen in: Journal of Neuroinflammation | Ausgabe 1/2020

Abstract

Multiple sclerosis (MS) is an autoimmune and demyelinating disease of the central nervous system characterised by incoordination, sensory loss, weakness, changes in bladder capacity and bowel function, fatigue and cognitive impairment, creating a significant socioeconomic burden. The pathogenesis of MS involves both genetic susceptibility and exposure to distinct environmental risk factors. The gene x environment interaction is regulated by epigenetic mechanisms. Epigenetics refers to a complex system that modifies gene expression without altering the DNA sequence. The most studied epigenetic mechanism is DNA methylation. This epigenetic mark participates in distinct MS pathophysiological processes, including blood–brain barrier breakdown, inflammatory response, demyelination, remyelination failure and neurodegeneration. In this study, we also accurately summarised a list of environmental factors involved in the MS pathogenesis and its clinical course. A literature search was conducted using MEDLINE through PubMED and Scopus. In conclusion, an exhaustive study of DNA methylation might contribute towards new pharmacological interventions in MS by use of epigenetic drugs.
Hinweise

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
5caC
5-Carboxylcytosine
5fC
5-Formylcytosine
5hmC
5-Hydroxymethylcytosine
5hmU
5-Hydroxymethyluracil
5mC
5-Methylcytosine
AHRR
Aryl-hydrocarbon receptor repressor
APC
Antigen presenting cells
ASC
Apoptosis-associated speck-like protein containing a C-terminal caspase recruitment domain
BBB
Blood–brain barrier
BDNF
Brain-derived neurotrophic factor
BER
Base excision repair
BS
Bisulphite
CDH1
E-cadherin
cfDNA
Circulating-free DNA
Cit
Citrullination
CNS
Central nervous system
CPZ
Cuprizone
CSF
Cerebrospinal fluid
CTR
Control
CYP1A1
Cytochrome P450 family 1 subfamily A member 1
DC
Dendritic cell
Dexras-1
Ras-related protein-1
DNAme
DNA methylation
DNMT
DNA methyltransferases
EAE
Experimental autoimmune/allergic encephalomyelitis
EBV
Epstein Barr virus
FGF
Fibroblast growth factor
FOXP3
Forkhead box P3
H2S
Hydrogen sulphide
HCB
Hexachlorobenzene
HDACs
Histone deacetylase inhibitors
HLA
Human leukocyte antigen
ICAM
Intercellular adhesion molecules
IFN-γ
Interferon-gamma
ILK
Integrin-linked kinase
LCFA
Long-chain fatty acids
LGMN
Legumain
LINE-1
Long interspersed element-1
MBD
Methyl CpG-binding protein
MBP
Myelin basic protein
MeDIP
Methylated-DNA immunoprecipitation
MHC
Major histocompatibility complex
MICB
MHC class I polypeptide-related sequence B
MOG
Myelin oligodendrocyte glycoprotein
MS
Multiple sclerosis
NaCl
Salt
NAWM
Normal appearing white matter
NFAS
Neurofascin
NF-kβ
Nuclear factor k-light-chain-enhancer of activated B cells
NH3
Ammonia
NH4
Ammonium
NKRs
Natural killer receptors
nNOS
Neuronal nitric oxide synthase
NR3C1
Nuclear subfamily 3 group C member 1 glucocorticoid receptor
OPCs
Oligodendrocyte progenitor cells
PAD2
Peptidil arginine deiminase 2
PBMCs
Peripheral blood mononuclear cells
PDGF
Platelet-derived growth factor
PLP
Proteolipid protein
PPMS
Primary progressive multiple sclerosis
ROS
Reactive oxygen species
RRBS
Reduced representation bisulphite sequencing
RRMS(r)
RRMS in remission
RRMS
Relapsing–remitting multiple sclerosis
RRMS(e)
RRMS in exacerbation
SAM
Methyl-donor S-adenosylmethionine
SCFA
Short-chain fatty acids
SHP-1
Src homology region 2 domain-containing phosphatase-1
SOCS1
Suppressor of cytokine signalling 1
SPMS
Secondary progressive multiple sclerosis
TAB-seq
TET-assisted bisulphate sequencing
TCRs
T cell receptors
TDG
Thymine DNA glycosylase
TET
Ten–eleven translocation
Th
T helper
TMEV
Theiler’s murine encephalomyelitis virus
VDR
Vitamin D receptor
VEGF-A
Vascular endothelial growth factor A
Vit D
Vitamin D

Background

Multiple sclerosis (MS) is an autoimmune, inflammatory, demyelinating and neurodegenerative disease of the central nervous system (CNS) [1]. As a result of myelin sheath destruction, the electric impulse between neurons is inefficient, and thus the initial symptoms appear [2]. Although symptoms differ from each MS patient, the most common ones include incoordination, sensory loss, weakness, changes in bladder capacity and bowel function, fatigue, and cognitive impairment [3]. Therefore, MS has serious negative effects on the health-, social, and work-related issues of patients and their families, creating a significant socioeconomic burden [4]. The aetiology of MS is still unknown and requires a close interaction between genetic susceptibility and exposure to environmental agents. Synergistic effect of these risk factors would be responsible for triggering autoimmunity in MS patients. For this reason, the underlying mechanisms involved in the MS pathogenesis can differ among patients. Common mechanisms are observed in the pathophysiology of disease and listed next. Autoreactive CD4+ T cells are activated in the periphery [1] by antigen-presenting cells (APC), that present via the class II major histocompatibility complex (MHC) receptor an amino acid similar to myelin peptides synthesised in the CNS. This interaction activates the differentiation of the CD4+ T naïve cells into CD4+ T helper cells [5]. Upon activation, the Th1 subtype produces interferon gamma (IFN-γ) [6], a cytokine responsible for recruiting CD8+ T cells, B cells and monocytes in the periphery [7]. Then, these proinflammatory cells migrate to the blood–brain barrier (BBB) throughout the bloodstream, where they can adhere to the BBB endothelium [8]. In a healthy brain, immune cells are circulating freely in the meninges orchestrating immune surveillance of the CNS [9]. In MS, the BBB displays an aberrant expression and organisation of the endothelial tight junctions [10] that favours massive lymphocyte trafficking into the brain [11]. Infiltrated CD4+ T cells in the CNS are reactivated upon interaction with the resident APCs [12]. Afterwards, the reactivated CD4+ T cells release a variety of proinflammatory cytokines and chemokines [13], resulting in astrogliosis [14] and microgliosis [15]. This process is exacerbated when infiltrated cytotoxic CD8+ T cells attack oligodendrocytes, causing their destruction and neuronal death [16]. In parallel, plasma B cells produce antibodies against CNS self-antigens, contributing to myelin sheath damage [17]. Plasma B cells in coordination with monocytes increase the local inflammatory response by reactivating the autoreactive CD4+ T cells [18] (Fig. 1). T cell–mediated axonal injury contributes to trophic/metabolic support deficiency from oligodendrocytes as well as a lack of energy by releasing soluble inflammatory molecules [19]. The pathophysiology of MS suggests a complex interaction between the genetic and environmental risk factors [20] regulated by epigenetic mechanisms. Epigenetics can provide a stable heritable base for understanding the underlying mechanisms involved in MS [21].

Genetic, epigenetic and environmental factors

The robust susceptibility loci that confer risk for MS are the human leukocyte antigen (HLA) system, which is located in the short arm of chromosome 6 [22]. However, only 27% of MS heritability can be explained by the genetic variants of the HLA system [23], which supports a prominent contribution of the environment to the MS pathogenesis. Indeed, Epstein Barr virus (EBV) infection, tobacco smoking, vitamin D deficiency, diet style and sun light exposure are critically involved in MS susceptibility [24, 25]. The individual genetic background in combination with the environmental risk factors increase the probability of developing MS. Epigenetic modifications do not alter the sequence of DNA, and they comprise distinct mechanisms such as DNA methylation (DNAme), histone modifications and micro-RNA [26]. Given the scope of this review, we focus our efforts on describing the contribution of DNAme in MS.

DNA methylation

DNAme is the most current epigenetic hallmark in human somatic cells [27]. The methylation of DNA occurs when a methyl group is transferred to the fifth carbon of a cytosine (5mC) through the action of DNA methyltransferases (DNMT). This process occurs mainly in CG dinucleotides, which are commonly found in the regulatory and promoter regions [28]. The addition of methyl groups in the promoter region contributes to gene silencing [28]. In mammals, DNMTs are mainly represented by DNMT1, DNMT3a and DNMT3b. DNMT1 acts during the cell cycle to maintain the DNAme pattern [29] and participates in the DNA repair system [30]. By contrast, DNMT3a and DNMT3b catalyse the de novo addition of a methyl group into a naked cytosine [31]. This fact occurs in cooperation with either the specific transcription factors or the binding transcription factors, which methylate all the CpG sites uncovered [32].
DNAme is a dynamic process that usually requires the removal of the methyl group (demethylation) to cope with environmental stimuli [33]. This process can be achieved in a passive or active manner. Passive DNA demethylation occurs when DNMT1 activity is misregulated and cannot maintain the integrity of the DNAme pattern during the DNA replication, leading to the incorporation of unmethylated cytosine into the genome [34]. Active DNA demethylation is achieved when a sequence of enzymatic reactions of oxidation and/or deamination modifies 5-methylcytosine (5mC) to obtain a naked cytosine. In the oxidation pathway, 5mC is oxidised by the ten–eleven translocation (TET) enzymes to 5-hydroxymethylcytosine (5hmC) [35], which can be further oxidised to 5-formylcytosine (5fC) and 5-carboxylcytosine (5caC) [36]. In the deamination pathway, AID or APOBEC deaminates 5hmC to 5-hydroxymethyluracil (5hmU) or 5mC to thymine [37]. Eventually, all these modified bases (5hmU, Thymine, 5fC, 5caC) can be recognised by thymine DNA glycosylase (TDG) [38] and converted to naked cytosine through the base excision repair pathway (Fig. 2).

Contribution of DNA methylation in patients with MS

Although the precise role of the DNAme in MS remains to be fully elucidated, several studies have reported differentially methylated regions in both the immune cells and brain tissue collected from MS patients (Table 1). Therefore, we summarise the current published research on DNAme as follows: BBB breakdown, inflammation, demyelination, remyelination failure and neurodegeneration.
Table 1
DNA methylation changes in MS
References
Comparison
Sample target
Method
Differentially methylated genes
[39]
MS vs CTR
CD8+ T cells
Illumina 450K array
ERG, FTL, DCAF4, NCAPH2, CDKN1C, ZNF462, CBX7, MIR492, HPS1, SASH1, MYL3, KCNG1, DYDC2, MEGF10, SP5, LMO3, SLC12A7, MORN1, IGF2BP1, PLCB3, ABCC4, CREG2, CDC42BPB, UGT1A10, TMEM125, ARHGAP22, DACH1, OR8B12, TMEM8C, BAI1, EIF2S1, CRTAC1, DHX36, C19orf41, DLGAP2, TNXB, PRDM8, HEATR2, WHSC2, CAMTA1, ALK, KCNQ2, SCTR, RHEB, LOC202181, RRP9, KRT75, DGKE, PLD5, ZC3H14.
[40]
RRMS vs CTR
CD4+ T cells
Illumina 450K array
MICA, MICB, HLA-DRB, MORN1, LCLAT1, PDCD1, MUC4, AHRR, ARSB, PCBD2, TGFBI, PCDHB13, PCDHB15, KIF25, CSGALNACT1, ADARB2, LDHAL6A, CORO1B, USP35, FUT4, ERC1, TCRA, PACS2, IL32, KCTD11, C17orf108, ARHGAP27, NPLOC4, SBNO2, GNG7, C21orf56, RIBC2.
[41]
Myelinated vs demyelinated
MS brains
Hippocampus
Illumina 450K array
MLLT4, PPIF, SCRT2, SNRNP40, ISLR2, MEF2A, PMEPA1, ABCA4, ADAMTS12, AHRR, BEST3, CASP7, CCL4L2, CPXM2, FBXW8, HLA-B, LOC145845, MEIS1, MGMT, MYO7A, NXN, PKP2, PQLC1, PSD3, SCN4B, SDK2, SMYD3, TGFBI, TMEM165, PON1, HDLBP, MKKS, TRIM26, TRPS1, KRTAP27-1, MGP, AJAP1, C1orf106, C2orf62, DSE, EIF2C2, GATA5, HLA-B, IGSF9B, INSC, KIAA1026, KIF25, LOC100292680, NFASC, RASA3, SDK1, SHISA2, SOLH, SORBS2, TAGLN3, TBX5, TM9SF1, TOP1MT, ZSCAN1, AKNA, EBPL, FLJ42709, HERC6, OR52M1, SFRP1, C22orf43, LOC285830, NAPEPLD, NHLH2, PLCH1, SERPINA9, SLFN13, TMEM132B, TTLL3, WDR81.
[42]
SPMS vs CTR
PMBCs
Microarray dataset and RT-PCR
DNMT3A, GADD45A, GADD45B, MBD4, APOBEC3D, APOBEC4, GADD45G, TET1, TDG, APOBEC3C, APOBEC2, MBD2, MBD3, APOBEC3A, DNMT3B, APOBEC1, TET2, TET3.
[43]
RRMS vs PPMS vs CTR
PMBCs
Illumina 450K array
RRMS vs CTR: ASB2, ATP11A, CACNA2D3, CERS5, ESRRG, FRMD4A, GNAS, HOXC4-HOXC6, IFITM5, ILDR1, KCNK15, KLHL35, LEFTY2, PLEKHA2, RPH3AL, WRAP73, ZFYVE28.PPMS vs CTR: ATG16L2, CES1, CSGALNACT2, CYB5D1, LSMD1, FAM110A, GDF7, HKR1, HLA-F, HOXB13, IGSF9B, ILDR1, LDB2, MTPN, LUZP6, NTN1, OPCML, OR2L13, RBM46, TBX1, TCP10L, TMEM44, VIPR2, WRAP73. RRMS vs SPMS: ABCC5, AKAP12, CARS, CBFA2T3, CCDC67, FAM110A, FRMD4A, GIMAP5, HIVEP3, ICAM5, KCNQ1, KLF4, LEFTY2, OLFM3, PTH1R, RASA3, RNF39, RPH3AL, TRAF3, USP35, XKR5.
[44]
Smoker vs non-smoker MS
PMBCs
Bisulphite Illumina Methylation 450k Beadchip
SRM, GNG12, GFI1, ANXA4, NFE2L2, ABLIM2, AHRR, SMIM3, CDKN1A, TPST1, CNTNAP2, SNTG1, MTSS1, PTK2, ZC3H3, ZMIZ1, PTGDR2, PRSS23, GRIK4, ETV6, RARG, LOC348021, CCDC88C, ITPK1, ANPEP, RARA, SMIM6, RECQL5, F2RL3, LINC00111, ACOT9.
[45]
MS vs CTR
NAWM
Direct
BS-sequencing
PAD2
[46]
RRMS vs CTR
cfDNA (whole blood)
BS-PCR sequencing assay
MBP3, WM1.
[47]
MS vs CTR
NAWM
Bisulphite Illumina Methylation 450k Beadchip
ALDOA, ATP1A2, BCAR1, BRK1, CDK5, CORO1A, CSF3, DLC1, DTNBP1, FGD2, FMNL1, MLST8, MYBPC3, MYH6, MYH7, MYO1F, OBSCN, PDGFA, PRKCZ, SHC1, SIPA1L1, SSH3, TPM3, ADA, AGAP1, ALDOA, ARHGEF16, ATP1A1, ATP1A2, ATP1A4, ATP5H, BIN1, DAB2IP, DLC1, FGD2, LDHC, MACROD1, MLST8, MYBPC3, MYH6, MYH7, NME4, NT5C, PLXNB1, PTPRN2, RASA3, SEPT9, SIPA1L1, TBCD, TK1, ACSBG1, ACSL1, ACTR8, ADA, AGAP1, AGPAT1, AGRP, AKAP8, ALDH3A1, ALDOA, AMH, ANGPT2, APBB1IP, APEX, ARHGEF16, ATF6B, ATP11A, ATP1A1, ATP1A2, ATP1A4, ATP6V0E1, ATRIP, BBS2, BCAR1, BCL2L2, BIN1, BIRC5, BPI, BRD4, BRK1, C4B, CACNA1D, CASKIN1, CBX4, CCL17, CCL22, CD37, CD59, CDH1, CDK5, CHST3, CHURC1, CLASP1, CLIC5, CORO1A, CREB5, CRY2, CSF3, CSNK1E, CX3CL1, CXXC5, CYP21A2, DAB2IP, DAND5, DCPS, DHRS3, DLC1, DLL1, DOK4, DOT1L, DSCAML1, DTNBP1, DYRK1B, E2F6, E4F1, EDN2, EFS, ENTPD2, ERCC3, F7, FAM109A, FGD2, FGFR3, FMNL1, GBX1, GDF10, GPR114, GPR56, GTF2H1, GYLTL1B, HDAC11, HEG1, HEXIM1, HEXIM2, HIGD1A, HIST3H3, HLA-DMA, IL17RB, IL25, IL34, INO80E, INPP5J, INTS1, IRAK2, ITPKB, JARID2, LIMD1, LMF1, LPCAT1, MAB21L2, MADD, MAML3, MAP3K14, MAPK3, MBP, MCF2L, MED24, MEIS2, MLLT10, MLST8, MT1A, MT1E, MT1F, MT1G, MT1M, MT2A, MT4, MTCH1, MTSS1L, MUSK, MYBPC3, MYH6, MYH7, MYO1F, NARFL, NCOR2, NDRG1, NLRP3, NOTCH4, NR1H3, NUP210, OBSCN, OTX2, PABPN1, PAG1, PBX2, PCSK6, PDGFA, PEG10, PHF21A, PIK3R1, PLEKHG3, PLLP, PLXNB1, POLD4, POLR2C, POU2F1, PPARA, PPIL2, PPP1R13B, PPP4C, PRAM1, PRDM16, PRKCH, PRKCZ, PTGDS, PTPRN2, RAD9A, RAI1, RASA3, RBP1, RFX5, RIN2, RNF187, RPA1, RRM2, RXRA, SACS, SEMA4C, SETD1A, SHC1, SHISA5, SIPA1L1, SLC17A7, SLC22A17, SLC39A13, SLC7A8, SMAD6, SOX1, SOX8, SPI1, SPOCK2, SREBF1, SSH3, SSTR5, SUN1, TACC3, TBCD, TBX6, TEAD2, TEF, TEP1, THRA, TLN2, TNRC6C, TPM3, TRAF2, TSNARE1, UBE2L3, USP19, VAC14, WHSC1, WISP1, WISP2, WNK2, ZBTB47, ZFP1, ZIC1, ZNF135, ZNF256, ZNF329, ZNF362, ZNF414, ZNF418, ZNF488, ZNF606, ZNF664, ZNF687, ADAMDEC1, AIF1, AIRE, B2M, BPI, C1QA, C1QB, C1QC, C4BPA, C4BPB, CCR6, CD19, CD37, CD4, CD7, CD81, CFD, DLG1, FCER2, HAMP, HLA-DMA, HLA-DMB, HLA-DOA, HLA-DOB, HLA-DQA2, HLA-DQB2, HLA-F, IRF6, IRF8, IRF9, JAK1, JAK3, KYNU, LAG3, LAT, LBP, LCP2, LGMN, LST1, LTA, LTB, MBL2, MICB, NCR3, OSM, PSMB8, PTPN22, RARA, RNF31, SECTM1, SLAMF7, STXBP2, TAP1, TAP2, TAPBP, TNF, TNIP2, B2M, C1QA, C1QB, C1QC, C4BPA, C4BPB, DLG1, FCER2, HLA-DMA, LAG3, LTA, MBL2, NCR3, SLAMF7, TAP1, TAP2, TNF, B2M, C1QA, C1QB, C1QC, C4BPA, C4BPB, DLG1, FCER2, HLA-DMA, LAG3, LAT, LTA, MBL2, NCR3, SLAMF7, STXBP2, TAP1, TAP2, TNF, B2M, FCER2, HAMP, LAG3, MBL2, NCR3, SLAMF7, STXBP2, TAP1, TAP2, BHLHE23, CTSZ, DLG1, DLL1, DLX1, DLX2, EDARADD, EPHB4, FOXL2, GLI1, GNAS, HOXC11, HOXC13, HOXC4, HOXC8, HOXC9, HOXD10, HOXD11, HOXD13, HOXD3, HOXD4, HOXD8, HOXD9, MSX1, PHLDA2, PPP1R13L, PTCD2, RARA, RUNX3, SOX1, SOX8, TBX3, TEAD2, TGM1, TH, TNF, TWIST1, WNT2, ZIC1.
[48]
RRMS and CTR
CD4+ T cells
CD8+ T cells
Whole blood
Illumina 450K array
CD4+T cells: DCX, RDH13, DNHD1, TEKT5, TXNL1, MAGI2, TTC30B, APC2, TMEM48, ANGPTL2, RALGPS1, USP29, C20orf151, DLL1 6, DACH2, INPP5A, LOC727677, SEMA5B, SUGT1L1, HOXB2, OR10J5, RBMS1, C20orf151, AEN. CD8+T cells: APC2, HOXA2, HRNBP3, HEXDC, NTRK3, DCX, TRIL, ARHGEF17, ESPNP, LHX5, TEKT5, LRRC43, CYP27C1, TMEM48, HHATL, AMMECR1, C19orf45, SRRM3, PSD3, PTPRN2, LOC654342, ARHGEF17, DNHD1, KIF1C, INCA1, VSIG1. Whole blood: DACH2, LAMA2, TTLL8, GALNT9, POU3F4, NLRP12, PLS3, ANKRD1, CLSTN2, MAGEB4, APC2, PCDHA7, TMEM27, DNHD1, LGI1, PTCHD2, MMD2, HHATL, TMEM48, NXPH1, TDRD9, CDX1, YTHDC2, RGPD1, PLGLB2.
[49]
RRMS vs PPMS vs SPMS vs CTR
Buffy coat
BS-sequencing
SHP-1
[50]
MS vs CTR
Whole blood
PBMCs
NAWM
Illumina 450K array
IL2RA
[51]
MS treatment-naïve vs 1 year IFN-b vs CTR
PMBCs
BS-PCR sequencing assay
LINE-1
[52]
Discordant twins (RRMS vs CTR)
CD4+ T cells
RRBS
TMEM1, PEX14.
[53]
RRMS(e)vs
RRMS(r) vs CTR
Serum
BS-PCR sequencing assay
MOG
[54]
RRMS vs CTR
cfDNA (serum)
BS-PCR sequencing assay
LINE-1
[55]
RRMS vs CTR
CD3+ T cells
BS-PCR sequencing assay
VDR
[56]
RRMS(e) vs
RRMS(r) vs CTR
cfDNA (plasma)
MethDet-56
microarray based assay
RRMS(r) vs CTR: CDH1, CDKN2A, CDKN2B, FAS, ICAM1, MCJ, MDGI, MUC2, MYF3, PAX5, PGK1, RB1, SOCS1, SYK, TP73. RRMS(e) vs CTR:BRCA1, CCND2, DAPK, FAS, FHIT, MCT1, MDGI, MCJ, CDKN2A, TP73, PGK1, PR-PROX.RRMS(r) vs RRMS(e): CDH1, CDKN2B, HIC1, PR-PROX, SYK.
[57]
RRMS(e) vs
RRMS(r) vs CTR
Whole Blood
Methylation-Specific Multiple Ligation Probe Amplification PCR
CDKN2A, SOCS1, RUNX3, NEUROG1.
[58]
Discordant twins
(MS vs CTR)
PMBCs
CD4+ T cells
Bisulphite Illumina Methylation 450k Beadchip
TMEM232, SEMA3C, YWHAGI, ZBTB16, MRI1.
[59]
RRMS and SPMS
vs CTR
PMBCs
BS-PCR sequencing assay
PAD2
[60]
RRMS and SPMS
vs CTR
PMBCs
EpiTyper assay
DNMT1, TET2
[61]
RRMS vs SPMS
vs CTR
CD4+ T cells
Illumina 450K array
VMP1, MIR21
MS multiple sclerosis, CTR control, RRMS relapsing–remitting multiple sclerosis, PPMS primary progressive multiple sclerosis, SPMS secondary progressive multiple sclerosis, RRMS(e) RRMS in exacerbation, RRMS(r) RRMS in remission, cfDNA circulating-free DNA, PBMCs peripheral blood mononuclear cells, BS bisulphite, RRBS reduced representation bisulphite sequencing, NAWM normal appearing white matter

BBB breakdown

Infiltration of the autoreactive proinflammatory cells across the BBB into the brain is one of the pathological features of MS [62]. The BBB is a selective semi-permeable endothelium that separates the CNS from the circulating blood. This barrier is composed of a monolayer of endothelial cells tightly bound mainly by cadherins [63] and intercellular adhesion molecule (ICAM) proteins [64]. Cadherins are calcium-dependent adhesion molecules importantly involved in cell–cell adhesion [63]. The disruption of cell–cell interaction mediated by cadherins leads to BBB permeability [63]. A hypermethylated pattern of E-cadherin (CDH1) may increase the BBB permeability in relapsing–remitting MS (RRMS) patients favouring lymphocyte infiltration into the brain, and lastly, disease progression [47, 56]. The other adhesion molecules expressing on the BBB endothelium are the ICAM family. In particular, ICAM-1 is essential for leukocyte crawling prior to diapedesis from the bloodstream to the CNS [65] and plays a remarkable role in T cell proliferation [66]. Liggett et al. (2010) reported a hypermethylation pattern for ICAM1 in cell-free plasma DNA derived from RRMS patients in response to clinical remission, indicating an impairment of the T cell extravasation into the brain as a consequence of immune response mitigation [56]. These findings are in accordance with the results reported in knockout mice for Icam1 subjected to the experimental autoimmune/allergic encephalomyelitis (EAE) model [66].

Inflammation

The first inflammatory event in MS is conducted when APC through the class II MHC complex presents a specific antigen to naïve CD4+ T cells, which favour T cell differentiation and the recruitment of proinflammatory cells into the CNS [5]. MHC, also known as human leukocyte antigen (HLA), is responsible for presenting non-self-antigens to the T cell receptors and natural killer receptors (NKRs) [67] facilitating the inflammatory response. Leukocytes use the HLA complex to distinguish self-proteins from exogenous components [68]. In MS, certain HLA genes showed an aberrant methylation pattern contributing to MS aetiology [47]. For example, the hypomethylation of MHC class I polypeptide-related sequence B (MICB) has been reported in normal appearing white matter (NAWM) [47] and CD4+ T cells in MS patients [40]. In MS, a ligand codified by MICB activates the NK and CD8+ T cell destruction [69]. Similarly, the HLA-F variant is actively expressed in the inflammatory reaction [67] as a result of its promoter demethylation [43, 47]. Aside from the HLA complex, changes in DNAme are found in other inflammatory pathways reported in MS. Specifically, global CG island hypermethylation of the Src homology region 2 domain-containing phosphatase-1 and the suppressor of cytokine signalling 1 might aggravate the course of MS through the overactivation of the immune-mediated response [49, 56, 57]. Adhesion molecules such as ICAM5 are markedly present in the cerebral and hippocampal neurons [70]. In MS, the extracellular domain of ICAM5 is cleaved and released into the cerebrospinal fluid (CSF) and blood, where it modulates the synthesis of proinflammatory cytokines (TNF-α, IL-1β), stimulates the expression of anti-inflammatory cytokine IL-10 and represses T cell activation [71]. As a result of disease progression, primary progressive MS (PPMS) patients with non-recoverable demyelination and neurodegeneration showed higher methylation levels for ICAM5 than RRMS patients [43]. This result suggests that an overactivation of the inflammatory response in MS may be attributable to the aberrant methylation pattern of certain anti-inflammatory genes.

Demyelination

In MS, demyelination occurs when the myelin sheath of neurons is damaged by the action of the immune system [72]. The attack of immune surveillance is mainly directed against the myelin basic protein (MBP) [73], a protein that stabilises and maintains the correct structure of the myelin sheath around the axon [74]. An extensive hippocampal demyelination simultaneously coincides with the lower number of methyl groups to the DNMT promoters with an increase in their mRNA levels and a decrease in their TET enzymes [41]. Under normal conditions, approximately 20% of the total MBP is citrullinated (MBP-Cit). Citrullination is a post-translational modification catalysed by peptidyl arginine deiminase 2 (PAD2) [75]. The addition of citrullin groups leads to the loss of myelin compaction [76], and particularly, the percentage of MBP-Cit increases drastically [77] along with the promoter demethylation and overexpression of PAD2 as a result of the clinical course [45, 59]. The processing of MBP self-antigens and their presentation by APCs to T cells occurs during the negative selection of autoreactive T cells in the thymus [78]. An increase in the legumain (LGMN) activity, an enzyme involved in the self-antigen processing, prevents the development of immune tolerance against MBP [79]. Interestingly, the demethylation of the LGMN promoter could be responsible for favouring autoimmunity in MS patients [47, 74].

Remyelination failure

Following the myelin destruction, the recruitment of oligodendrocyte progenitor cells (OPCs) is necessary to rescue the demyelinated axons [80]. However, in MS patients, this process is not completely achieved [81], contributing to progressive neurodegeneration. The origin of this failure is not fully understood, but some hypotheses have been postulated in this regard [81]. Briefly, remyelination may be incomplete because of the inadequate recruitment of OPCs into the demyelinated lesion, an impairment of the OPC differentiation into myelinating oligodendrocytes, or the dysfunctions in oligodendrocytes when they attempt to wrap axons [81, 82]. In adults, OPC migration and recruitment require several growth factors including the platelet-derived growth factor (PDGF) and the fibroblast growth factor (FGF) [8385]. Both growth factors are significantly methylated in the NAWM of MS patients [47]. In this regard, the addition of methyl groups to the DNA may be accompanied by the lower expression of PDGF and FGF during disease progression, thus gaining mechanistic insight into the oligodendrocyte dysfunction. Wnt signalling pathway is involved in the differentiation of precursor oligodendocytes into mature myelinating oligodendocytes, affecting remyelination of axons [86]. Fancy et al. (2009) found that MS demyelinating lesions display an activation of the Wnt signalling impairing the remyelination process due to lack of proliferation of premyelinating oligodendrocytes [86]. Interestingly, some of the negative regulators of the Wnt pathway (histone deacetylase inhibitors, or Shisha genes) [87, 88] are hypermethylated in the brain [47, 41]. This hypermethylation would facilitate Wnt pathway activation declining efficiency of endogenous remyelination. This hypothesis is reinforced by Chen et al. (2011). They reported that mice lacking HDAC1/2 displayed severe myelin deficiency [89]. The hypermethylation of neurofascin (NFAS) has been related to the improper myelin wrapping of paranodal segments and might respond to the pathophysiological mechanisms underlying MS [90, 91]. Self-antibodies against NFAS have been detected in MS patients [92]. In the hippocampus while neurodegeneration is ongoing, Chomyk et al. (2017) reported a decrease in the WD repeat domain 81 and AT-hook transcription factor methylation status, which could lead to a lower mRNA expression and neuronal injury [41].

Neurodegeneration

Progressive axonal loss contributes to brain atrophy and neurological disability [93]. Kulakova et al. (2016) examined the DNAme of the peripheral blood mononuclear cells (PBMCs) collected from secondary progressive (SPMS) and RRMS patients, providing a time course DNAme pattern [43]. They found a cluster of 21 differentially methylated genes. Gene ontology analysis revealed that four of these genes are hypermethylated while neurodegeneration is ongoing and involved in biological processes, such as cell proliferation (PTH1R, CBFA2T3, KLF4) and nuclear factor κ-light-chain-enhancer of activated B cell (NF-kβ) pathway (KLF4, TRAF3). NF-kβ is a pleiotropic regulator of neuroinflammation, neuronal protection and neurotoxicity [94] involved in distinct pathological events mediated by the glia once the earliest signs of MS are present [95]. During the course of neurodegeneration, an imbalance between the expression of enzymes participating in the addition and removal of methyl groups to the DNA has been reported in MS patients. Indeed, Fagone et al. (2016) found an upregulation of methyl CpG-binding protein 2 and 4 and a downregulation of enzymes such as TET3 and TDG in SPMS patients [92].

Environmental and lifestyle risk factors

The influence of the environment on MS pathophysiology has been largely studied in twins discordant for the disease [52]. This divergence has been attributed to the effect of environmental risk factors through epigenetic modifications [96]. Thus, Handel et al. (2010) reported an additive deleterious effect of smoking, ultraviolet (UV) exposure and HLA alleles on MS prevalence [20]. Scientific evidence shows that environmental factors, such as smoking, Epstein Barr infection, vitamin D level, organic solvent and chemical pollutant exposure, diet style, gut microbiota, exercise and stress, are involved in the development and/or course of MS (Fig. 3).

Smoking

Smoking confers the risk for developing MS, and it is associated with disease onset and progression [97]. Degelman and Herman (2017) found a significant association between smoking frequency and conversion from RRMS to SPMS forms [98]. Nearly 98 chemical compounds of tobacco, including nicotine, cyanide and nitric oxide, are hazardous [99]. For example, nicotine increases the permeability of the BBB [100] in the earlier stages of MS [101], cyanide contributes to CNS demyelination [102] and nitric oxide promotes neurodegeneration [103]. Tobacco contains dioxins that activate the aryl-hydrocarbon receptor pathway [104], which modulates neuroinflammation [105] and Th17 and Treg activities [106] becoming a key player in the MS aetiology and disease progression. A well-designed study performed by Zeilinger et al. (2013) reported the differences in the DNAme of current, former and never smokers [107]. They found that the aryl-hydrocarbon receptor repressor (AHRR) was highly demethylated among the current smokers, leading to the inhibition of AhR signalling pathway and thus enhancing neuroinflammatory and neurodegeneration events [105, 106]. AhR can act as protective or deleterious pathway depending on the cell type where is expressed. In EAE model, AhR activation has a protective role in dendritic cells (DC), astrocytes and foxp3+ T reg cells, while it has a pro-inflammatory effect in Th17 cells [108]. AhR deletion is known to exacerbate EAE disease [109], altering myelin-associated proteins and increasing the production of proinflammatory cytokines [110]. In MS patients, low levels of AhR have been measured in serum, along with a reduced AhR activity in demyelinating lesions during disease progression [111]. Interestingly, Laquinimod, a phase III drug for MS treatment that activated AhR pathway, was shown to reduce brain atrophy in MS patients by counteracting the neuroinflammatory reaction [112].

Epstein Barr virus

The latent form of EBV is present in almost 90% of the world population within memory B cells [113]; more than 99% of MS patients are seropositive for EBV [24]. Infection with EBV increases the risk of developing MS by ~ 3.6-fold [114], and its effect can be exacerbated when interacts with other risk factor, such as HLA risk variants, rocketing the odd ratio for MS up to ~ 15-fold [115]. We can speculate that a cross-reaction may occur between the myelin self-antigens and certain viral proteins of the EBV in MS [116]. This hypothesis is underpinned by the identification of two EBV peptides (EBNA-1 and BRRF2) in the CSF of MS patients [117]. Indeed, the CD8+ T cells derived from MS patients displayed reactivity against some latent EBV proteins [117]. The reactivation of latent EBV in memory B cells may occur through specific epigenetic modifications. For example, the use of 5-azacytidine, a DNMT inhibitor, can switch the EBV latent form to a reactivated form [118].

Vitamin D

Vitamin D is synthesised upon exposure to UV [119] or can be obtained from diet [120]. Vitamin D deficiency is considered a risk factor for MS even before birth, altering the structure of embryonic tissues as a result of low maternal vitamin D [121]. Indeed, the prevalence of MS is lower near the equator, where UV radiation is at its maximum, than at higher and lower latitudes [122]. Low levels of vitamin D have been associated with a higher frequency of relapses [123] and disability [121].
Current evidence points out that the active form of vitamin D acts as an inducer of the promoter demethylation of multiple genes [124]. In accordance with these findings, Rawson et al. (2012) reported that a high intake of vitamin D is associated with lower methylation levels of certain genes involved in the Wnt signalling pathway [125], which could favour neurogenesis and neuronal plasticity [126]. In T cells from RRMS whole peripheral blood, the vitamin V receptor (VDR) showed a promoter demethylation pattern associated with an overexpression of VDR mRNA [55]. Nevertheless, the underlying mechanism behind vitamin D and its effect on demethylation remains unknown and requires further investigation.

Organic solvents and pesticides

Organic solvents are hydrocarbon compounds commonly used worldwide. The prolonged exposure to these compounds has severe effects on health and may be clinically important in autoimmune diseases [127]. Organic solvents are highly hydrophilic and lipophilic molecules able to pass through the BBB into the brain, resulting in distinct myelin pathologies [128]. Exposure to organic solvents induces changes in the DNAme of the immune system [129] and certain genes involved in cell survival [130]. Hexachlorobenzene, a pesticide widely used until 1965, modulates the expression of E-cadherin through the activation of the integrin-linked kinase signalling [131]. E-cadherin plays an important role in BBB integrity, and its promoter region is highly methylated in MS patients [49], facilitating the infiltration of immune cell into the brain. To evaluate the precise contribution of this pesticide in the context of MS, we should determine the exposure time frame and if it occurred before or after the diagnosis of disease. Remarkably, the precise effect of organic solvents on DNAme in MS warrants further investigation. However, exposure to organic solvents and the presence of HLA risk alleles present a fourfold increased risk for MS in comparison with exposure to organic solvents alone, and the additive effect of organic solvents, HLA risk alleles and smoking increase the risk 20-fold [132].

Diet

Recent evidence points out that food intake is important in the pathogenesis of MS and other autoimmune diseases. Seasonal variations in food intake in pregnant mothers may interfere with foetal development and confer the risk of MS in the later life of the offspring (for a review, see [133]). The intake of some nutrients, such as long-chain fatty acids and salt, is known to act as pro-inflammatory molecules, whereas other nutrients, such as short-chain fatty acids and flavonoids, possess anti-inflammatory properties [134]. A type of flavonoid called quercetin represses the capacity of monocytes to cross the BBB [135] and reduces the synthesis of proinflammatory cytokines produced by monocytes in MS [136]. Limited data are available on the intake of nutrients and their effect on DNAme. However, certain cofactors obtained from diet are required to maintain DNAme homeostasis. For example, vitamin B, folate, methionine, choline and zinc are essential to maintain the levels of Dnmt1 [137] and methyl-donor S-adenosylmethionine [137]. By contrast, vitamin C can act as a cofactor for TET enzymes [138]. In particular, vitamin B12 is necessary for the proper function of the CNS through the conversion of homocysteine to methionine. Methionine regulates the expression of DNA methyltransferase 3A (DNMT3A) [139] and participates in the transcription of pro-inflammatory genes and the formation of myelin sheath [140]. Interestingly, vitamin B12 and folate deficiency is reported in RRMS concomitantly with elevated levels of homocysteine. A misbalance in DNAme metabolism has been reported in MS patients [139, 141]. Indeed, some authors found that the levels of methyl group donors were lower in plasma [139] and post-mortem grey matter [141] in MS.
The elevated consumption of salt has many noxious effects, including the production of reactive oxygen species [142], deregulation of regulatory T cells [143] and macrophage [144], disruption of BBB permeability [145] and higher probability of new brain lesions in RRMS patients [146]. After the ingestion of a diet containing high levels of salt, patients with an autoimmune disease showed demethylation and elevated levels of hydroxymethylation in CD4+ T cells as a result of TET2 overexpression [25]. Consequently, the high consumption of salt may be a risk factor for MS. Note that the use of standardised diets or alternative therapies cannot substitute conventional MS treatments, but the intake of healthy food may ameliorate the inflammatory and physical status of MS patients. Therefore, it can be hypothesised that some components of diet may interfere with the DNAme metabolism even before childbirth or during lifetime contributing to both disease aetiology and progression.

Gut microbiota

Gut microbiota is a complex ecosystem of microorganisms that establish a symbiotic relationship with the host by favouring the vitamin production and fermentation of some components of diet [147]. However, the dysbiosis of microbiota increases the risk of developing autoimmune diseases [148, 149]. DC from germ-free mice subjected to the EAE model were less reactive to stimulated proinflammatory T cells than conventionally colonised germ mice [149]. Berer et al. (2011) found that gut microbiota in cooperation with myelin auto-antigens is necessary to stimulate the immune response [148], consistent with [149]. Dysbiosis has been associated with an inflammatory phenotype in MS patients [150, 151]. Resident microbiota can alter the epigenetic signature of the host through the production of specific metabolites (Table 2). Pregnant women showed a different DNAme profile according to their predominant microbiota in the gut [166]. Therefore, dysbiosis of microbiota can be involved in disease onset by an overactivation of T cells [148, 149] or exacerbating inflammatory events in patients diagnosed with MS [150, 151]. However, the contribution of microbiota in disease aetiology and progression warrants further investigation.
Table 2
List of metabolites released by microbiota
Metabolite
Effect on DNA methylation
p-Cresol
It induces the expression of DNA methyltransferases 1, 3a, and 3b and it is associated with CpG hypermethylation of Klotho gene [152], a regulator of vitamin D metabolism [153].
Hydrogen sulphide (H2S)
Involved in the neutralisation of ROS. It increases DNA methylation [154].
Riboflavin (vitamin B2)
Pyridoxine (vitamin B6)
Cobalamin (vitamin B12)
Cofactor involved in DNA methylation metabolism [155, 156].
Folate (vitamin B9)
It acts as a methyl donor involved in DNA methylation metabolism [155, 156].
It reduces the activity of DNA methyltransferase [157].
Choline
It acts as a methyl donor that can be recruited by human gut microbiota, reducing its availability [158].
Involved in DNA methylation and gene expression in murine colitis model, an inflammatory disease [159].
Betaine
It acts as a methyl donor involved in DNA methylation reactions [156, 160].
Associated with changes in DNA methyltransferases and coupled with changes in DNA methylation [161].
Ammonium (NH4)
Inverse correlation between faecal NH3 and LINE-1 gene methylation [162].
α-ketoglutarate
Involved in (de)methylation as a co-factor of histone demethylases and TET family [163, 164].
L-ascorbic acid (vitamin C)
It exerts a strong influence on active DNA demethylation. It enhances TET-mediated generation of 5-hydroxymethylation [165].
ROS reactive oxygen species, NH3 ammonia, LINE-1 long interspersed element-1, TET ten–eleven translocation
Adapted from Mischke et al. [147]

Physical activity

Physical exercise has been demonstrated to produce changes in the leukocyte DNAme pattern [167] and thus changes in gene expression [168]. In the context of neurodegenerative diseases, the influence of physical exercise has a direct effect on the brain-derived neurotrophic factor (BDNF) [169, 170]. In a healthy brain, BDNF is mainly expressed in neurons [171]. However, following a demyelinating insult, this gene is transcriptionally active in astrocytes [172], regulatory T cells, B cells and monocytes [172, 173], favouring brain plasticity [174] and myelin formation [175]. Recently, Briken et al. (2016) found elevated protein levels of serum BDNF after 30 min of exercise in progressive MS patients [170], and this result was probably attributable to the demethylation of its promoter region [169]. Similarly, 2 weeks of physical exercise promotes the overexpression of TET1 and the demethylation of the vascular endothelial growth factor A (VEGF-A) [176]. VEGF-A may potentiate neurogenesis and neuroprotection in the EAE model as postulated by [177]. The overexpression of apoptosis-associated, speck-like protein containing a C-terminal caspase recruitment domain gene (ASC) activates inflammatory signalling and may exacerbate MS progression [178]. In addition, Nakajima et al. (2010) found that following 6 months of moderate exercise, the mRNA levels of ASC were lower because its promoter region was hypermethylated [179]. Therefore, the current data support the notion that moderate exercise can reduce pro-inflammatory cytokines and improve the clinical MS course. However, some studies failed to validate this evidence [180182]. The findings reported here point out that moderate exercise can ameliorate some symptoms but cannot stop the progression of MS.

Stress

Stressful life events have a negative effect on MS by increasing the risk of clinical exacerbation and disease progression [183]. However, the contribution of stress in the pathophysiology of the disease remains under discussion. A well-designed study by Liu et al. (2009) reported a strong association between stress and MS aetiology [184]. In accordance with these findings, Babenko et al. (2015) demonstrated that prenatal stress causes a demethylation of the nuclear subfamily 3 group C member 1 glucocorticoid receptor (NR3C1) [185] and changes in the nervous, immune and musculoskeletal systems [186]. Interestingly, the differences in NR3C1 gene expression have been reported in MS patients [187], suggesting that early life stressors can present susceptibility to developing MS in adulthood.

DNA methylation in animal models of MS

None of the current experimental animal models can reproduce the complexity and heterogeneity of MS. In particular, both disease onset and clinical course in animals differ considerably from those in humans [188]. However, in vivo experimental models are widely used to understand certain aspects of the disease. In general, we consider three models to study MS pathophysiology: (a) EAE, (b) Theiler’s murine encephalomyelitis virus (TMEV) and (c) use of toxins such as cuprizone (CPZ) or lysolecithin. However, not all of them have been addressed to study the changes in DNAme. For example, no current study has been conducted on DNAme in the TMEV model.

EAE model

EAE is a well-established model of autoimmunity induced by the subcutaneous injection of self-antigens derived from myelin proteins, such as the myelin oligodendrocyte glycoprotein (MOG) [189] and the proteolipid protein (PLP) [190]. Catanzaro et al. (2016) characterised the DNAme profile in the striatum of EAE mice showing a global DNA hypomethylation of interneurons positive for neuronal nitric oxide synthase [191]. Interestingly, they found a demethylation of Ras-related protein-1 (Dexras-1) in parallel with elevated levels of iron inside the cells and thus neurotoxicity and neuronal death. Furthermore, they reported that the hypomethylation of Dexras-1 was reverted when mice were subjected to an enriched environment in their home cage, emphasising an epigenetic-mediated effect [191]. Recently, Noori-Zadeh et al. (2017) found that the promoter region for forkhead box P3 was hypermethylated in T cells collected from EAE mice [192], thus indicating a dysfunction of regulatory T lineage and the lack of auto-immune tolerance [193].

Toxin-induced demyelination

Demyelination can be induced by copper-chelating agents (e.g. CPZ) or lysolecithin [194]. In the CPZ model, young adult mice fed with this neurotoxicant showed a significant loss of mature oligodendrocytes, astrocytosis, microgliosis and demyelination, followed by spontaneous remyelination [195]. The CPZ model was used by Olsen et al. (2019) to identify novel biomarkers related to the demyelination course [53]. Specifically, they isolated circulating-free DNA from mice blood at the end of the CPZ treatment, and identified a specific methylation pattern associated with oligondendrocyte apoptosis. Conversely, the use of lysolecithin in mice induced demyelination accompanied by a high expression of DNMT1 in the OPCs at the early stages of remyelination. By contrast, DNMT3A is highly expressed in oligodendrocytes at the later stages when remyelination is achieved. The study revealed a global hypermethylation in the oligodendrocyte lineage during remyelination, demonstrating that DNMT1 plays a crucial role in the proliferation and differentiation of OPCs into mature oligodendrocytes, while DNMT3A has a dominant role in the remyelination phase [196].

Conclusions

MS is an inflammatory autoimmune disease of the CNS caused by a complex interaction between genetic and environmental factors [20]. Emerging evidence indicates that DNAme actively participates in gene x environment interactions [33]. As previously mentioned, several studies showed an aberrant DNAme profile in relapsing–remitting forms and in progressive MS forms. Remarkably, most of the studies reported in this work were based on the bisulphite technique (Table 1). However, this approximation does not discriminate between 5-methylcytosine and 5-hydroxymethylcytosine and thus may contribute to a misinterpretation of the data. To avoid this bias, we recommend an alternative method for studying DNAme, such as the methylated-DNA immunoprecipitation and the TET-assisted bisulphate sequencing. As far as we know, genetic factors can explain approximately 30% of worldwide MS prevalence [23], and the remaining 70% may correspond to the influence of environmental risk factors. As described in this study, UV radiation, cigarette smoking and infection with the Epstein Barr virus are clinically relevant for MS, although other environmental factors, such as diet style, microbiota profile, exposure to organic solvents and pollutants, exercise and long-term stress, have a clear effect in MS. All of the aforementioned risk factors can modify the DNAme pattern in humans, but further studies are required to expand our knowledge of the molecular basis of the disease and elucidate the underlying mechanisms behind MS pathophysiology. Furthermore, DNA methylation is currently the best surrogate marker for epigenetic change in disease, because methylation alterations track with disease state. DNA methylation markers can also indicate success or failure of drug treatment, are stable in isolated DNA, and can be measured by a variety of quantitative and qualitative methods. We postulate that epigenetic DNAme marks described in the context of the disease can potentially be used in a specific, substantial, and credible way in clinical interventions. It is conceivable that, in the near future, we will be able to design drugs modifying DNAme metabolism to stop the progression of MS.

Acknowledgements

Not applicable.
Not applicable.
All authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Ghasemi N, Razavi S, Nikzad E. Multiple sclerosis: pathogenesis, symptoms, diagnoses and cell-based therapy citation. Cell J. 2017;19:1–10.PubMed Ghasemi N, Razavi S, Nikzad E. Multiple sclerosis: pathogenesis, symptoms, diagnoses and cell-based therapy citation. Cell J. 2017;19:1–10.PubMed
2.
Zurück zum Zitat De Stefano N, Narayanan S, Francis GS, Arnaoutelis R, Tartaglia MC, Antel JP, et al. Evidence of axonal damage in the early stages of multiple sclerosis and its relevance to disability. Arch Neurol. 2001;58:65–70.PubMedCrossRef De Stefano N, Narayanan S, Francis GS, Arnaoutelis R, Tartaglia MC, Antel JP, et al. Evidence of axonal damage in the early stages of multiple sclerosis and its relevance to disability. Arch Neurol. 2001;58:65–70.PubMedCrossRef
3.
Zurück zum Zitat Compston A, Coles A. Seminar multiple sclerosis; 2008. Compston A, Coles A. Seminar multiple sclerosis; 2008.
6.
Zurück zum Zitat Sattler A, Wagner U, Rossol M, Sieper J, Wu P, Krause A, et al. Cytokine-induced human IFN-γ-secreting effector-memory Th cells in chronic autoimmune inflammation. Blood. 2009;113:1948–56.PubMedCrossRef Sattler A, Wagner U, Rossol M, Sieper J, Wu P, Krause A, et al. Cytokine-induced human IFN-γ-secreting effector-memory Th cells in chronic autoimmune inflammation. Blood. 2009;113:1948–56.PubMedCrossRef
7.
Zurück zum Zitat Young HA, Hardy KJ. Role of interferon-γ in immune cell regulation. J Leukoc Biol. 1995;58:373–81.PubMedCrossRef Young HA, Hardy KJ. Role of interferon-γ in immune cell regulation. J Leukoc Biol. 1995;58:373–81.PubMedCrossRef
8.
Zurück zum Zitat Takeshita Y, Ransohoff RM. Inflammatory cell trafficking across the blood-brain barrier: chemokine regulation and in vitro models. Immunol Rev. 2012;248:228–39.PubMedPubMedCentralCrossRef Takeshita Y, Ransohoff RM. Inflammatory cell trafficking across the blood-brain barrier: chemokine regulation and in vitro models. Immunol Rev. 2012;248:228–39.PubMedPubMedCentralCrossRef
10.
Zurück zum Zitat Leech S, Kirk J, Plumb J, McQuaid S. Persistent endothelial abnormalities and blood-brain barrier leak in primary and secondary progressive multiple sclerosis. Neuropathol Appl Neurobiol. 2007;33:86–98.PubMedCrossRef Leech S, Kirk J, Plumb J, McQuaid S. Persistent endothelial abnormalities and blood-brain barrier leak in primary and secondary progressive multiple sclerosis. Neuropathol Appl Neurobiol. 2007;33:86–98.PubMedCrossRef
15.
Zurück zum Zitat Najafi S, Mirshafiey A. The effect of activated microglia in progression of multiple sclerosis. Int Trends Immun. 2015;3:96–104. Najafi S, Mirshafiey A. The effect of activated microglia in progression of multiple sclerosis. Int Trends Immun. 2015;3:96–104.
18.
Zurück zum Zitat Jelcic I, Al Nimer F, Wang J, Lentsch V, Planas R, Jelcic I, et al. Memory B cells activate brain-homing, autoreactive CD4+ T cells in multiple sclerosis. Cell. 2018;175:85–100 e23.PubMedPubMedCentralCrossRef Jelcic I, Al Nimer F, Wang J, Lentsch V, Planas R, Jelcic I, et al. Memory B cells activate brain-homing, autoreactive CD4+ T cells in multiple sclerosis. Cell. 2018;175:85–100 e23.PubMedPubMedCentralCrossRef
19.
Zurück zum Zitat Stassart RM, Möbius W, Nave KA, Edgar JM. The axon-myelin unit in development and degenerative disease. Front Neurosci. 2018;12:467. Stassart RM, Möbius W, Nave KA, Edgar JM. The axon-myelin unit in development and degenerative disease. Front Neurosci. 2018;12:467.
20.
Zurück zum Zitat Handel AE, Handunnetthi L, Giovannoni G, Ebers GC, Ramagopalan S V. Genetic and environmental factors and the distribution of multiple sclerosis in Europe. Eur J Neurol. 2010;17:1210–4.PubMedCrossRef Handel AE, Handunnetthi L, Giovannoni G, Ebers GC, Ramagopalan S V. Genetic and environmental factors and the distribution of multiple sclerosis in Europe. Eur J Neurol. 2010;17:1210–4.PubMedCrossRef
21.
Zurück zum Zitat Meaney MJ. Epigenetics and the biological definition of gene X environment interactions. Child Dev. 2010;81:41–79.PubMedCrossRef Meaney MJ. Epigenetics and the biological definition of gene X environment interactions. Child Dev. 2010;81:41–79.PubMedCrossRef
22.
23.
Zurück zum Zitat Lill CM. Recent advances and future challenges in the genetics of multiple sclerosis. Front Neurol. 2014;5:1–5.CrossRef Lill CM. Recent advances and future challenges in the genetics of multiple sclerosis. Front Neurol. 2014;5:1–5.CrossRef
24.
Zurück zum Zitat Ascherio A, Munger KL. Environmental risk factors for multiple sclerosis. Part I: the role of infection. Ann Neurol. 2007;61:288–99.PubMedCrossRef Ascherio A, Munger KL. Environmental risk factors for multiple sclerosis. Part I: the role of infection. Ann Neurol. 2007;61:288–99.PubMedCrossRef
25.
Zurück zum Zitat Wu H, Huang X, Qiu H, Zhao M, Liao W, Yuan S, et al. High salt promotes autoimmunity by TET2-induced DNA demethylation and driving the differentiation of Tfh cells. Sci Rep. 2016;6:1–14.CrossRef Wu H, Huang X, Qiu H, Zhao M, Liao W, Yuan S, et al. High salt promotes autoimmunity by TET2-induced DNA demethylation and driving the differentiation of Tfh cells. Sci Rep. 2016;6:1–14.CrossRef
27.
Zurück zum Zitat Ehrlich M, Gama-Sosa MA, Huang LH, Midgett RM, Kuo KC, Mccune RA, et al. Amount and distribution of 5-methylcytosine in human DNA from different types of tissues or cells. Nucleic Acids Res. 1982;10:2709–21.PubMedPubMedCentralCrossRef Ehrlich M, Gama-Sosa MA, Huang LH, Midgett RM, Kuo KC, Mccune RA, et al. Amount and distribution of 5-methylcytosine in human DNA from different types of tissues or cells. Nucleic Acids Res. 1982;10:2709–21.PubMedPubMedCentralCrossRef
28.
Zurück zum Zitat Weber M, Hellmann I, Stadler MB, Ramos L, Pääbo S, Rebhan M, et al. Distribution, silencing potential and evolutionary impact of promoter DNA methylation in the human genome. Nat Genet. 2007;39:457–66.PubMedCrossRef Weber M, Hellmann I, Stadler MB, Ramos L, Pääbo S, Rebhan M, et al. Distribution, silencing potential and evolutionary impact of promoter DNA methylation in the human genome. Nat Genet. 2007;39:457–66.PubMedCrossRef
29.
31.
Zurück zum Zitat Van Emburgh BO, Robertson KD. Modulation of Dnmt3b function in vitro by interactions with Dnmt3L, Dnmt3a and Dnmt3b splice variants. Nucleic Acids Res. 2011;39:4984–5002.PubMedPubMedCentralCrossRef Van Emburgh BO, Robertson KD. Modulation of Dnmt3b function in vitro by interactions with Dnmt3L, Dnmt3a and Dnmt3b splice variants. Nucleic Acids Res. 2011;39:4984–5002.PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Kagiwada S, Kurimoto K, Hirota T, Yamaji M, Saitou M. Replication-coupled passive DNA demethylation for the erasure of genome imprints in mice. EMBO J. 2013;32:340–53.PubMedCrossRef Kagiwada S, Kurimoto K, Hirota T, Yamaji M, Saitou M. Replication-coupled passive DNA demethylation for the erasure of genome imprints in mice. EMBO J. 2013;32:340–53.PubMedCrossRef
36.
Zurück zum Zitat Ito S, Shen L, Dai Q, Wu SC, Collins LB, Swenberg JA, et al. Tet proteins can convert 5-Methylcytosine to 5-Formylcytosine and 5-Carboxylcytosine. Science. 2011;333:1300–3.PubMedPubMedCentralCrossRef Ito S, Shen L, Dai Q, Wu SC, Collins LB, Swenberg JA, et al. Tet proteins can convert 5-Methylcytosine to 5-Formylcytosine and 5-Carboxylcytosine. Science. 2011;333:1300–3.PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat Cortellino S, Xu J, Sannai M, Moore R, Caretti E, Cigliano A, et al. Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair. Cell. 2011;146:67–79.PubMedPubMedCentralCrossRef Cortellino S, Xu J, Sannai M, Moore R, Caretti E, Cigliano A, et al. Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair. Cell. 2011;146:67–79.PubMedPubMedCentralCrossRef
39.
40.
Zurück zum Zitat Graves MC, Benton M, Lea RA, Boyle M, Tajouri L, Macartney-Coxson D, et al. Methylation differences at the HLA-DRB1 locus in CD4+ T-cells are associated with multiple sclerosis. Mult Scler J. 2014;20:1033–41.PubMedCrossRef Graves MC, Benton M, Lea RA, Boyle M, Tajouri L, Macartney-Coxson D, et al. Methylation differences at the HLA-DRB1 locus in CD4+ T-cells are associated with multiple sclerosis. Mult Scler J. 2014;20:1033–41.PubMedCrossRef
42.
Zurück zum Zitat Fagone P, Mangano K, di Marco R, Touil-Boukoffa C, Chikovan T, Signorelli S, et al. Expression of DNA methylation genes in secondary progressive multiple sclerosis. J Neuroimmunol. 2016;290:66–9.PubMedCrossRef Fagone P, Mangano K, di Marco R, Touil-Boukoffa C, Chikovan T, Signorelli S, et al. Expression of DNA methylation genes in secondary progressive multiple sclerosis. J Neuroimmunol. 2016;290:66–9.PubMedCrossRef
43.
Zurück zum Zitat Kulakova OG, Kabilov MR, Danilova LV, Popova EV, Baturina OA, Tsareva EY, et al. Whole-genome DNA methylation analysis of peripheral blood mononuclear cells in multiple sclerosis patients with different disease courses. Acta Naturae. 2016;8:103–10.PubMedPubMedCentralCrossRef Kulakova OG, Kabilov MR, Danilova LV, Popova EV, Baturina OA, Tsareva EY, et al. Whole-genome DNA methylation analysis of peripheral blood mononuclear cells in multiple sclerosis patients with different disease courses. Acta Naturae. 2016;8:103–10.PubMedPubMedCentralCrossRef
44.
Zurück zum Zitat Marabita F, Almgren M, Sjöholm LK, Kular L, Liu Y, James T, et al. Smoking induces DNA methylation changes in multiple sclerosis patients with exposure-response relationship. Sci Rep. 2017;7:1–15. Marabita F, Almgren M, Sjöholm LK, Kular L, Liu Y, James T, et al. Smoking induces DNA methylation changes in multiple sclerosis patients with exposure-response relationship. Sci Rep. 2017;7:1–15.
45.
Zurück zum Zitat Mastronardi FG, Noor A, Wood DD, Paton T, Moscarello MA. Peptidyl argininedeiminase 2 CpG island in multiple sclerosis white matter is hypomethylated. J Neurosci Res. 2007;85:2006–16.PubMedCrossRef Mastronardi FG, Noor A, Wood DD, Paton T, Moscarello MA. Peptidyl argininedeiminase 2 CpG island in multiple sclerosis white matter is hypomethylated. J Neurosci Res. 2007;85:2006–16.PubMedCrossRef
48.
Zurück zum Zitat Bos SD, Page CM, Andreassen BK, Elboudwarej E, Gustavsen MW, Briggs F, et al. Genome-wide DNA methylation profiles indicate CD8+ T cell hypermethylation in multiple sclerosis. PLoS One. 2015;10:1–16.PubMedPubMedCentralCrossRef Bos SD, Page CM, Andreassen BK, Elboudwarej E, Gustavsen MW, Briggs F, et al. Genome-wide DNA methylation profiles indicate CD8+ T cell hypermethylation in multiple sclerosis. PLoS One. 2015;10:1–16.PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat Pinto-Medel MJ, Oliver-Martos B, Urbaneja-Romero P, Hurtado-Guerrero I, Ortega-Pinazo J, Serrano-Castro P, et al. Global methylation correlates with clinical status in multiple sclerosis patients in the first year of IFNbeta treatment. Sci Rep. 2017;7:1–9. Pinto-Medel MJ, Oliver-Martos B, Urbaneja-Romero P, Hurtado-Guerrero I, Ortega-Pinazo J, Serrano-Castro P, et al. Global methylation correlates with clinical status in multiple sclerosis patients in the first year of IFNbeta treatment. Sci Rep. 2017;7:1–9.
52.
53.
Zurück zum Zitat Olsen JA, Kenna LA, Tipon RC, Spelios MG, Stecker MM, Akirav EM. A minimally-invasive blood-derived biomarker of oligodendrocyte cell-loss in multiple sclerosis. EBioMedicine. 2016;10:227–35.PubMedPubMedCentralCrossRef Olsen JA, Kenna LA, Tipon RC, Spelios MG, Stecker MM, Akirav EM. A minimally-invasive blood-derived biomarker of oligodendrocyte cell-loss in multiple sclerosis. EBioMedicine. 2016;10:227–35.PubMedPubMedCentralCrossRef
54.
Zurück zum Zitat Dunaeva M, Derksen M, Pruijn GJM. LINE-1 Hypermethylation in serum cell-free DNA of relapsing remitting multiple sclerosis patients. Mol Neurobiol. 2018;55:4681–8.PubMedPubMedCentralCrossRef Dunaeva M, Derksen M, Pruijn GJM. LINE-1 Hypermethylation in serum cell-free DNA of relapsing remitting multiple sclerosis patients. Mol Neurobiol. 2018;55:4681–8.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Ayuso T, Aznar P, Soriano L, Olaskoaga A, Roldán M, Otano M, et al. Vitamin D receptor gene is epigenetically altered and transcriptionally up-regulated in multiple sclerosis. PLoS One. 2017;12:1–10.PubMedPubMedCentralCrossRef Ayuso T, Aznar P, Soriano L, Olaskoaga A, Roldán M, Otano M, et al. Vitamin D receptor gene is epigenetically altered and transcriptionally up-regulated in multiple sclerosis. PLoS One. 2017;12:1–10.PubMedPubMedCentralCrossRef
57.
Zurück zum Zitat Sokratous M, Dardiotis E, Bellou E, Tsouris Z, Michalopoulou A, Dardioti M, et al. CpG island methylation patterns in relapsing-remitting multiple sclerosis. J Mol Neurosci. 2018;64:478–84.PubMedCrossRef Sokratous M, Dardiotis E, Bellou E, Tsouris Z, Michalopoulou A, Dardioti M, et al. CpG island methylation patterns in relapsing-remitting multiple sclerosis. J Mol Neurosci. 2018;64:478–84.PubMedCrossRef
58.
59.
Zurück zum Zitat Calabrese R, Zampieri M, Mechelli R, Annibali V, Guastafierro T, Ciccarone F, et al. Methylation-dependent PAD2 upregulation in multiple sclerosis peripheral blood. Mult Scler J. 2012;18:299–304.CrossRef Calabrese R, Zampieri M, Mechelli R, Annibali V, Guastafierro T, Ciccarone F, et al. Methylation-dependent PAD2 upregulation in multiple sclerosis peripheral blood. Mult Scler J. 2012;18:299–304.CrossRef
60.
Zurück zum Zitat Calabrese R, Valentini E, Ciccarone F, Guastafierro T, Bacalini MG, Ricigliano VAG, et al. TET2 gene expression and 5-hydroxymethylcytosine level in multiple sclerosis peripheral blood cells. Biochim Biophys Acta - Mol Basis Dis. 1842;2014:1130–6.CrossRef Calabrese R, Valentini E, Ciccarone F, Guastafierro T, Bacalini MG, Ricigliano VAG, et al. TET2 gene expression and 5-hydroxymethylcytosine level in multiple sclerosis peripheral blood cells. Biochim Biophys Acta - Mol Basis Dis. 1842;2014:1130–6.CrossRef
61.
Zurück zum Zitat Ruhrmann S, Ewing E, Piket E, Kular L, Cetrulo Lorenzi JC, Fernandes SJ, et al. Hypermethylation of MIR21 in CD4+ T cells from patients with relapsingremitting multiple sclerosis associates with lower miRNA-21 levels and concomitant up-regulation of its target genes. Mult Scler J. 2017;24:1288–1300.PubMedCrossRef Ruhrmann S, Ewing E, Piket E, Kular L, Cetrulo Lorenzi JC, Fernandes SJ, et al. Hypermethylation of MIR21 in CD4+ T cells from patients with relapsingremitting multiple sclerosis associates with lower miRNA-21 levels and concomitant up-regulation of its target genes. Mult Scler J. 2017;24:1288–1300.PubMedCrossRef
63.
Zurück zum Zitat On NH, Kiptoo P, Siahaan TJ, Miller DW. Modulation of blood-brain barrier permeability in mice using synthetic E-cadherin peptide. Mol Pharm. 2014;11:974–81.PubMedPubMedCentralCrossRef On NH, Kiptoo P, Siahaan TJ, Miller DW. Modulation of blood-brain barrier permeability in mice using synthetic E-cadherin peptide. Mol Pharm. 2014;11:974–81.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Montesinos ML. Cell adhesion molecules: implications in neurological diseases; 2014. Montesinos ML. Cell adhesion molecules: implications in neurological diseases; 2014.
65.
Zurück zum Zitat Greenwood J, Heasman SJ, Alvarez JI, Prat A, Lyck R, Engelhardt B. Review: Leucocyte-endothelial cell crosstalk at the blood-brain barrier: a prerequisite for successful immune cell entry to the brain. Neuropathol Appl Neurobiol. 2011;37:24–39.PubMedCrossRef Greenwood J, Heasman SJ, Alvarez JI, Prat A, Lyck R, Engelhardt B. Review: Leucocyte-endothelial cell crosstalk at the blood-brain barrier: a prerequisite for successful immune cell entry to the brain. Neuropathol Appl Neurobiol. 2011;37:24–39.PubMedCrossRef
67.
Zurück zum Zitat Dulberger CL, McMurtrey CP, Hölzemer A, Neu KE, Liu V, Steinbach AM, et al. Human leukocyte antigen F presents peptides and regulates immunity through interactions with NK cell receptors. Immunity. 2017;46:1018–29 e7.PubMedPubMedCentralCrossRef Dulberger CL, McMurtrey CP, Hölzemer A, Neu KE, Liu V, Steinbach AM, et al. Human leukocyte antigen F presents peptides and regulates immunity through interactions with NK cell receptors. Immunity. 2017;46:1018–29 e7.PubMedPubMedCentralCrossRef
68.
Zurück zum Zitat Frohn C, Fricke L, Puchta JC, Kirchner H. The effect of HLA-C matching on acute renal transplant rejection. Nephrol Dial Transplant. 2001;16:355–60.CrossRef Frohn C, Fricke L, Puchta JC, Kirchner H. The effect of HLA-C matching on acute renal transplant rejection. Nephrol Dial Transplant. 2001;16:355–60.CrossRef
70.
Zurück zum Zitat Yoshihara Y, Oka S, Nemoto Y, Watanabe Y, Nagata S, Kagamiyama H, et al. An ICAM-related neuronal glycoprotein, telencephalin, with brain segment-specific expression. Neuron. 1994;12:541–53.PubMedCrossRef Yoshihara Y, Oka S, Nemoto Y, Watanabe Y, Nagata S, Kagamiyama H, et al. An ICAM-related neuronal glycoprotein, telencephalin, with brain segment-specific expression. Neuron. 1994;12:541–53.PubMedCrossRef
73.
Zurück zum Zitat Karin N. Reversal of experimental autoimmune encephalomyelitis by a soluble peptide variant of a myelin basic protein epitope: T cell receptor antagonism and reduction of interferon gamma and tumor necrosis factor alpha production. J Exp Med. 2004;180:2227–37.CrossRef Karin N. Reversal of experimental autoimmune encephalomyelitis by a soluble peptide variant of a myelin basic protein epitope: T cell receptor antagonism and reduction of interferon gamma and tumor necrosis factor alpha production. J Exp Med. 2004;180:2227–37.CrossRef
74.
Zurück zum Zitat Krogsgaard M, Wucherpfennig KW, Cannella B, Hansen BE, Svejgaard A, Pyrdol J, et al. Visualization of myelin basic protein (Mbp) T cell epitopes in multiple sclerosis lesions using a monoclonal antibody specific for the human histocompatibility leukocyte antigen (Hla)-Dr2–Mbp 85–99 complex. J Exp Med. 2000;191:1395–412.PubMedPubMedCentralCrossRef Krogsgaard M, Wucherpfennig KW, Cannella B, Hansen BE, Svejgaard A, Pyrdol J, et al. Visualization of myelin basic protein (Mbp) T cell epitopes in multiple sclerosis lesions using a monoclonal antibody specific for the human histocompatibility leukocyte antigen (Hla)-Dr2–Mbp 85–99 complex. J Exp Med. 2000;191:1395–412.PubMedPubMedCentralCrossRef
75.
Zurück zum Zitat Falcão AM, Meijer M, Scaglione A, Rinwa P, Agirre E, Liang J, et al. PADI2-mediated citrullination is required for efficient oligodendrocyte differentiation and myelination. bioRxiv. 2018. https://doi.org/10.1101/425348. Falcão AM, Meijer M, Scaglione A, Rinwa P, Agirre E, Liang J, et al. PADI2-mediated citrullination is required for efficient oligodendrocyte differentiation and myelination. bioRxiv. 2018. https://​doi.​org/​10.​1101/​425348.
76.
Zurück zum Zitat Pritzker LB, Joshi S, Harauz G, Moscarello MA. Deimination of myelin basic protein. 2. Effect of methylation of MBP on its deimination by peptidylarginine deiminase. Biochemistry. 2000;39:5382–8.PubMedCrossRef Pritzker LB, Joshi S, Harauz G, Moscarello MA. Deimination of myelin basic protein. 2. Effect of methylation of MBP on its deimination by peptidylarginine deiminase. Biochemistry. 2000;39:5382–8.PubMedCrossRef
77.
Zurück zum Zitat Wood DD, Bilbao JM, O’Connors P, Moscarello MA. Acute multiple sclerosis (Marburg type) is associated with developmentally immature myelin basic protein. Ann Neurol. 1996;40:18–24.PubMedCrossRef Wood DD, Bilbao JM, O’Connors P, Moscarello MA. Acute multiple sclerosis (Marburg type) is associated with developmentally immature myelin basic protein. Ann Neurol. 1996;40:18–24.PubMedCrossRef
78.
Zurück zum Zitat Watts C, Matthews SP, Mazzeo D, Manoury B, Moss CX. Asparaginyl endopeptidase: case history of a class II MHC compartment protease. Immunol Rev. 2005;207:218–28.PubMedCrossRef Watts C, Matthews SP, Mazzeo D, Manoury B, Moss CX. Asparaginyl endopeptidase: case history of a class II MHC compartment protease. Immunol Rev. 2005;207:218–28.PubMedCrossRef
79.
Zurück zum Zitat Manoury B, Mazzeo D, Fugger L, Viner N, Ponsford M, Streeter H, et al. Destructive processing by asparagine endopeptidase limits presentation of a dominant T cell epitope in MBP. Nat Immunol. 2002;3:169–74.PubMedCrossRef Manoury B, Mazzeo D, Fugger L, Viner N, Ponsford M, Streeter H, et al. Destructive processing by asparagine endopeptidase limits presentation of a dominant T cell epitope in MBP. Nat Immunol. 2002;3:169–74.PubMedCrossRef
81.
Zurück zum Zitat Franklin RJM. Why does remyelination fail in multiple sclerosis? Nat Rev Neurosci. 2002;3:705–14.PubMedCrossRef Franklin RJM. Why does remyelination fail in multiple sclerosis? Nat Rev Neurosci. 2002;3:705–14.PubMedCrossRef
82.
Zurück zum Zitat Kuhlmann T, Miron V, Cuo Q, Wegner C, Antel J, Brück W. Differentiation block of oligodendroglial progenitor cells as a cause for remyelination failure in chronic multiple sclerosis. Brain. 2008;131:1749–58.PubMedCrossRef Kuhlmann T, Miron V, Cuo Q, Wegner C, Antel J, Brück W. Differentiation block of oligodendroglial progenitor cells as a cause for remyelination failure in chronic multiple sclerosis. Brain. 2008;131:1749–58.PubMedCrossRef
83.
Zurück zum Zitat Redwine JM, Armstrong RC. In vivo proliferation of oligodendrocyte progenitors expressing PDGFαR during early remyelination. J Neurobiol. 1998;37:413–28.PubMedCrossRef Redwine JM, Armstrong RC. In vivo proliferation of oligodendrocyte progenitors expressing PDGFαR during early remyelination. J Neurobiol. 1998;37:413–28.PubMedCrossRef
84.
Zurück zum Zitat Hinks GL, Franklin RJM. Distinctive patterns of PDGF-A, FGF-2, IGF-I, and TGF-β1 gene expression during remyelination of experimentally-induced spinal cord demyelination. Mol Cell Neurosci. 1999;14:153–68.PubMedCrossRef Hinks GL, Franklin RJM. Distinctive patterns of PDGF-A, FGF-2, IGF-I, and TGF-β1 gene expression during remyelination of experimentally-induced spinal cord demyelination. Mol Cell Neurosci. 1999;14:153–68.PubMedCrossRef
85.
Zurück zum Zitat Messersmith DJ, Murtie JC, Le TQ, Frost EE, Armstrong RC. Fibroblast growth factor 2 (FGF2) and FGF receptor expression in an experimental demyelinating disease with extensive remyelination. J Neurosci Res. 2000;62:241–56.PubMedCrossRefPubMedCentral Messersmith DJ, Murtie JC, Le TQ, Frost EE, Armstrong RC. Fibroblast growth factor 2 (FGF2) and FGF receptor expression in an experimental demyelinating disease with extensive remyelination. J Neurosci Res. 2000;62:241–56.PubMedCrossRefPubMedCentral
86.
Zurück zum Zitat Fancy SPJ, Baranzini SE, Zhao C, Yuk D-I, Irvine K-A, Kaing S, et al. Dysregulation of the Wnt pathway inhibits timely myelination and remyelination in the mammalian CNS. Genes Dev. 2009;23:1571–85.PubMedPubMedCentralCrossRef Fancy SPJ, Baranzini SE, Zhao C, Yuk D-I, Irvine K-A, Kaing S, et al. Dysregulation of the Wnt pathway inhibits timely myelination and remyelination in the mammalian CNS. Genes Dev. 2009;23:1571–85.PubMedPubMedCentralCrossRef
88.
Zurück zum Zitat Yamamoto A, Nagano T, Takehara S, Hibi M, Aizawa S. Shisa promotes head formation through the inhibition of receptor protein maturation for the caudalizing factors, Wnt and FGF. Cell. 2005;120:223–35.PubMedCrossRef Yamamoto A, Nagano T, Takehara S, Hibi M, Aizawa S. Shisa promotes head formation through the inhibition of receptor protein maturation for the caudalizing factors, Wnt and FGF. Cell. 2005;120:223–35.PubMedCrossRef
89.
Zurück zum Zitat Chen Y, Wang H, Yoon SO, Xu X, Hottiger MO, Svaren J, et al. HDAC-mediated deacetylation of NF-κB is critical for Schwann cell myelination. Nat Neurosci. 2011;14:437–41.PubMedPubMedCentralCrossRef Chen Y, Wang H, Yoon SO, Xu X, Hottiger MO, Svaren J, et al. HDAC-mediated deacetylation of NF-κB is critical for Schwann cell myelination. Nat Neurosci. 2011;14:437–41.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Bhat MA, Rios JC, Lu Y, Garcia-Fresco GP, Ching W, Martin MS, et al. Axon-glia interactions and the domain organization of myelinated axons requires Neurexin IV/Caspr/Paranodin. Neuron. 2001;30:369–83.PubMedCrossRef Bhat MA, Rios JC, Lu Y, Garcia-Fresco GP, Ching W, Martin MS, et al. Axon-glia interactions and the domain organization of myelinated axons requires Neurexin IV/Caspr/Paranodin. Neuron. 2001;30:369–83.PubMedCrossRef
91.
Zurück zum Zitat Sherman DL, Tait S, Melrose S, Johnson R, Zonta B, Court FA, et al. Neurofascins are required to establish axonal domains for saltatory conduction. Neuron. 2005;48:737–42.PubMedCrossRef Sherman DL, Tait S, Melrose S, Johnson R, Zonta B, Court FA, et al. Neurofascins are required to establish axonal domains for saltatory conduction. Neuron. 2005;48:737–42.PubMedCrossRef
93.
Zurück zum Zitat Linington C, Engelhardt B, Kapocs G, Lassman H. Induction of persistently demyelinated lesions in the rat following the repeated adoptive transfer of encephalitogenic T cells and demyelinating antibody. J Neuroimmunol. 1992;40:219–24.PubMedCrossRef Linington C, Engelhardt B, Kapocs G, Lassman H. Induction of persistently demyelinated lesions in the rat following the repeated adoptive transfer of encephalitogenic T cells and demyelinating antibody. J Neuroimmunol. 1992;40:219–24.PubMedCrossRef
95.
Zurück zum Zitat Akama KT, Albanese C, Pestell RG, Van Eldik L. Amyloid beta-peptide stimulates nitric oxide production in astrocytes through an NFkappaB-dependent mechanism. Proc Natl Acad Sci U S A. 1998;95:5795–800. Akama KT, Albanese C, Pestell RG, Van Eldik L. Amyloid beta-peptide stimulates nitric oxide production in astrocytes through an NFkappaB-dependent mechanism. Proc Natl Acad Sci U S A. 1998;95:5795–800.
96.
Zurück zum Zitat Castillo-Fernandez JE, Spector TD, Bell JT. Epigenetics of discordant monozygotic twins: Implications for disease. Genome Med. 2014;6:1–16.CrossRef Castillo-Fernandez JE, Spector TD, Bell JT. Epigenetics of discordant monozygotic twins: Implications for disease. Genome Med. 2014;6:1–16.CrossRef
97.
Zurück zum Zitat Healy BC, Ali EN, Guttmann CRG, Chitnis T, Glanz BI, Buckle G, et al. Smoking and disease progression in multiple sclerosis. Arch Neurol. 2009;66:858–64.PubMedPubMedCentralCrossRef Healy BC, Ali EN, Guttmann CRG, Chitnis T, Glanz BI, Buckle G, et al. Smoking and disease progression in multiple sclerosis. Arch Neurol. 2009;66:858–64.PubMedPubMedCentralCrossRef
98.
Zurück zum Zitat Degelman ML, Herman KM. Smoking and multiple sclerosis: a systematic review and meta-analysis using the Bradford Hill criteria for causation. Mult Scler Relat Disord. 2017;17:207–16.PubMedCrossRef Degelman ML, Herman KM. Smoking and multiple sclerosis: a systematic review and meta-analysis using the Bradford Hill criteria for causation. Mult Scler Relat Disord. 2017;17:207–16.PubMedCrossRef
99.
Zurück zum Zitat Talhout R, Schulz T, Florek E, van Benthem J, Wester P, Opperhuizen A. Hazardous compounds in tobacco smoke. Int J Environ Res Public Health. 2011;8:613–28.PubMedPubMedCentralCrossRef Talhout R, Schulz T, Florek E, van Benthem J, Wester P, Opperhuizen A. Hazardous compounds in tobacco smoke. Int J Environ Res Public Health. 2011;8:613–28.PubMedPubMedCentralCrossRef
100.
Zurück zum Zitat Hawkins BT, Abbruscato TJ, Egleton RD, Brown RC, Huber JD, Campos CR, et al. Nicotine increases in vivo blood-brain barrier permeability and alters cerebral microvascular tight junction protein distribution. Brain Res. 2004;1027:48–58.PubMedCrossRef Hawkins BT, Abbruscato TJ, Egleton RD, Brown RC, Huber JD, Campos CR, et al. Nicotine increases in vivo blood-brain barrier permeability and alters cerebral microvascular tight junction protein distribution. Brain Res. 2004;1027:48–58.PubMedCrossRef
101.
Zurück zum Zitat Minagar A, Alexander JS. Blood-brain barrier disruption in multiple sclerosis. Multiple Sclerosis. 2003;9:540–9.PubMedCrossRef Minagar A, Alexander JS. Blood-brain barrier disruption in multiple sclerosis. Multiple Sclerosis. 2003;9:540–9.PubMedCrossRef
102.
Zurück zum Zitat Funata N, Song SY, Okeda R, Funata M, Higashino F. A study of experimental cyanide encephalopathy in the acute phase -physiological and neuropathological correlation. Acta Neuropathol. 1984;64:99–107.PubMedCrossRef Funata N, Song SY, Okeda R, Funata M, Higashino F. A study of experimental cyanide encephalopathy in the acute phase -physiological and neuropathological correlation. Acta Neuropathol. 1984;64:99–107.PubMedCrossRef
103.
Zurück zum Zitat Wang L, Hagemann TL, Kalwa H, Michel T, Messing A, Feany MB. Nitric oxide mediates glial-induced neurodegeneration in Alexander disease. Nat Commun. 2015;6:8966. Wang L, Hagemann TL, Kalwa H, Michel T, Messing A, Feany MB. Nitric oxide mediates glial-induced neurodegeneration in Alexander disease. Nat Commun. 2015;6:8966.
104.
Zurück zum Zitat Kitamura M, Kasai A. Cigarette smoke as a trigger for the dioxin receptor-mediated signaling pathway. Cancer Lett. 2007;252:184–94.PubMedCrossRef Kitamura M, Kasai A. Cigarette smoke as a trigger for the dioxin receptor-mediated signaling pathway. Cancer Lett. 2007;252:184–94.PubMedCrossRef
105.
Zurück zum Zitat Quintana FJ. Regulation of central nervous system autoimmunity by the aryl hydrocarbon receptor. Semin Immunopathol. 2013;35:627–35.PubMedCrossRef Quintana FJ. Regulation of central nervous system autoimmunity by the aryl hydrocarbon receptor. Semin Immunopathol. 2013;35:627–35.PubMedCrossRef
106.
Zurück zum Zitat Quintana FJ, Basso AS, Iglesias AH, Korn T, Farez MF, Bettelli E, et al. Control of Treg and TH17 cell differentiation by the aryl hydrocarbon receptor. Nature. 2008;453:65–71.PubMedCrossRef Quintana FJ, Basso AS, Iglesias AH, Korn T, Farez MF, Bettelli E, et al. Control of Treg and TH17 cell differentiation by the aryl hydrocarbon receptor. Nature. 2008;453:65–71.PubMedCrossRef
107.
Zurück zum Zitat Zeilinger S, Kühnel B, Klopp N, Baurecht H, Kleinschmidt A, Gieger C, et al. Tobacco smoking leads to extensive genome-wide changes in DNA methylation. PLoS One. 2013;8:e63812.PubMedPubMedCentralCrossRef Zeilinger S, Kühnel B, Klopp N, Baurecht H, Kleinschmidt A, Gieger C, et al. Tobacco smoking leads to extensive genome-wide changes in DNA methylation. PLoS One. 2013;8:e63812.PubMedPubMedCentralCrossRef
108.
Zurück zum Zitat Neavin DR, Liu D, Ray B, Weinshilboum RM. The Role of the Aryl Hydrocarbon Receptor (AHR) in Immune and Inflammatory Diseases. Int J Mol Sci. 2018;19(12):3851.PubMedCentralCrossRef Neavin DR, Liu D, Ray B, Weinshilboum RM. The Role of the Aryl Hydrocarbon Receptor (AHR) in Immune and Inflammatory Diseases. Int J Mol Sci. 2018;19(12):3851.PubMedCentralCrossRef
109.
Zurück zum Zitat Duarte JH, Di Meglio P, Hirota K, Ahlfors H, Stockinger B. Differential influences of the aryl hydrocarbon receptor on Th17 mediated responses in vitro and in vivo. PLoS One. 2013;8:e79819.PubMedPubMedCentralCrossRef Duarte JH, Di Meglio P, Hirota K, Ahlfors H, Stockinger B. Differential influences of the aryl hydrocarbon receptor on Th17 mediated responses in vitro and in vivo. PLoS One. 2013;8:e79819.PubMedPubMedCentralCrossRef
110.
Zurück zum Zitat Juricek L, Carcaud J, Pelhaitre A, Riday TT, Chevallier A, Lanzini J, et al. AhR-deficiency as a cause of demyelinating disease and inflammation. Sci Rep. 2017;7:9794.PubMedPubMedCentralCrossRef Juricek L, Carcaud J, Pelhaitre A, Riday TT, Chevallier A, Lanzini J, et al. AhR-deficiency as a cause of demyelinating disease and inflammation. Sci Rep. 2017;7:9794.PubMedPubMedCentralCrossRef
111.
Zurück zum Zitat Rothhammer V, Borucki DM, Garcia Sanchez MI, Mazzola MA, Hemond CC, Regev K, et al. Dynamic regulation of serum aryl hydrocarbon receptor agonists in MS. Neurol Neuroimmunol Neuroinflamm. 2017;4:e359.PubMedPubMedCentralCrossRef Rothhammer V, Borucki DM, Garcia Sanchez MI, Mazzola MA, Hemond CC, Regev K, et al. Dynamic regulation of serum aryl hydrocarbon receptor agonists in MS. Neurol Neuroimmunol Neuroinflamm. 2017;4:e359.PubMedPubMedCentralCrossRef
113.
Zurück zum Zitat Hatton OL, Harris-Arnold A, Schaffert S, Krams SM, Martinez OM. The interplay between Epstein-Barr virus and B lymphocytes: implications for infection, immunity, and disease. Immunol Res. 2014;58:268–76.PubMedPubMedCentralCrossRef Hatton OL, Harris-Arnold A, Schaffert S, Krams SM, Martinez OM. The interplay between Epstein-Barr virus and B lymphocytes: implications for infection, immunity, and disease. Immunol Res. 2014;58:268–76.PubMedPubMedCentralCrossRef
114.
Zurück zum Zitat Ascherio A, Munger KL, Lennette ET, Spiegelman D, Hernán MA, Olek MJ, et al. Epstein-Barr virus antibodies and risk of multiple sclerosis: a prospective study. J Am Med Assoc. 2001;286:3083–8.CrossRef Ascherio A, Munger KL, Lennette ET, Spiegelman D, Hernán MA, Olek MJ, et al. Epstein-Barr virus antibodies and risk of multiple sclerosis: a prospective study. J Am Med Assoc. 2001;286:3083–8.CrossRef
115.
Zurück zum Zitat Olsson T, Barcellos LF, Alfredsson L. Interactions between genetic, lifestyle and environmental risk factors for multiple sclerosis. Nat Rev Neurol. 2017;13:25–36.PubMedCrossRef Olsson T, Barcellos LF, Alfredsson L. Interactions between genetic, lifestyle and environmental risk factors for multiple sclerosis. Nat Rev Neurol. 2017;13:25–36.PubMedCrossRef
116.
Zurück zum Zitat Niller HH, Wolf H, Minarovits J. Regulation and dysregulation of Epstein - Barr virus latency: implications for the development of autoimmune diseases. Autoimmunity. 2008;41:298–328.PubMedCrossRef Niller HH, Wolf H, Minarovits J. Regulation and dysregulation of Epstein - Barr virus latency: implications for the development of autoimmune diseases. Autoimmunity. 2008;41:298–328.PubMedCrossRef
117.
Zurück zum Zitat Cepok S, Zhou D, Srivastava R, Nessler S, Stei S, Büssow K, et al. Identification of Epstein-Barr virus proteins as putative targets of the immune response in multiple sclerosis. J Clin Invest. 2005;115:1352-60.PubMedPubMedCentralCrossRef Cepok S, Zhou D, Srivastava R, Nessler S, Stei S, Büssow K, et al. Identification of Epstein-Barr virus proteins as putative targets of the immune response in multiple sclerosis. J Clin Invest. 2005;115:1352-60.PubMedPubMedCentralCrossRef
118.
Zurück zum Zitat Robertson KD, Hayward SD, Ling PD, Samid D, Robertson KD, Hayward SD, et al. Transcriptional activation of the Epstein-Barr virus latency C promoter after 5-azacytidine treatment: evidence that demethylation at a single CpG site is crucial. Mol Cell Biol. 1995;15:6150–9.PubMedPubMedCentralCrossRef Robertson KD, Hayward SD, Ling PD, Samid D, Robertson KD, Hayward SD, et al. Transcriptional activation of the Epstein-Barr virus latency C promoter after 5-azacytidine treatment: evidence that demethylation at a single CpG site is crucial. Mol Cell Biol. 1995;15:6150–9.PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Holick MF. Sunlight and vitamin D for bone health and prevention of autoimmune diseases, cancers, and cardiovascular disease... Vitamin D and Health in the 21st Century: proceedings of a conference held in Bethesda, MD, October 9-10, 2003. Am J Clin Nutr. 2004;80:1678S.PubMedCrossRef Holick MF. Sunlight and vitamin D for bone health and prevention of autoimmune diseases, cancers, and cardiovascular disease... Vitamin D and Health in the 21st Century: proceedings of a conference held in Bethesda, MD, October 9-10, 2003. Am J Clin Nutr. 2004;80:1678S.PubMedCrossRef
121.
Zurück zum Zitat Smolders J, Damoiseaux J, Menheere P, Hupperts R. Vitamin D as an immune modulator in multiple sclerosis, a review. J Neuroimmunol. 2008;194:7–17.PubMedCrossRef Smolders J, Damoiseaux J, Menheere P, Hupperts R. Vitamin D as an immune modulator in multiple sclerosis, a review. J Neuroimmunol. 2008;194:7–17.PubMedCrossRef
122.
Zurück zum Zitat Sloka S, Silva C, Pryse-Phillips W, Patten S, Metz L, Yong VW. A quantitative analysis of suspected environmental causes of MS. Can J Neurol Sci. 2011;38:98–105.CrossRef Sloka S, Silva C, Pryse-Phillips W, Patten S, Metz L, Yong VW. A quantitative analysis of suspected environmental causes of MS. Can J Neurol Sci. 2011;38:98–105.CrossRef
123.
Zurück zum Zitat Simpson S, Taylor B, Blizzard L, Ponsonby AL, Pittas F, Tremlett H, et al. Higher 25-hydroxyvitamin D is associated with lower relapse risk in multiple sclerosis. Ann Neurol. 2010;68:193–203. Simpson S, Taylor B, Blizzard L, Ponsonby AL, Pittas F, Tremlett H, et al. Higher 25-hydroxyvitamin D is associated with lower relapse risk in multiple sclerosis. Ann Neurol. 2010;68:193–203.
124.
Zurück zum Zitat Fetahu IS, Höbaus J, Kállay E. Vitamin D and the epigenome. Front Physiol. 2014;5:1–13. Fetahu IS, Höbaus J, Kállay E. Vitamin D and the epigenome. Front Physiol. 2014;5:1–13.
125.
Zurück zum Zitat Rawson JB, Sun Z, Dicks E, Daftary D, Parfrey PS, Green RC, et al. Vitamin D intake is negatively associated with promoter methylation of the Wnt antagonist gene DKK1 in a large group of colorectal cancer patients. Nutr Cancer. 2012;64:919–28.PubMedPubMedCentralCrossRef Rawson JB, Sun Z, Dicks E, Daftary D, Parfrey PS, Green RC, et al. Vitamin D intake is negatively associated with promoter methylation of the Wnt antagonist gene DKK1 in a large group of colorectal cancer patients. Nutr Cancer. 2012;64:919–28.PubMedPubMedCentralCrossRef
126.
Zurück zum Zitat Inestrosa NC, Varela-Nallar L. Wnt signalling in neuronal differentiation and development. Cell Tissue Res. 2015;359:215–23.PubMedCrossRef Inestrosa NC, Varela-Nallar L. Wnt signalling in neuronal differentiation and development. Cell Tissue Res. 2015;359:215–23.PubMedCrossRef
127.
Zurück zum Zitat Barragán-Martínez C, Speck-Hernández CA, Montoya-Ortiz G, Mantilla RD, Anaya JM, Rojas-Villarraga A, et al. Organic solvents as risk factor for autoimmune diseases: a systematic review and meta-analysis. PLoS One. 2012;7:e51506.PubMedPubMedCentralCrossRef Barragán-Martínez C, Speck-Hernández CA, Montoya-Ortiz G, Mantilla RD, Anaya JM, Rojas-Villarraga A, et al. Organic solvents as risk factor for autoimmune diseases: a systematic review and meta-analysis. PLoS One. 2012;7:e51506.PubMedPubMedCentralCrossRef
128.
Zurück zum Zitat Al-Hajri Z, Del Bigio MR. Brain damage in a large cohort of solvent abusers. Acta Neuropathol. 2010;119:435–45.PubMedCrossRef Al-Hajri Z, Del Bigio MR. Brain damage in a large cohort of solvent abusers. Acta Neuropathol. 2010;119:435–45.PubMedCrossRef
129.
Zurück zum Zitat Godderis L, De Raedt K, Tabish AM, Poels K, Maertens N, De Ruyck K, et al. Epigenetic changes in lymphocytes of solvent-exposed individuals. Epigenomics. 2012;4:269–77.PubMedCrossRef Godderis L, De Raedt K, Tabish AM, Poels K, Maertens N, De Ruyck K, et al. Epigenetic changes in lymphocytes of solvent-exposed individuals. Epigenomics. 2012;4:269–77.PubMedCrossRef
130.
Zurück zum Zitat Hong JY, Yu SY, Kim SY, Ahn JJ, Kim Y, Kim GW, et al. Association analysis of toluene exposure time with high-throughput mRNA expressions and methylation patterns using in vivo samples. Environ Res. 2016;146:59–64.PubMedCrossRef Hong JY, Yu SY, Kim SY, Ahn JJ, Kim Y, Kim GW, et al. Association analysis of toluene exposure time with high-throughput mRNA expressions and methylation patterns using in vivo samples. Environ Res. 2016;146:59–64.PubMedCrossRef
131.
Zurück zum Zitat Wu C, Keightley SY, Leung-Hagesteijn C, Radeva G, Coppolino M, Goicoechea S, et al. Integrin-linked protein kinase regulates fibronectin matrix assembly, E- cadherin expression, and tumorigenicity. J Biol Chem. 1998;273:528–36.PubMedCrossRef Wu C, Keightley SY, Leung-Hagesteijn C, Radeva G, Coppolino M, Goicoechea S, et al. Integrin-linked protein kinase regulates fibronectin matrix assembly, E- cadherin expression, and tumorigenicity. J Biol Chem. 1998;273:528–36.PubMedCrossRef
133.
Zurück zum Zitat Bagur MJ, Murcia MA, Jiménez-Monreal AM, Tur JA, Bibiloni MM, Alonso GL, et al. Influence of diet in multiple sclerosis: a systematic review. Adv Nutr An Int Rev J. 2017;8:463-47.CrossRef Bagur MJ, Murcia MA, Jiménez-Monreal AM, Tur JA, Bibiloni MM, Alonso GL, et al. Influence of diet in multiple sclerosis: a systematic review. Adv Nutr An Int Rev J. 2017;8:463-47.CrossRef
134.
Zurück zum Zitat Wang WW, Lu L, Bao TH, Zhang HM, Yuan J, Miao W, et al. Scutellarin alleviates behavioral deficits in a mouse model of multiple sclerosis, possibly through protecting neural stem cells. J Mol Neurosci. 2016;58:210–20.PubMedCrossRef Wang WW, Lu L, Bao TH, Zhang HM, Yuan J, Miao W, et al. Scutellarin alleviates behavioral deficits in a mouse model of multiple sclerosis, possibly through protecting neural stem cells. J Mol Neurosci. 2016;58:210–20.PubMedCrossRef
137.
Zurück zum Zitat Ross SA, Poirier L. Proceedings of the Trans-HHS Workshop: diet, DNA methylation processes and health. J Nutr. 2002;132:2329S-2332SPubMedCrossRef Ross SA, Poirier L. Proceedings of the Trans-HHS Workshop: diet, DNA methylation processes and health. J Nutr. 2002;132:2329S-2332SPubMedCrossRef
140.
Zurück zum Zitat Strickland FM, Hewagama A, Wu A, Sawalha AH, Delaney C, Hoeltzel MF, et al. Diet influences expression of autoimmune-associated genes and disease severity by epigenetic mechanisms in a transgenic mouse model of lupus. Arthritis Rheum. 2013;65:1872–81.CrossRef Strickland FM, Hewagama A, Wu A, Sawalha AH, Delaney C, Hoeltzel MF, et al. Diet influences expression of autoimmune-associated genes and disease severity by epigenetic mechanisms in a transgenic mouse model of lupus. Arthritis Rheum. 2013;65:1872–81.CrossRef
142.
Zurück zum Zitat Zhou X, Zhang L, Ji WJ, Yuan F, Guo ZZ, Pang B, et al. Variation in dietary salt intake induces coordinated dynamics of monocyte subsets and monocyteplatelet aggregates in humans: implications in end organ inflammation. PLoS One. 2013;8;e60332. Zhou X, Zhang L, Ji WJ, Yuan F, Guo ZZ, Pang B, et al. Variation in dietary salt intake induces coordinated dynamics of monocyte subsets and monocyteplatelet aggregates in humans: implications in end organ inflammation. PLoS One. 2013;8;e60332.
143.
Zurück zum Zitat Hernandez AL, Kleinewietfeld M, David A, Hernandez AL, Kitz A, Wu C, et al. Sodium chloride inhibits the suppressive function of FOXP3 + regulatory T cells. J Clin Invest. 2015;125:4212–22.PubMedPubMedCentralCrossRef Hernandez AL, Kleinewietfeld M, David A, Hernandez AL, Kitz A, Wu C, et al. Sodium chloride inhibits the suppressive function of FOXP3 + regulatory T cells. J Clin Invest. 2015;125:4212–22.PubMedPubMedCentralCrossRef
145.
146.
Zurück zum Zitat Farez MF, Fiol MP, Gaitán MI, Quintana FJ, Correale J. Sodium intake is associated with increased disease activity in multiple sclerosis. J Neurol Neurosurg Psychiatry. 2015;86:26–31.PubMedCrossRef Farez MF, Fiol MP, Gaitán MI, Quintana FJ, Correale J. Sodium intake is associated with increased disease activity in multiple sclerosis. J Neurol Neurosurg Psychiatry. 2015;86:26–31.PubMedCrossRef
147.
Zurück zum Zitat Mischke M, Plösch T. The gut microbiota and their metabolites: potential implications for the host epigenome. Adv Exp Med Biol. 2016;902:33–44.PubMedCrossRef Mischke M, Plösch T. The gut microbiota and their metabolites: potential implications for the host epigenome. Adv Exp Med Biol. 2016;902:33–44.PubMedCrossRef
151.
Zurück zum Zitat Berer K, Gerdes LA, Cekanaviciute E, Jia X, Xiao L, Xia Z, et al. Gut microbiota from multiple sclerosis patients enables spontaneous autoimmune encephalomyelitis in mice. Proc Natl Acad Sci. 2017;114:10719–24.PubMedCrossRefPubMedCentral Berer K, Gerdes LA, Cekanaviciute E, Jia X, Xiao L, Xia Z, et al. Gut microbiota from multiple sclerosis patients enables spontaneous autoimmune encephalomyelitis in mice. Proc Natl Acad Sci. 2017;114:10719–24.PubMedCrossRefPubMedCentral
152.
Zurück zum Zitat Sun CY, Chang SC, Wu MS. Suppression of Klotho expression by protein-bound uremic toxins is associated with increased DNA methyltransferase expression and DNA hypermethylation. Kidney Int. 2012;81:640–50.PubMedPubMedCentralCrossRef Sun CY, Chang SC, Wu MS. Suppression of Klotho expression by protein-bound uremic toxins is associated with increased DNA methyltransferase expression and DNA hypermethylation. Kidney Int. 2012;81:640–50.PubMedPubMedCentralCrossRef
153.
Zurück zum Zitat Torres PU, Prié D, Molina-Blétry V, Beck L, Silve C, Friedlander G. Klotho: An antiaging protein involved in mineral and vitamin D metabolism. Kidney Int. 2007;71:730–7.PubMedCrossRef Torres PU, Prié D, Molina-Blétry V, Beck L, Silve C, Friedlander G. Klotho: An antiaging protein involved in mineral and vitamin D metabolism. Kidney Int. 2007;71:730–7.PubMedCrossRef
154.
Zurück zum Zitat Afanas’ev I. New nucleophilic mechanisms of Ros-dependent epigenetic modifications: comparison of aging and cancer. Aging Dis. 2014;5:52–62.PubMedCrossRef Afanas’ev I. New nucleophilic mechanisms of Ros-dependent epigenetic modifications: comparison of aging and cancer. Aging Dis. 2014;5:52–62.PubMedCrossRef
155.
Zurück zum Zitat Anderson OS, Sant KE, Dolinoy DC. Nutrition and epigenetics: an interplay of dietary methyl donors, one-carbon metabolism and DNA methylation. J Nutr Biochem. 2012;23:853–9.PubMedPubMedCentralCrossRef Anderson OS, Sant KE, Dolinoy DC. Nutrition and epigenetics: an interplay of dietary methyl donors, one-carbon metabolism and DNA methylation. J Nutr Biochem. 2012;23:853–9.PubMedPubMedCentralCrossRef
156.
Zurück zum Zitat Kalhan SC. One-carbon metabolism, fetal growth and long-term consequences. In: Nestle Nutrition Institute Workshop Series; 2013. Kalhan SC. One-carbon metabolism, fetal growth and long-term consequences. In: Nestle Nutrition Institute Workshop Series; 2013.
157.
Zurück zum Zitat Ly A, Lee H, Chen J, Sie KKY, Renlund R, Medline A, et al. Effect of maternal and postweaning folic acid supplementation on mammary tumor risk in the offspring. Cancer Res. 2011;71:988–97.PubMedCrossRef Ly A, Lee H, Chen J, Sie KKY, Renlund R, Medline A, et al. Effect of maternal and postweaning folic acid supplementation on mammary tumor risk in the offspring. Cancer Res. 2011;71:988–97.PubMedCrossRef
158.
Zurück zum Zitat Dumas M-E, Barton RH, Toye A, Cloarec O, Blancher C, Rothwell A, et al. Metabolic profiling reveals a contribution of gut microbiota to fatty liver phenotype in insulin-resistant mice. Proc Natl Acad Sci. 2006;103:12511–6.CrossRef Dumas M-E, Barton RH, Toye A, Cloarec O, Blancher C, Rothwell A, et al. Metabolic profiling reveals a contribution of gut microbiota to fatty liver phenotype in insulin-resistant mice. Proc Natl Acad Sci. 2006;103:12511–6.CrossRef
159.
Zurück zum Zitat Schaible TD, Harris RA, Dowd SE, Smith CW, Kellermayer R. Maternal methyl-donor supplementation induces prolonged murine offspring colitis susceptibility in association with mucosal epigenetic and microbiomic changes. Hum Mol Genet. 2011;20:1687–96.PubMedPubMedCentralCrossRef Schaible TD, Harris RA, Dowd SE, Smith CW, Kellermayer R. Maternal methyl-donor supplementation induces prolonged murine offspring colitis susceptibility in association with mucosal epigenetic and microbiomic changes. Hum Mol Genet. 2011;20:1687–96.PubMedPubMedCentralCrossRef
160.
Zurück zum Zitat Canani RB, Costanzo MD, Leone L. The epigenetic effects of butyrate: potential therapeutic implications for clinical practice. Clin Epigenetics. 2012;4:1–7. Canani RB, Costanzo MD, Leone L. The epigenetic effects of butyrate: potential therapeutic implications for clinical practice. Clin Epigenetics. 2012;4:1–7.
161.
Zurück zum Zitat Kanai Y, Hirohashi S. Alterations of DNA methylation associated with abnormalities of DNA methyltransferases in human cancers during transition from a precancerous to a malignant state. Carcinogenesis. 2007;28:2434–42.PubMedCrossRef Kanai Y, Hirohashi S. Alterations of DNA methylation associated with abnormalities of DNA methyltransferases in human cancers during transition from a precancerous to a malignant state. Carcinogenesis. 2007;28:2434–42.PubMedCrossRef
162.
Zurück zum Zitat Worthley DL, Whitehall VLJ, Le Leu RK, Irahara N, Buttenshaw RL, Mallitt KA, et al. DNA methylation in the rectal mucosa is associated with crypt proliferation and fecal short-chain fatty acids. Dig Dis Sci. 2011;56:387–96.PubMedCrossRef Worthley DL, Whitehall VLJ, Le Leu RK, Irahara N, Buttenshaw RL, Mallitt KA, et al. DNA methylation in the rectal mucosa is associated with crypt proliferation and fecal short-chain fatty acids. Dig Dis Sci. 2011;56:387–96.PubMedCrossRef
164.
Zurück zum Zitat Wang L, Chang J, Varghese D, Dellinger M, Kumar S, Best AM, et al. A small molecule modulates Jumonji histone demethylase activity and selectively inhibits cancer growth. Nat Commun. 2013;4:2035. Wang L, Chang J, Varghese D, Dellinger M, Kumar S, Best AM, et al. A small molecule modulates Jumonji histone demethylase activity and selectively inhibits cancer growth. Nat Commun. 2013;4:2035.
165.
Zurück zum Zitat Minor EA, Court BL, Young JI, Wang G. Ascorbate induces ten-eleven translocation (Tet) methylcytosine dioxygenase-mediated generation of 5-hydroxymethylcytosine. J Biol Chem. 2013;288:13669–74.PubMedCrossRefPubMedCentral Minor EA, Court BL, Young JI, Wang G. Ascorbate induces ten-eleven translocation (Tet) methylcytosine dioxygenase-mediated generation of 5-hydroxymethylcytosine. J Biol Chem. 2013;288:13669–74.PubMedCrossRefPubMedCentral
166.
Zurück zum Zitat Kumar H, Lund R, Laiho A, Lundelin K, Ley RE, Isolauri E, et al. Gut microbiota as an epigenetic regulator: pilot study based on whole-genome methylation analysis. MBio. 2014;5:e02113–14. Kumar H, Lund R, Laiho A, Lundelin K, Ley RE, Isolauri E, et al. Gut microbiota as an epigenetic regulator: pilot study based on whole-genome methylation analysis. MBio. 2014;5:e02113–14.
168.
Zurück zum Zitat Radom-Aizik S, Zaldivar F, Leu SY, Adams GR, Oliver S, Cooper DM. Effects of exercise on microRNA expression in young males peripheral blood mononuclear cells. Clin Transl Sci. 2012;5:32–8.PubMedPubMedCentralCrossRef Radom-Aizik S, Zaldivar F, Leu SY, Adams GR, Oliver S, Cooper DM. Effects of exercise on microRNA expression in young males peripheral blood mononuclear cells. Clin Transl Sci. 2012;5:32–8.PubMedPubMedCentralCrossRef
169.
Zurück zum Zitat Gomez-Pinilla F, Zhuang Y, Feng J, Ying Z, Fan G. Exercise impacts brain-derived neurotrophic factor plasticity by engaging mechanisms of epigenetic regulation. Eur J Neurosci. 2011;33:383–90.PubMedCrossRef Gomez-Pinilla F, Zhuang Y, Feng J, Ying Z, Fan G. Exercise impacts brain-derived neurotrophic factor plasticity by engaging mechanisms of epigenetic regulation. Eur J Neurosci. 2011;33:383–90.PubMedCrossRef
171.
172.
Zurück zum Zitat Stadelmann C, Kerschensteiner M, Misgeld T, Brück W, Hohlfeld R, Lassmann H. BDNF and gp145trkB in multiple sclerosis brain lesions: neuroprotective interactions between immune and neuronal cells? Brain. 2002;125:75–85.PubMedCrossRef Stadelmann C, Kerschensteiner M, Misgeld T, Brück W, Hohlfeld R, Lassmann H. BDNF and gp145trkB in multiple sclerosis brain lesions: neuroprotective interactions between immune and neuronal cells? Brain. 2002;125:75–85.PubMedCrossRef
173.
Zurück zum Zitat Kerschensteiner M, Gallmeier E, Behrens L, Leal VV, Misgeld T, Klinkert WE, et al. Activated human T cells, B cells, and monocytes produce brainderived neurotrophic factor in vitro and in inflammatory brain lesions: a neuroprotective role of inflammation? J Exp Med. 1999;189:865–70.PubMedPubMedCentralCrossRef Kerschensteiner M, Gallmeier E, Behrens L, Leal VV, Misgeld T, Klinkert WE, et al. Activated human T cells, B cells, and monocytes produce brainderived neurotrophic factor in vitro and in inflammatory brain lesions: a neuroprotective role of inflammation? J Exp Med. 1999;189:865–70.PubMedPubMedCentralCrossRef
174.
Zurück zum Zitat Vaynman S, Ying Z, Gomez-Pinilla F. Hippocampal BDNF mediates the efficacy of exercise on synaptic plasticity and cognition. Eur J Neurosci. 2004;20:2580–90.PubMedCrossRef Vaynman S, Ying Z, Gomez-Pinilla F. Hippocampal BDNF mediates the efficacy of exercise on synaptic plasticity and cognition. Eur J Neurosci. 2004;20:2580–90.PubMedCrossRef
175.
Zurück zum Zitat Xiao J, Wong AW, Willingham MM, Van Den Buuse M, Kilpatrick TJ, Murray SS. Brain-derived neurotrophic factor promotes central nervous system myelination via a direct effect upon oligodendrocytes. NeuroSignals. 2011;18:186–202.PubMedCrossRef Xiao J, Wong AW, Willingham MM, Van Den Buuse M, Kilpatrick TJ, Murray SS. Brain-derived neurotrophic factor promotes central nervous system myelination via a direct effect upon oligodendrocytes. NeuroSignals. 2011;18:186–202.PubMedCrossRef
176.
Zurück zum Zitat Sølvsten CAE, De Paoli F, Christensen JH, Nielsen AL. Voluntary physical exercise induces expression and epigenetic remodeling of VegfA in the rat hippocampus. Mol Neurobiol. 2016;55:567–82.PubMedCrossRef Sølvsten CAE, De Paoli F, Christensen JH, Nielsen AL. Voluntary physical exercise induces expression and epigenetic remodeling of VegfA in the rat hippocampus. Mol Neurobiol. 2016;55:567–82.PubMedCrossRef
177.
Zurück zum Zitat Lin W. Neuroprotective effects of vascular endothelial growth factor A in the experimental autoimmune encephalomyelitis model of multiple sclerosis. Neural Regen Res. 2017;12:70–1.PubMedPubMedCentralCrossRef Lin W. Neuroprotective effects of vascular endothelial growth factor A in the experimental autoimmune encephalomyelitis model of multiple sclerosis. Neural Regen Res. 2017;12:70–1.PubMedPubMedCentralCrossRef
178.
Zurück zum Zitat Keane RW, Dietrich WD, de Rivero Vaccari JP. Inflammasome proteins as biomarkers of multiple sclerosis. Front Neurol. 2018;9:135. Keane RW, Dietrich WD, de Rivero Vaccari JP. Inflammasome proteins as biomarkers of multiple sclerosis. Front Neurol. 2018;9:135.
179.
Zurück zum Zitat Nakajima K, Takeoka M, Mori M, Hashimoto S, Sakurai A, Nose H, et al. Exercise effects on methylation of ASC gene. Int J Sports Med. 2010;31:671–5.PubMedCrossRef Nakajima K, Takeoka M, Mori M, Hashimoto S, Sakurai A, Nose H, et al. Exercise effects on methylation of ASC gene. Int J Sports Med. 2010;31:671–5.PubMedCrossRef
180.
Zurück zum Zitat Slattery ML, Curtin K, Sweeney C, Levin TR, Potter J, Wolff RK, et al. Diet and lifestyle factor associations with CpG island methylator phenotype and BRAF mutations in colon cancer. Int J Cancer. 2006;120:656–63.CrossRef Slattery ML, Curtin K, Sweeney C, Levin TR, Potter J, Wolff RK, et al. Diet and lifestyle factor associations with CpG island methylator phenotype and BRAF mutations in colon cancer. Int J Cancer. 2006;120:656–63.CrossRef
181.
Zurück zum Zitat Zhang FF, Cardarelli R, Carroll J, Zhang S, Fulda KG, Gonzalez K, et al. Physical activity and global genomic DNA methylation in a cancer-free population. Epigenetics. 2011;6:293–9.PubMedPubMedCentralCrossRef Zhang FF, Cardarelli R, Carroll J, Zhang S, Fulda KG, Gonzalez K, et al. Physical activity and global genomic DNA methylation in a cancer-free population. Epigenetics. 2011;6:293–9.PubMedPubMedCentralCrossRef
184.
Zurück zum Zitat Liu XJ, Ye HX, Li WP, Dai R, Chen D, Jin M. Relationship between psychosocial factors and onset of multiple sclerosis. Eur Neurol. 2009;62:130–6.PubMedPubMedCentralCrossRef Liu XJ, Ye HX, Li WP, Dai R, Chen D, Jin M. Relationship between psychosocial factors and onset of multiple sclerosis. Eur Neurol. 2009;62:130–6.PubMedPubMedCentralCrossRef
187.
Zurück zum Zitat Matysiak M, Makosa B, Walczak A, Selmaj K. Patients with multiple sclerosis resisted to glucocorticoid therapy: abnormal expression of heat-shock protein 90 in glucocorticoid receptor complex. Mult Scler. 2008;14:919–26.PubMedCrossRef Matysiak M, Makosa B, Walczak A, Selmaj K. Patients with multiple sclerosis resisted to glucocorticoid therapy: abnormal expression of heat-shock protein 90 in glucocorticoid receptor complex. Mult Scler. 2008;14:919–26.PubMedCrossRef
188.
Zurück zum Zitat Constantinescu CS, Farooqi N, O’Brien K, Gran B. Experimental autoimmune encephalomyelitis (EAE) as a model for multiple sclerosis (MS). Br J Pharmacol. 2011;164:1079–106.PubMedPubMedCentralCrossRef Constantinescu CS, Farooqi N, O’Brien K, Gran B. Experimental autoimmune encephalomyelitis (EAE) as a model for multiple sclerosis (MS). Br J Pharmacol. 2011;164:1079–106.PubMedPubMedCentralCrossRef
189.
Zurück zum Zitat Barthelmes J, Tafferner N, Kurz J, de Bruin N, Parnham MJ, Geisslinger G, et al. Induction of experimental autoimmune encephalomyelitis in mice and evaluation of the disease-dependent distribution of immune cells in various tissues. J Vis Exp. 2016:1–10. http://dx.doi.org/10.3791/53933. Barthelmes J, Tafferner N, Kurz J, de Bruin N, Parnham MJ, Geisslinger G, et al. Induction of experimental autoimmune encephalomyelitis in mice and evaluation of the disease-dependent distribution of immune cells in various tissues. J Vis Exp. 2016:1–10. http://​dx.​doi.​org/​10.​3791/​53933.
192.
Zurück zum Zitat Noori-Zadeh A, Mesbah-Namin SA, Saboor-Yaraghi AA. Epigenetic and gene expression alterations of FOXP3 in the T cells of EAE mouse model of multiple sclerosis. J Neurol Sci. 2017;375:203–8.PubMedCrossRef Noori-Zadeh A, Mesbah-Namin SA, Saboor-Yaraghi AA. Epigenetic and gene expression alterations of FOXP3 in the T cells of EAE mouse model of multiple sclerosis. J Neurol Sci. 2017;375:203–8.PubMedCrossRef
193.
Zurück zum Zitat Fontenot JD, Gavin MA, Rudensky AY. Foxp3 programs the development and function of CD4+CD25+ regulatory T cells. J Immunol. 2017;198:986–92. Fontenot JD, Gavin MA, Rudensky AY. Foxp3 programs the development and function of CD4+CD25+ regulatory T cells. J Immunol. 2017;198:986–92.
194.
Zurück zum Zitat Blakemore WF, Franklin RJM. Remyelination in experimental models of toxin-induced demyelination. Curr Top Microbiol Immunol. 2008;318:193–212. Blakemore WF, Franklin RJM. Remyelination in experimental models of toxin-induced demyelination. Curr Top Microbiol Immunol. 2008;318:193–212.
195.
Zurück zum Zitat Matsushima GK, Morell P. The neurotoxicant, cuprizone, as a model to study demyelination and remyelination in the central nervous system. Brain Pathol. 2001;11:107–16.CrossRefPubMedCentral Matsushima GK, Morell P. The neurotoxicant, cuprizone, as a model to study demyelination and remyelination in the central nervous system. Brain Pathol. 2001;11:107–16.CrossRefPubMedCentral
196.
Zurück zum Zitat Moyon S, Ma D, Huynh JL, Coutts DJC, Zhao C, Casaccia P, et al. Efficient remyelination requires DNA methylation. eneuro. 2017;4:ENEURO.0336–16.2017.PubMedPubMedCentralCrossRef Moyon S, Ma D, Huynh JL, Coutts DJC, Zhao C, Casaccia P, et al. Efficient remyelination requires DNA methylation. eneuro. 2017;4:ENEURO.0336–16.2017.PubMedPubMedCentralCrossRef
197.
Zurück zum Zitat Minor EA, Court BL, Young JI, Wang G. Ascorbate induces ten-eleven translocation (Tet) methylcytosine dioxygenase-mediated generation of 5-hydroxymethylcytosine. J Biol Chem. 2013;288:13669–74.PubMedPubMedCentralCrossRef Minor EA, Court BL, Young JI, Wang G. Ascorbate induces ten-eleven translocation (Tet) methylcytosine dioxygenase-mediated generation of 5-hydroxymethylcytosine. J Biol Chem. 2013;288:13669–74.PubMedPubMedCentralCrossRef
Metadaten
Titel
Aberrant DNA methylation profile exacerbates inflammation and neurodegeneration in multiple sclerosis patients
verfasst von
Naiara Celarain
Jordi Tomas-Roig
Publikationsdatum
01.12.2020
Verlag
BioMed Central
Erschienen in
Journal of Neuroinflammation / Ausgabe 1/2020
Elektronische ISSN: 1742-2094
DOI
https://doi.org/10.1186/s12974-019-1667-1

Weitere Artikel der Ausgabe 1/2020

Journal of Neuroinflammation 1/2020 Zur Ausgabe

Neu in den Fachgebieten Neurologie und Psychiatrie

Blutdrucksenkung schon im Rettungswagen bei akutem Schlaganfall?

31.05.2024 Apoplex Nachrichten

Der optimale Ansatz für die Blutdruckkontrolle bei Patientinnen und Patienten mit akutem Schlaganfall ist noch nicht gefunden. Ob sich eine frühzeitige Therapie der Hypertonie noch während des Transports in die Klinik lohnt, hat jetzt eine Studie aus China untersucht.

Nicht Creutzfeldt Jakob, sondern Abführtee-Vergiftung

29.05.2024 Hyponatriämie Nachrichten

Eine ältere Frau trinkt regelmäßig Sennesblättertee gegen ihre Verstopfung. Der scheint plötzlich gut zu wirken. Auf Durchfall und Erbrechen folgt allerdings eine Hyponatriämie. Nach deren Korrektur kommt es plötzlich zu progredienten Kognitions- und Verhaltensstörungen.

Schutz der Synapsen bei Alzheimer

29.05.2024 Morbus Alzheimer Nachrichten

Mit einem Neurotrophin-Rezeptor-Modulator lässt sich möglicherweise eine bestehende Alzheimerdemenz etwas abschwächen: Erste Phase-2-Daten deuten auf einen verbesserten Synapsenschutz.

Hörschwäche erhöht Demenzrisiko unabhängig von Beta-Amyloid

29.05.2024 Hörstörungen Nachrichten

Hört jemand im Alter schlecht, nimmt das Hirn- und Hippocampusvolumen besonders schnell ab, was auch mit einem beschleunigten kognitiven Abbau einhergeht. Und diese Prozesse scheinen sich unabhängig von der Amyloidablagerung zu ereignen.