25
REVIEW PAPER Genetic syndromes caused by mutations in epigenetic genes Marı ´a Berdasco Manel Esteller Received: 10 December 2012 / Accepted: 18 January 2013 / Published online: 31 January 2013 Ó Springer-Verlag Berlin Heidelberg 2013 Abstract The orchestrated organization of epigenetic factors that control chromatin dynamism, including DNA methylation, histone marks, non-coding RNAs (ncRNAs) and chromatin-remodeling proteins, is essential for the proper function of tissue homeostasis, cell identity and development. Indeed, deregulation of epigenetic profiles has been described in several human pathologies, including complex diseases (such as cancer, cardiovascular and neurological diseases), metabolic pathologies (type 2 dia- betes and obesity) and imprinting disorders. Over the last decade it has become increasingly clear that mutations of genes involved in epigenetic mechanism, such as DNA methyltransferases, methyl-binding domain proteins, his- tone deacetylases, histone methylases and members of the SWI/SNF family of chromatin remodelers are linked to human disorders, including Immunodeficiency Centro- meric instability Facial syndrome 1, Rett syndrome, Rubinstein–Taybi syndrome, Sotos syndrome or alpha- thalassemia/mental retardation X-linked syndrome, among others. As new members of the epigenetic machinery are described, the number of human syndromes associated with epigenetic alterations increases. As recent examples, muta- tions of histone demethylases and members of the non- coding RNA machinery have recently been associated with Kabuki syndrome, Claes-Jensen X-linked mental retardation syndrome and Goiter syndrome. In this review, we describe the variety of germline mutations of epigenetic modifiers that are known to be associated with human disorders, and dis- cuss the therapeutic potential of epigenetic drugs as pallia- tive care strategies in the treatment of such disorders. Introduction Chromatin dynamism is critical to basic cellular processes such as gene transcription, DNA replication, DNA recombination and DNA repair. DNA accessibility is modulated by epigenetic mechanisms that ultimately alter the structure of the chromatin and provide binding sites for a wide variety of regulatory proteins. The orchestrated organization of epigenetic factors, including DNA meth- ylation, histone marks, non-coding RNAs (ncRNAs), and their associated chromatin proteins, is essential for development and cellular differentiation. For instance, extensive chromatin remodeling occurs on a global level during early development. DNA methylation patterns undergo genome-wide alterations that occur immediately after fertilization and during early-preimplantation devel- opment, together with histone modification changes, such as increased H3K9me with differentiation (Reik 2007). Epigenetic factors also guarantee the activation and maintenance of specific differentiation programs in adult somatic cells (Berdasco and Esteller 2010). The active role of epigenetic factors in controlling cellular differen- tiation is supported by spontaneous cell differentiation after treatment with demethylating agents or histone M. Berdasco M. Esteller (&) Cancer Epigenetics Group, Cancer Epigenetics and Biology Program (PEBC), Bellvitge Biomedical Research Institute (IDIBELL), 3rd Floor, Hospital Duran i Reynals, Av. Gran Via 199-203, 08908 L’Hospitalet de LLobregat Barcelona, Catalonia, Spain e-mail: [email protected] M. Esteller Department of Physiological Sciences II, School of Medicine, University of Barcelona, Barcelona, Catalonia, Spain M. Esteller Institucio ´ Catalana de Recerca I Estudis Avanc ¸ats (ICREA), 08010 Barcelona, Catalonia, Spain 123 Hum Genet (2013) 132:359–383 DOI 10.1007/s00439-013-1271-x

Genetic syndromes caused by mutations in epigenetic genes · genetic syndromes featuring mutations in the DNA-related machinery that often cause neurological disorders are dis-cussed

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

REVIEW PAPER

Genetic syndromes caused by mutations in epigenetic genes

Marıa Berdasco • Manel Esteller

Received: 10 December 2012 / Accepted: 18 January 2013 / Published online: 31 January 2013

� Springer-Verlag Berlin Heidelberg 2013

Abstract The orchestrated organization of epigenetic

factors that control chromatin dynamism, including DNA

methylation, histone marks, non-coding RNAs (ncRNAs)

and chromatin-remodeling proteins, is essential for the

proper function of tissue homeostasis, cell identity and

development. Indeed, deregulation of epigenetic profiles

has been described in several human pathologies, including

complex diseases (such as cancer, cardiovascular and

neurological diseases), metabolic pathologies (type 2 dia-

betes and obesity) and imprinting disorders. Over the last

decade it has become increasingly clear that mutations of

genes involved in epigenetic mechanism, such as DNA

methyltransferases, methyl-binding domain proteins, his-

tone deacetylases, histone methylases and members of the

SWI/SNF family of chromatin remodelers are linked to

human disorders, including Immunodeficiency Centro-

meric instability Facial syndrome 1, Rett syndrome,

Rubinstein–Taybi syndrome, Sotos syndrome or alpha-

thalassemia/mental retardation X-linked syndrome, among

others. As new members of the epigenetic machinery are

described, the number of human syndromes associated with

epigenetic alterations increases. As recent examples, muta-

tions of histone demethylases and members of the non-

coding RNA machinery have recently been associated with

Kabuki syndrome, Claes-Jensen X-linked mental retardation

syndrome and Goiter syndrome. In this review, we describe

the variety of germline mutations of epigenetic modifiers that

are known to be associated with human disorders, and dis-

cuss the therapeutic potential of epigenetic drugs as pallia-

tive care strategies in the treatment of such disorders.

Introduction

Chromatin dynamism is critical to basic cellular processes

such as gene transcription, DNA replication, DNA

recombination and DNA repair. DNA accessibility is

modulated by epigenetic mechanisms that ultimately alter

the structure of the chromatin and provide binding sites

for a wide variety of regulatory proteins. The orchestrated

organization of epigenetic factors, including DNA meth-

ylation, histone marks, non-coding RNAs (ncRNAs), and

their associated chromatin proteins, is essential for

development and cellular differentiation. For instance,

extensive chromatin remodeling occurs on a global level

during early development. DNA methylation patterns

undergo genome-wide alterations that occur immediately

after fertilization and during early-preimplantation devel-

opment, together with histone modification changes, such

as increased H3K9me with differentiation (Reik 2007).

Epigenetic factors also guarantee the activation and

maintenance of specific differentiation programs in adult

somatic cells (Berdasco and Esteller 2010). The active

role of epigenetic factors in controlling cellular differen-

tiation is supported by spontaneous cell differentiation

after treatment with demethylating agents or histone

M. Berdasco � M. Esteller (&)

Cancer Epigenetics Group, Cancer Epigenetics and Biology

Program (PEBC), Bellvitge Biomedical Research Institute

(IDIBELL), 3rd Floor, Hospital Duran i Reynals, Av. Gran

Via 199-203, 08908 L’Hospitalet de LLobregat Barcelona,

Catalonia, Spain

e-mail: [email protected]

M. Esteller

Department of Physiological Sciences II, School of Medicine,

University of Barcelona, Barcelona, Catalonia, Spain

M. Esteller

Institucio Catalana de Recerca I Estudis Avancats (ICREA),

08010 Barcelona, Catalonia, Spain

123

Hum Genet (2013) 132:359–383

DOI 10.1007/s00439-013-1271-x

deacetylase inhibitors (reviewed in Berdasco and Esteller

2011). Treatment with the demethylation agent 5-aza-20-deoxycytidine promotes differentiation of different types

of adult stem cells into cardiac myogenic or osteogenic

cells by enhancing the expression of lineage genes. In a

similar manner, histone deacetylase inhibitor trichostatin

enhances chondrogenic or neural differentiation of stem

cells, reinforcing the epigenetic control of differentiation.

Furthermore, the essential role of these factors is reflected

in the fact that altered profiles of epigenetic marks often

lead to defaults in cellular homeostasis and development

of human diseases. Genetic alterations could explain the

causes of several monogenic diseases. However, the

genetic basis underlying the origin of complex and mul-

tifactorial diseases remains largely unknown and the

importance of the role of non-genetic mechanisms,

including epigenetic mechanisms or posttranslational

protein modifications, is increasingly being realized.

Cancer has been the best characterized complex human

disease associated with epigenetic defects (Berdasco and

Esteller 2010), but the list of complex diseases carrying

epigenetic defaults has been increasing rapidly in recent

years. Epigenetic studies have now been made of complex

diseases such as obesity, type 2 diabetes mellitus, car-

diovascular diseases and neurological disorders. These

pathogenic mechanisms are particularly interesting

because the epigenetic effects may also be affected by

aspects of the environment such as diet and lifestyle,

raising the possibility of ‘‘resetting’’ the altered epigenetic

marks. Deleterious epigenetic profiles could be a conse-

quence of mutations in the ‘‘writers’’, that is to say,

dysfunctional enzymes that are responsible for putting in

and out the epigenetic marks. Defective epigenetic

machinery has been observed in cancer initiation and

progression. Furthermore, germline mutations of epige-

netic modifiers contribute to the development of human

diseases including intellectual disability (review in: Fro-

yen et al. 2006; Kramer and van Bokhoven 2009;

Franklin and Mansuy 2011). The aim of the present

review is to provide an overview of these disorders

grouped by the type of epigenetic change involved:

(i) alterations in DNA methylation players; (ii) mutations

in histone modifiers; (iii) disruption of chromatin-remod-

eling complexes, and (iv) mutations in non-coding RNA

processing machinery.

Genetic disorders linked to DNA methylation defects

DNA methylation, or the addition of a methyl group to a

cytosine, is a key epigenetic player that has long been

considered the genome’s fifth base (Portela and

Esteller 2010). In mammals this reaction occurs at CpG

(deoxycytidine-phosphate-deoxyguanosine) sites located

throughout the genome, but there are certain areas, known

as CpG islands, that are enriched in CpG dinucleotides

(especially in promoter regions). Non-CpG islands of the

human genome are usually methylated and prevent geno-

mic instability phenomena, such as the movement of

transposable elements (Berdasco and Esteller 2010). Nor-

mal methylation at these sequences is also necessary for

X-chromosome inactivation in females and genomic

imprinting. Conversely, CpG islands are usually unme-

thylated, being closely related to the expression of house-

keeping genes. It is estimated that only 6 % of the human

CpG islands are methylated and, consequently, silenced,

being essential for maintaining tissue-specific patterns

during development and differentiation. By our current

understanding, this ‘‘DNA methylation code’’ seems to be

an oversimplification, since, in recent years, new genomic

contexts outside of CpG islands, known as CpG shores,

have emerged as candidates for regulating gene expression

of tissue-specific genes. The technological advance in

studying DNA methylation will provide insight into the

role of 5-methylcytosine patterns with respect to their

density, location and function, amongst other features.

Additionally the recently discovered cytosine modification

5-hydroxymethyl-20-deoxycytidine (5hmC) needs to be

further studied to determine its implications for normal and

diseased epigenetic regulation.

The enzymes responsible for introducing the methyl

group into a cytosine are DNA methyltransferases

(DNMTs). Three major proteins with DNMT activity have

been identified in mammals: DNMT1, DNMT3A and

DNMT3B. DNMT1 is a widely expressed maintenance

DNMT that recognizes hemimethylated DNA and is

responsible for maintaining the existing methylation pat-

terns after DNA replication. By contrast, DNMT3 enzymes

are de novo DNMTs that introduce methyl groups into

previously unmethylated cytosines. These enzymes intro-

duce a methyl group into the genome, but this ‘‘writing’’

must be interpreted (read) by the rest of the cellular

machinery (i.e., transcription factors, DNA polymerases,

chromatin-remodeling proteins, epigenetic enzymes, etc.).

Additional members of the DNMT family without meth-

yltransferase activity have been reported, such as DNMT2

or DNMT3L. DNMT3L lacks the amino acid sequence

necessary for methyltransferase but it seems to be required

for the establishment of maternal genomic imprints

(Aapola et al. 2002).

Methyl-CpG-binding domain (MBD) proteins are one of

the DNA methylation-associated proteins that could be

recruited to methylated DNA and in turn facilitate the

recruitment of histone modifiers and chromatin-remodeling

complexes (Portela and Esteller 2010). Evidence is

mounting of the role of DNA methylation in modulating

360 Hum Genet (2013) 132:359–383

123

cognitive functions of the central nervous system, such as

learning and memory, and of how dysregulation of DNMTs

activities can give rise to neurological disorders (Liu et al.

2009; Urdinguio et al. 2009; Feng et al. 2010). Some of the

genetic syndromes featuring mutations in the DNA-related

machinery that often cause neurological disorders are dis-

cussed in this section (Table 1).

DNMT1 mutations and disorders of the central

and peripheral nervous system

Hereditary sensory and autonomic neuropathy type 1 with

dementia and hearing loss (HSAN1; MIM #614116) is a

degenerative disorder of the central and peripheral nervous

system. Its clinical manifestations consist of sensory

impairment, sudomotor dysfunction (loss of sweating),

dementia and sensorineural hearing loss. HSAN1 is

inherited in an autosomal dominant manner, although the

proportion of patients with de novo mutations is unknown,

DNMT1 being the only gene in which mutations of exons

20 and 21 are known to cause HSAN1 (Klein et al. 2011;

Winkelmann et al. 2012). Molecular genetic testing to

screen for three mutations in exon 21 of DNMT1

(p.Ala570Val, p.Gly605Ala and p.Val606phe) is available

for research purposes. Mutations are present within the

targeting-sequence domain of DNMT1 that regulates the

binding of the enzyme to chromatin during the S-phase and

is responsible for maintaining this association during the

G2-M phases (Fatemi et al. 2001; Song et al. 2012).

DNMT1 is strongly expressed in postmitotic neurons and

plays important roles in neuronal differentiation, migration

and central neural connection (Feng et al. 2010). The

functional involvement of DNMT1 mutations has been

assessed in in vitro studies in HeLa cells (Klein et al.

2011), so that cells carrying mutations in the DNMT1

targeting sequence showed abnormal heterochromatin

binding of DNMT1 during the G2 phase and were pre-

maturely degraded. Abolishing DNMT1 function affects

DNA methylation cellular levels: first, a lower level of

global methylation (8 %) has been measured in mutant

cells (i.e., satellite 2 methylation was reduced); second,

site-specific hypermethylation at specific loci has also been

found (Klein et al. 2011). Additionally, DNMT1 is required

for CD4? differentiation into T regulatory cells (Jose-

fowicz et al. 2009), and a link between the absence of

CD4? T regulatory cells and the autoimmune response in

neurological syndromes has been proposed (Winkelmann

et al. 2012). The findings from several studies together

suggest that DNMT1 participates in a precise mechanism

of dynamic regulation of neuronal survival, but additional

efforts will be needed to elucidate its pathogenic mecha-

nisms and to explain the phenotypic variation observed

between individuals bearing different mutations.

DNMT3 mutations and immunodeficiency centromeric

instability facial syndrome 1

Immunodeficiency centromeric instability facial syndrome

1 (ICF1, MIM #242860) is a rare autosomal recessive

disorder characterized by immune defects in association

with centromere instability and facial anomalies. Several

chromosomal abnormalities have been described, including

the juxtacentromeric heterochromatin formation of chro-

mosomes 1, 9 and 16, an increased frequency of somatic

recombination between the arms of these chromosomes,

and a marked tendency to form multibranched configura-

tions (Ehrlich 2003). 60 % of ICF1 patients carry muta-

tions in the de novo DNA methyltransferase DNMT3B

(Xu et al. 1999; Lana et al. 2012). Mutations in the ZBTB24

gene, which encodes a transcription factor, are responsible

for the ICF type 2 phenotype (de Greef et al. 2011).

Hypomethylation in ICF patients commonly affect specific

non-coding repetitive sequences (satellites 2 and 3, subt-

elomeric sequences and Alu sequences), imprinted genes

and genes located in constitutive and facultative hetero-

chromatin (Xu et al. 1999; Yehezkel et al. 2008; Brun et al.

2011), causing chromatin decondensation and chromosome

instability. Recently, whole-genome bisulfite sequencing of

an ICF patient harboring mutated DNMT3B and one heal-

thy control have been performed to assess DNA methyla-

tion at base pair resolution (Heyn et al. 2012). The authors

concluded that ICF patients have 42 % less global DNA

methylation, especially in inactive heterochromatic regions

(in accordance with previous studies). Interestingly, the

methylation status of transcriptional active loci and rRNA

repeats did not change, suggesting that there is a selective

pressure to maintain the stability of these genomic struc-

tures (Heyn et al. 2012). In addition to methylation studies,

the altered expression of more than 700 genes in ICF1

patients has been described, especially genes related to

immune function, development and neurogenesis (Jin et al.

2008). Interestingly, half the upregulated genes were hy-

pomethylated (compared with normal cells) in parallel with

the loss of the histone repressive H3K27 trimethylation

mark and the gain of the histone active marks H3K9

acetylation and H3K4 trimethylation (Jin et al. 2008). Not

only the protein-coding genes are altered; a dramatic loss

of methylation (from 80 to 30 %) was found in hetero-

chromatic genes, which are usually aberrantly hypome-

thylated in cancer cells, although the hypomethylation was

not always associated with their activation (Brun et al.

2011). In contrast to the dysregulation of protein-coding

genes, no changes in histone marks associated with het-

erochromatic genes could be found (Brun et al. 2011).

Finally, the genomic instability generated in ICF patients

also resulted in replication defects, including shortening of

the S-phase, a higher global replication fork speed and

Hum Genet (2013) 132:359–383 361

123

Table 1 List of human disorders associated with germline mutations in epigenetic genes

Gene GeneIDa Cytogenetic

location

Function Disease OMIMb References

Mutations associated with DNA methylation

DNMT1 1786 19p13.2 DNMT Hereditary sensory and autonomic neuropathy type 1

(HSAN1)

614116 Klein et al. (2011),

Winkelmann et al.

(2012)

DNMT3b 1789 20q11.21 DNMT Immunodeficiency–centromeric instability–facial

anomalies syndrome 1 (ICF1)

242860 Xu et al. (1999);

Lana et al. (2012)

MECP2 4204 Xq28 MDB Rett syndrome 312750 Amir et al. (1999),

Moretti and Zoghbi

(2006)

MECP2 4204 Xq28 MDB Angelman syndrome 105830 Watson et al. (2001)

Mutations associated with histone modifications

MYST4 23522 10q22.2 HAT Genitopatellar syndrome 606170 Campeau et al.

(2012a), Simpson

et al. (2012)

MYST4 23522 10q22.2 HAT Say-Barber-Biesecker-Young-Simpson syndrome

(SBBYS)

603736 Clayton-Smith et al.

(2011)

CREBBP 1387 16p13.3 HAT Rubinstein–Taybi syndrome 1 180849 Petrij et al. (1995),

Tsai et al. (2011)

EP300 2033 22q13.2 HAT Rubinstein–Taybi syndrome 2 613684 Arany et al. (1994),

Bartsch et al.

(2010)

HDAC4 14063 2q37.3 HDAC Brachydactyly-mental retardation syndrome

(BDMR)

600430 Williams et al.

(2010), Morris

et al. (2012)

EHMT1 79813 9q34.3 HMT Kleefstra syndrome 610253 Kleefstra et al.

(2009)

EZH2 2146 7q36.1 HMT Weaver syndrome 2 (WVS2) 614421 Gibso et al. (2012)

MLL2 8085 12q13.12 HMT Kabuki syndrome 1 147920 Ng et al. (2010),

Hannibal et al.

(2011)

NSD1 64324 5q35.2 HDMT Sotos syndrome 117550 Tatton-Brown and

Rahman (2007)

NSD1 64324 5q35.2 HDMT Weaver syndrome 1 (WVS1) 277590 Douglas et al. (2003)

NSD1 64324 5q35.2 HDMT Beckwith–Wiedemann syndrome 130650 Baujat et al. (2004)

JMJD3 7403 Xp11.3 HDMT Kabuki syndrome 2 300867 Lederer et al. (2012)

PHF8 23133 Xp11.22 HDMT Siderius X-Linked Mental Retardation Syndrome

(MRXSSD)

300263 Laumonnier et al.

(2005), Koivisto

et al. (2007)

JARID1C 8242 Xp11.22 HDMT Claes-Jensen X-linked Mental retardation syndrome 300534 Claes et al. (2000),

Jensen et al. (2005)

Mutations associated with chromatin- remodeling factors

ATXN7 6314 3p14.1 STAGA-

HAT

complex

Spinocerebellar ataxia 7 164500 Garden and La

Spada (2008)

ATRX 546 Xq21.1 SWI/SNF

complex

Alpha-thalassemia X-linked mental retardation

syndrome

301040 Gibbons et al. (1995)

ATRX 546 Xq21.1 SWI/SNF

complex

Mental retardation-hypotonic facies syndrome,

X-linked

309580 Abidi et al. (2005)

ATRX 546 Xq21.1 SWI/SNF

complex

Alpha-thalassemia myelodysplasia syndrome 300448 Gibbons et al. (2003)

ERCC6 2074 10q11.23 SWI/SNF

complex

Cockayne syndrome, type B (CSB) 133540 Laugel et al. (2010)

ERCC8 1161 5q12.1 SWI/SNF

complex

Cockayne syndrome, type A (CSA) 216400 Henning et al. (1995)

362 Hum Genet (2013) 132:359–383

123

earlier replication of heterochromatic genes in S-phase

(Lana et al. 2012). To conclude, the loss of DNMT3B

function and the interaction of DNMT3B with histone

modifications, together with the variable clinical features

of the patients, make ICF samples an ideal model for

investigating the epigenetic network and their molecular

consequences in several biological pathways (gene tran-

scription, DNA replication and recombination, among

others).

MeCP2 genetic alterations and Rett syndrome

Rett syndrome (MIM #312750) is a progressive neurode-

velopmental disorder characterized by arrested develop-

ment between 6 and 18 months of age, regression of

acquired skills, loss of speech, unusual stereotyped move-

ments and intellectual disability (Zachariah and Rastegar

2012). It affects predominantly females, occurring at a

frequency of 1:10,000 live births, but male patients with

Rett syndrome and variable phenotype (i.e., severe to

moderate congenital encephalopathy, infantile death or

psychiatric manifestations) have also been described (Ravn

et al. 2003; Moretti and Zoghbi 2006). Mutations in the

X-linked gene encoding methyl-CpG-binding protein 2

(MeCP2), including missense, frameshift and nonsense

mutations and intragenic deletions, account for the condi-

tion in up to 96 % of Rett syndrome patients (Amir et al.

1999; Moretti and Zoghbi 2006). Interestingly, the increase

in MeCP2 dosage due to duplications of the locus and

surrounding areas also causes the neurological disorder

Lubs X-linked mental retardation syndrome (MRXSL,

MIM #300260) (Van Esch et al. 2005). MeCP2 mutations

in Rett syndrome patients arise in the germline although the

phenotypic alterations in the neurological system appear at

Table 1 continued

Gene GeneIDa Cytogenetic

location

Function Disease OMIMb References

SMARCB1 6598 22q11.23 SWI/SNF

complex

Coffin-Siris syndrome 135900 Tsurusaki et al.

(2012)

SMARCA4 6597 19p13.2 SWI/SNF

complex

Coffin-Siris syndrome 135900 Tsurusaki et al.

(2012)

SMARCA2 6596 9p24.3 SWI/SNF

complex

Coffin-Siris syndrome 135900 Tsurusaki et al.

(2012)

SMARCA2 6596 9p24.3 SWI/SNF

complex

Nicolaides-Baraitser syndrome 601358 Van Houdt et al.

(2012)

ARID1A 8289 1p36.11 SWI/SNF

complex

Coffin-Siris syndrome 135900 Tsurusaki et al.

(2012)

ARID1B 57492 6q25.3 SWI/SNF

complex

Coffin-Siris syndrome 135900 Tsurusaki et al.

(2012)

SRCAP 10847 16p11.2 SRCAP

complex

Floating-Harbor syndrome 136140 Hood et al.

(2012)

CDH7 55636 8q12.1 CHD

complex

Coloboma of the eye, Heart Anomaly, choanal

Atresia, Retardation, Genital And Ear Anomalies

syndrome (CHARGE)

214800 Sanlaville et al.

(2006)

CDH7 55636 8q12.1 CHD

complex

Hypogonadotropic hypogonadism 2 with or without

anosmia

147950 Kim et al.

(2008)

Mutations associated with non-coding RNAs

DICER1 23405 14q32.13 miRNA

processing

Goiter, multinodular 1, with or without Sertoli-

Leydig cell tumors

138800 Rio Frio et al.

(2011)

TDP-43 23435 1p36.22 miRNA

processing

Amyotrophic lateral sclerosis 612069 Ling et al.

(2010)

FMRP 2332 Xq27.3 miRNA

processing

Fragile X syndrome 300624 Edbauer et al.

(2010)

DGCR8 54487 22q11.21 miRNA

processing

DiGeorge syndrome 188400 Shiohama et al.

(2003), Stark et al.

(2008)

CHD chromodomain helicase DNA-binding protein, DNMT DNA methyltransferase, HAT histone acetyltransferase, HDAC histone deacetylase,

HDMT histone demethylase, MDB methyl DNA-binding domain protein, HMT histone methyltransferase, SRCAP Snf2-related CREBBP

activator protein, STAGA SPT3/TAF9/GCN5 transcription coactivator complex, SWI/SNF SWItch/sucrose non fermentablea Identification in Entrez Gene database (http://www.ncbi.nlm.nih.gov/gene)b Identification in Online Mendelian Inheritance in Man (http://www.ncbi.nlm.nih.gov/omim/)

Hum Genet (2013) 132:359–383 363

123

early postnatal stages (Zachariah and Rastegar 2012).

Recently, abrogation of MeCP2 function in adult mice has

been found to result in severe neurological symptoms

commonly observed in Rett syndrome, such as global

shrinkage of the brain, increased neuronal cell density,

retraction of dendritic arbors, reduction of synaptic proteins

and altered astrocytic development, among others (Nguyen

et al. 2012). This is further evidence of the involvement of

MeCP2 in regression from a normal mature brain to a Rett-

like brain. This dynamism is of vital importance and sug-

gests opportunities for reverting the Rett phenotype. In this

regard, Rett syndrome is not a neurodegenerative disorder,

neurons do not die, opening the opportunities of phenotypic

reversion by means of MeCP2 restoration (Giacometti et al.

2007; Guy et al. 2007; Tropea et al. 2009).

MeCP2 protein is widely expressed in several tissues but

the highest level of expression has been observed in the

brain (Zachariah and Rastegar 2012). Although a higher

level of MeCP2 expression has been described in neurons,

especially in postmitotic neurons, deregulation of MeCP2

expression in glia cells also contributes to the progression

of Rett syndrome (Ballas et al. 2009). MeCP2 was initially

identified as a transcriptional repressor that binds to

methylated CpG dinucleotides and recruits corepressors

such as mSin3 and HDACs (Jones et al. 1998). However, in

recent years, studies have concluded that MeCP2 may act

either as a repressor or an activator, depending on its

interaction proteins (Chahrour et al. 2008; Yasui et al.

2007). Chahrour et al. (2008) examined the gene expres-

sion profiles in the hypothalamus of mice lacking or

overexpressing MeCP2 and, contrary to expectation, con-

firmed that MeCP2 occupancy preferentially occurs in

active genes (85 %) and is associated with binding to the

transcriptional activator CREB1. In accordance with this

finding, similar genome-wide analyses of MeCP2 binding,

CpG methylation and gene expression showed that MeCP2

binds methylated and unmethylated DNA preferentially to

actively expressed genes (Yasui et al. 2007). Several

MeCP2 target genes (UBE3A, DLX5, BDNF and PRODH)

have been identified (Samaco et al. 2005; Horike et al.

2005; Chang et al. 2006; Urdinguio et al. 2008), although

no direct link between MeCP2-dependent expression of

these genes and the phenotypic abnormality of Rett syn-

drome has so far been found. A role for oxidative stress as

a mechanism underlying the Rett phenotype has been

suggested (De Felice et al. 2012) mainly on the basis of

two observations: (i) oxidation of either a single guanine to

8-oxoG or of a single 5mC to 5hmC, significantly inhibits

(by at least an order of magnitude) binding of MeCP2 to

the oligonucleotide duplex (Valinluck et al. 2004); (ii)

several MeCP2 target genes affect the oxidative stress

response, such as BDNF, CREB or Prodh (Chang et al.

2006; Urdinguio et al. 2008). Finally, analysis of the

posttranslational modifications of MeCP2 proteins could

also provide insight into the effect on synaptic plasticity

mediated by the regulation of specific genes. Recent work

has shown that phosphorylation of MeCP2 at serine 421 is

induced by membrane depolarization and leads to the

regulation of BDNF transcription (Zhou et al. 2006).

Interestingly, a mouse model showed that phosphorylation

at serine 421 has widespread effects on synaptic plasticity

(Li et al. 2011). Neuronal activity also influences dephos-

phorylation of MeCP2 at serine 80, which alters the tran-

scription of several genes (Tao et al. 2009). Considering

these results together suggests that further research is

warranted into the role of these MeCP2 posttranslational

modifications (and other phosphorylation sites) in neuronal

plasticity. It should be stressed that recent evidence sug-

gests that MeCP2 is not only involved in transcriptional

regulation, but also possibly in RNA splicing (Young et al.

2005), chromatin condensation (similar to H1 function)

(Ishibashi et al. 2008) and the silencing of repetitive ele-

ments (Muotri et al. 2010). These findings demonstrate that

MeCP2 function and its involvement in Rett syndrome

might be more complex than previously appreciated.

Finally, MeCP2 mutations are also linked to a broad

spectrum of neurological disorders (Van Esch et al. 2005;

Villard 2007), and to autism (Carney et al. 2003), Angel-

man syndrome (Watson et al. 2001) and Prader–Willi

syndrome (Samaco et al. 2004). As mentioned before,

duplications or triplications on chromosome Xq28 con-

taining the MeCp2 region are also associated with the Lubs

X-linked mental retardation syndrome (Van Esch et al.

2005). Taking together, it must be highlighted that both

down- and overexpression of MeCP2 result in altered

neuron function, an aspect that must be especially con-

sidered for therapeutic purposes based on MeCP2 restora-

tion. Although it is clear that MeCP2 deficiency affects the

brain function, a definitive molecular pathology of the

MeCP2-associated disorders remains elusive. In this con-

cern, progresses are currently being carried out mainly due

to the development of appropriate experimental systems,

such as stem cell-based system allowing the synchronous

differentiation of neuronal progenitors in wild-type or

mutant MeCP2 (Yazdani et al. 2010) or by the develop-

ment of several mouse models that reproduce many traits

of Rett syndrome (Na et al. 2012). Hopefully their findings

will help us better understand the many facets of the

pathobiology of the disease.

Genetic alteration of histone modifiers

The histone modification network is very complex.

Post-transcriptional histone modifications can occur in

various histone proteins (e.g., H2B, H3, H4) and variants

364 Hum Genet (2013) 132:359–383

123

(e.g., H3.3) and affect different histone residues (lysine,

arginine, serine) located in their N-terminal tails (Esteller

2008; Bannister and Kouzarides 2011). Several chemical

groups [methyl, acetyl, phosphate, small ubiquitin-related

modifier (SUMO) and ADP-ribose] may be added in dif-

ferent degrees depending on the chemical modification

(mono-, di- or trimethylation) (Bannister and Kouzarides

2011). Cross-talk between histone marks can occur within

the same residue, in the same tail or among different his-

tone tails (Portela and Esteller 2010) and, as a conse-

quence, the functional significance of histone modifications

depends on the combination of all marks (the ‘‘histone

code’’). Furthermore, we must not forget that an additional

level of complexity exists due to the communication

between the epigenetic marks involving DNA, histone and

chromatin-related proteins. Histone modifications are

involved in gene transcription, although the consequence of

each mark depends on the residue affected and the type of

modification. In general, acetylation of lysines is associated

with transcriptional activation. However, methylation of

lysine 4 of the H3 histone is associated with active tran-

scription whereas methylation of lysines 9 and 27 is

associated with gene silencing (Bannister and Kouzarides

2011). Gene transcription is not the only characteristic that

is controlled in this way; histone modifications are a

mechanism for controlling chromatin structure and they

also affect more global biological processes such as DNA

repair, DNA replication, alternative splicing and chromo-

some condensation (Portela and Esteller 2010).

Addition of chemical groups to histone residues is a very

dynamic and reversible process catalyzed by two sets of

enzymes (and their protein complexes) that have antago-

nistic activities, enzymes that covalently attach the chem-

ical groups and others for removing them (Fig. 1).

Acetyltransferases (HATs) and histone deacetylases

(HDACs) are among the least specific histone modifiers

because they are able to modify several residues. Con-

versely, histone methyltransferases (HMTs), histone

demethylases (HDMTs) and kinases have higher specificity

(Portela and Esteller 2010). Genetic alterations of histone

modifier enzymes are frequently linked to human diseases.

In this regard, aberrations in the histone modification

profiles associated with cancer could be a consequence of

the genetic disruption of the epigenetic machinery

Normal development

…K… DRLVKRHRKAGGKGLGKGGKGRGS91 135812161820

- N- terminal

P PMeAcAc

Ac

MeAc

Ac

Me

CBP

P300

pCAFSWI/SNF

DNMT3B

PRMT1

CARM1

HDAC1HDAC2

SINA3MeCP2

CH3

H4

Pri-mRNA

Pre-mRNA

DroshaXPO5/mRNA

DICER

miRNA

RISC

Genetic alterations of epigenetic genes

Writers

Ac

Me

P

5mC DNMTs

HATs

HMTs

Kinases

Readers

Ac

Me

Bromodomains

ChromodomainsPHD domainsPWWP domainsTudor domains

5mC MBDs

Erasers

Ac

Me

P

HDACs

HDMTs

Phosphatases

Epigenetic therapies

HDAC inhibitorsHDMT inhibitors

DNA demethylating agentsOthers

miRNA

Fig. 1 Epigenetic mechanisms disrupted in human disorders. Epige-

netic mechanisms regulate chromatin function and cell identity.

Appropriate activity of enzymes controlling DNA methylation,

histone modifications and non-coding RNAs controls the temporal

and spatial patterns of gene expression, DNA repair and DNA

replication. Their deregulation may contribute to human diseases.

Epigenetic-based therapies, such as histone deacetylase inhibitors, can

partially alter the phenotype of the disease by recovering the aberrant

epigenetic patterns, and are a promising area in pharmacological

research. Chemical modifications at histone H4 are shown as a

representative example of histone marks. 5-mC; 5-methylcytosine; Achistone acetylation, DNMTs DNA methyltransferases, HATs histone

acetyltransferases, HDACs histone deacetylases, HDMTs histone

demethylases, HMTs histone methyltransferases, MBDs methyl-

binding domain proteins, Me histone methylation, P histone phos-

phorylation, PHD plant homeodomain, PWWP proline–tryptophan–

tryptophan–proline domain, XPO5 exportin-5

Hum Genet (2013) 132:359–383 365

123

(Berdasco and Esteller 2010). Several hematological

malignancies can be associated with chromosomal trans-

locations in the coding region of HATs (i.e., CBP-MOZ) or

HMTs (i.e., mixed-lineage leukemia 1 (MLL1), or nuclear

receptor binding SET domain protein 1 (NSD1). In solid

tumors, both HMT genes [such as EZH2, mixed-lineage

leukemia 2 (MLL2)] and DNMTs [i.e., Jumonji domain-

containing protein 2C (JMJD2C/GASC1)] are known to be

amplified (Berdasco and Esteller 2010). In this section we

will focus on non-tumoral human diseases in which the

epigenetic profile changes as a consequence of genetic

alterations in histone modifiers (Table 1).

MYST4 acetyltransferase (KAT6B) mutations

in Genitopatellar syndrome and Say-Barber-Biesecker-

Young-Simpson syndrome

Genitopatellar syndrome (GPS, MIM #606170) is a rare

skeletal dysplasia consisting of microcephaly, severe psy-

chomotor retardation and craniofacial defects, associated

with congenital flexion contractures of the lower extremi-

ties, abnormal or missing patellae, and urogenital anoma-

lies (Reardon 2002). To date it has been described in only

18 subjects (Campeau et al. 2012a). Whole-exome

sequencing identified mutations in the KAT6B acetyl-

transferase that lead to protein truncation (Campeau et al.

2012a; Simpson et al. 2012). All mutations are heterozy-

gous, exhibit autosomal dominant inheritance and occur in

the proximal portion of the last exon (Campeau et al.

2012a). KAT6B is a member of the MYST family of

proteins containing a conserved acetyltransferase domain

(Champagne et al. 1999). The absence of decreased

expression of the KAT6B transcript has been described in

GPS patients (Campeau et al. 2012a); however, GPS sub-

jects were characterized by decreased global acetylation of

histone H3 and H4 (Simpson et al. 2012).

Heterozygous mutations of KAT6B have also been

described in patients with Say-Barber-Biesecker-Young-

Simpson syndrome (SBBYS, MIM #603736) (Clayton-

Smith et al. 2011). The main clinical features associated

with the syndrome are distinctive facial appearance, severe

hypotonia and feeding problems, associated skeletal prob-

lems, cardiac defects, severe intellectual disability, delayed

motor milestones, and significantly impaired speech

(Clayton-Smith et al. 2011). Unlike with GPS syndrome,

mutations in SBBYS patients are located throughout the

gene or more distally in the last exon. Although GPS and

SBBYS share several clinical features, such as severe

intellectual disability, cardiac defects and genital abnor-

malities, the range of phenotypic alteration varies between

the two syndromes (Campeau et al. 2012b). Campeau and

collaborators created a database for establishing correla-

tions between phenotype and genotype in patients carrying

KAT6B mutations. Their preliminary results suggested that

the common features are a consequence of haploinsuffi-

ciency of the C-terminal region, whereas the unique phe-

notypes of GPS could arise from the expression of a

truncated protein that acquires new cellular functions

(Campeau et al. 2012b).

The molecular mechanisms of abrogation of KAT6B

function that lead to defective neural development are

not currently known. The Querkopf mouse model that

only expressed a 5 % of the MYST4 levels quantified for

wild-type mice leads to a phenotype commonly observed

in human syndromes: brain development defects, facial

dysmorphisms and alteration of bone growth (Thomas

et al. 2000), supporting the role of KAT6B in cerebral

cortex development. According to this idea, KAT6B is

expressed in developing cerebral cortex, adult neural

stem cells, osteoblasts and germ cells (Merson et al.

2006). A study performed in one Noonan-like phenotype

patient, with a chromosomal breakpoint in KAT6B

resulting in 50 % gene expression, and in Querkopf mice,

described a reduction of H3 acetylation that specifically

dysregulates the expression of genes in the MAPK

pathway (Kraft et al. 2011).

Genetic disruption of ep300/CBP acetyltransferases

in Rubinstein–Taybi syndrome

The Rubinstein–Taybi syndrome (RSTS, MIM #180849) is

a well-defined disease characterized by postnatal growth

deficiency, microcephaly, specific facial characteristics,

broad thumbs, big toes and intellectual disability. An

increased risk of tumors (mainly leukemia in childhood and

meningioma in adulthood) has been observed (Hennekam

2006). Although the exact molecular etiology of RSTS is

not clearly understood, it is widely accepted that RSTS is

associated with breakpoints, translocations, mutations and

microdeletions of chromosome 16p13.3 (Lacombe et al.

1992). Petrij et al. (1995) were the first to report that

mutations in the gene encoding the CREB binding protein

(CBP), located in the aforementioned region, could cause

RSTS. Recent investigations employing larger series of

RSTS samples detected CBP mutations in 45–55 % of

patients (Roelfsema et al. 2005; Tsai et al. 2011).

Furthermore, CBP has a homolog located at 22q13.2,

known as the E1A-binding protein p300 (p300) (Arany

et al. 1994). The exact frequency of genetic alterations of

p300 in RTS is not yet known, but some sources estimate it

to be 3 % (Bartholdi et al. 2007; Bartsch et al. 2010). Since

a cytogenetic or molecular abnormality in p300/CBP could

be detected in about 55 % of patients, further work to

explain the cause of the syndrome in the remaining 45 % of

patients without a ‘‘classic’’ genetic abnormality is still

needed.

366 Hum Genet (2013) 132:359–383

123

CREB binding protein was first identified as a nuclear

transcription coactivator that binds specifically to CREB

when it is phosphorylated (Chrivia et al. 1993), while p300

was originally described by protein-interaction assays with

the adenoviral E1A oncoprotein (Eckner et al. 1994).

Although both proteins are highly homologous (63 %

homology at the amino acid level) and have common

interaction partners, they have distinct cellular functions

and cannot always replace one another (Viosca et al. 2010).

There are two biological mechanisms whereby defects in

p300/CBP function could cause the RTS symptoms:

(i) CBP and p300 proteins act as cofactors for over 300

transcriptional factors, including several regulators

involved in neuronal activity, such as c-Fos, c-Jun, CREB

or NF-jß, known oncoproteins (myb), transforming viral

proteins (E1A, E6, large T antigen) and tumor-suppressor

proteins (p53, E2F, RB, Smads, RUNX, BRCA1) (Chan

and La Thangue 2001; Kasper et al. 2011); (ii) both pro-

teins have HAT activity that targets the N-terminal tails of

histones and contributes to transcriptional activation by

relaxing the structure of the nucleosomes (Ogryzko et al.

1996). CBP and ep300 have some common activities, such

as the acetylation of H4K5, H3K14, H3K18, H3K27 and

H3K56 (Das et al. 2009; Jin et al. 2011). However, they

also have unique properties such as substrate specificity

profiles that could explain functional differences of both

enzymes (McManus and Hendzel 2003). RSTS disorder

has been modeled in mice, and several heterozygous

cbp ± mice (homozygous cbp -/- mutants are embryonic

lethal) have been generated (Bourtchouladze et al. 2003;

Alarcon et al. 2004). These model animals exhibit deficits

in long-term memory and cognitive impairments reminis-

cent of human RSTS neural symptoms, confirming the role

of CBP in the etiology of the disease (Alarcon et al. 2004;

Korzus et al. 2004). At the molecular level, these mice

have reduced HAT activity, decreased acetylation of spe-

cific histone proteins and impaired CBP-dependent gene

expression (Alarcon et al. 2004; Korzus et al. 2004). The

role of CBP in neural differentiation and development has

been recently demonstrated in CBP genetically modified

models (Wang et al. 2010). Phosphorylation of CBP by

atypical protein kinase C f is necessary for CBP binding to

neural promoters, followed by histone acetylation and

transcriptional activation, leading to neural differentiation

of stem cell precursors (Wang et al. 2010). This mechanism

could explain how CBP alterations can result in cognitive

dysfunction of RSTS patients.

HDAC4 histone deacetylase mutations

in Brachydactyly-mental retardation syndrome

Brachydactyly-mental retardation syndrome (BDMR, MIM

#600430) is a complex disease that presents a wide

spectrum of clinical features, such as intellectual disabilities,

developmental delays, sleep disturbance, craniofacial and

skeletal abnormalities (including brachydactyly type E),

cardiac defects and autism (Aldred et al. 2004). Hetero-

zygous mutations in the HDAC4 deacetylase gene located

on chromosome 2q37.2 have been reported in BMRS

subjects (Williams et al. 2010; Morris et al. 2012). HDAC4

acts as a corepressor for transcription factors regulating the

expression of genes from the osteogenic, chondrogenic,

myogenic and neurogenic differentiation pathways (Miska

et al. 2001; Arnold et al. 2007; Chen and Cepko 2009).

HDAC4 is essential for the repression of RUNX2 and

MEF2 transcription factors in normal bone development

(Arnold et al. 2007). Indeed, mice with a deleted MEFC2

gene have impaired chondrogenic and osteogenic devel-

opment that antagonizes the phenotype of the Hdac4

mutant mice, which is similar to the human BDMR phe-

notype (Arnold et al. 2007; Rajan et al. 2009). These

results suggest that haploinsufficiency of HDAC4 causes

BDMR through its ability to regulate important master

genes of cellular differentiation.

Mutations in histone methyltransferase EHMT1

in Kleefstra syndrome

Kleefstra syndrome (MIM #610253), previously known as

9q subtelomeric deletion syndrome, is characterized

by severe intellectual disability, hypotonia, brachy(micro)

cephaly, epileptic seizures, flat face with hypertelorism,

synophrys, anteverted nares, everted lower lip, carp mouth

with macroglossia, and heart defects (Willemsen et al.

2012). The mutational landscape of Kleefstra patients

includes a microdeletion in the distal long arm of chro-

mosome 9q or intragenic loss of function mutations in the

histone methyltransferase EHMT1 (Kleefstra et al. 2006,

2009), both leading to haploinsufficiency of EHMT1.

EHMT1 is a specific HMT for lysine 9 at histone H3 and is

involved in gene repression (i.e., the NF-kB gene) (Ogawa

et al. 2002; Ea et al. 2012). It has not been established how

EHMT1 disruption results in the phenotypic skills of

Kleefstra syndrome, but recent research on EHMT mutants

in Drosophila melanogaster demonstrates that learning and

memory defects could be restored after re-expression of

EHMT (Kramer et al. 2011). Interestingly, the same

authors have recently identified new genetic mutations

affecting epigenetic modifiers in Kleefstra syndrome sub-

jects without mutations in the EHMT1 gene, including the

methyl-binding domain MBD5, the histone methyltrans-

ferase MLL3 or the chromatin-remodeling factor

SMARCB1 (Kleefstra et al. 2012). These findings highlight

the crucial role of epigenetic modifiers in brain develop-

ment and strongly emphasize the need to explore this area

of research further.

Hum Genet (2013) 132:359–383 367

123

NSD1 histone methyltransferase genetic alterations

in Sotos syndrome

Sotos syndrome (MIM #117550) is an autosomal dominant

condition characterized by overgrowth that results in tall

stature and macrocephaly, a distinctive facial appearance,

learning disability (Lapunzina 2005; Tatton-Brown and

Rahman 2007), and an increased incidence of malignant

neoplasms (Rahman 2005). The distinctive head shape and

size has led to Sotos syndrome sometimes being called

cerebral gigantism (Tatton-Brown and Rahman 2007).

Most cases of NSD1 mutational mechanisms, including

truncating, missense and splice-site mutations and

deletions, result in loss of function of the NSD1 protein

(Tatton-Brown and Rahman 2007). Mutations in the

nuclear receptor SET domain containing protein-1 gene

(NSD1), which encodes a histone methyltransferase of

lysine residues H3K36 and H4-K20, are found in patients

exhibiting the clinical symptoms of Sotos syndrome

(Kurotaki et al. 2002; Rayasam et al. 2003). Recently,

mutations in the NFIX gene were associated with a Sotos

syndrome-like phenotype without NSD1 mutations (Yon-

eda et al. 2012).

NSD1 is essential for early postimplantation of embryos

and NSD1 homozygous mutants are embryonic lethal,

although heterozygous mutant NSD1 are viable and fertile

(Rayasam et al. 2003). The NSD1 protein contains a

su(var)3-9, enhancer-of-zeste, trithorax (SET) domain

responsible for HMT activity and other functional domains,

including plant homeodomain (PHD) and proline–trypto-

phan–tryptophan–proline (PWWP) domains, both of which

are involved in a protein–protein interaction (Kurotaki

et al. 2001). Additionally, the potential of NSD1 to mono-

and dimethylate lysines K218 and K221 of the p65 subunit

of the immune response gene NF-kB has been noted (Lu

et al. 2010). NSD1 has previously been shown to interact

with nuclear receptors, such as Nizp1, a DNA-binding

transcriptional repressor (Huang et al. 1998; Nielsen et al.

2004), but to date there has been no evidence clarifying

whether NSD1 mutations contribute to deregulation of

brain function. It has become clear that NSD1 is a versatile

protein that can act as a corepressor or coactivator,

depending on the cellular context (Huang et al. 1998;

Pasillas et al. 2011). According to this idea, binding of

NSD1 PHD domains to target genes is guided by the

presence of specific histone marks of promoters (Pasillas

et al. 2011), specifically methylation at lysine H3K4 and

H3K9. By binding to trimethylated H3K9, NSD1 can

recognize genes that are transcriptionally repressed and

interact with other repression complexes (i.e., DNMT1 or

HP1), whereas its interaction with trimethylated H3K4

allows binding to active genes (Pasillas et al. 2011). In

addition, patients with genetic disruption of the NSD1 gene

have an increased risk of developing malignancy before

adulthood, including neuroblastoma, Wilms tumors and

hematological malignancies (Rahman 2005). NSD1 also

has a tumor-suppressor function (Berdasco et al. 2009).

NSD1 function is abrogated in neuroblastoma and glioma

cells by transcriptional silencing associated with CpG

island-promoter hypermethylation, and restoration of its

expression demonstrates that its tumor-suppressor features

are mediated by a mechanism dependent on MEIS1

expression (Berdasco et al. 2009).

Finally, occasional individuals have NSD1 defects that

overlap clinically with Sotos syndrome and other condi-

tions, such as Weaver syndrome 1 (WVS1, MIM #277590)

(Douglas et al. 2003). Beckwith–Wiedemann syndrome

(BWS, MIM #130650) is, like Sotos syndrome, an over-

growth syndrome. It is cause by deregulation of imprinted

growth-regulatory genes within the 11p15 region. Inter-

estingly, correlations between the two syndromes have

been found: first, unexplained Beckwith–Wiedemann

patients could be related to NSD1 deletions or mutations

(2/52 cases) and secondly, 11p15 anomalies (including the

KCNQ10T1 imprinting center) were identified in Sotos

syndrome cases (2/52) (Baujat et al. 2004; Mayo et al.

2012). A potential role for NSD1 in imprinting of the

11p15 region is suggested, although the molecular basis for

this association is not known.

EZH2 histone methyltransferase mutations

and Weaver syndrome

Weaver syndrome 2 (WVS2, MIM #614421) is an over-

growth syndrome characterized by tall stature, advanced

bone age, macrocephaly, hypertelorism, learning disabili-

ties and dysmorphic facial features (Weaver et al. 1974). A

predisposition to hematological malignancies has also been

reported (Basel-Vanagaite 2010). Heterozygous mutations

in the histone methyltransferase EZH2 gene on chromo-

some 7q36.1 have been identified in Weaver syndrome

patients (Gibson et al. 2012). EZH2 protein is a member of

the polycomb repressive complex 2 (PRC2), together with

SUZ12 and EED, which catalyses the trimethylation of

lysine H3K27 (Kirmizis et al. 2004). Mammalian EZH2

has critical roles in X-chromosome inactivations, genomic

imprinting during germline development, stem cell main-

tenance and cell lineage determination, including osteo-

genesis, myogenesis and hematogenesis (Chou et al. 2011;

Wyngaarden et al. 2011). A role for EZH2 in regulating the

circadian-clock functions has been suggested (Etchegaray

et al. 2006). In addition, mice with targeted mutant EZH2

beta cells have reduced beta cell proliferation and beta cell

mass (Chen et al. 2009), whereas mice with EZH2 mutant

satellite cells exhibit defects in muscle regeneration (Juan

et al. 2011). Some characteristics of these phenotypes are

368 Hum Genet (2013) 132:359–383

123

shared with that of human Weaver syndrome, such as

defects in limb development (Wyngaarden et al. 2011).

However, to date, few studies have provided any insight into

the specific contribution of EZH2 mutations in the Weaver

phenotype. As described above, some patients with Weaver

syndrome have a mutated NSD1 gene that is responsible for

the overgrowth Sotos syndrome (Douglas et al. 2003). There

are some clinical features common to both syndromes (i.e.,

developmental delay, overgrowth and macrocephaly), but

some features are specific to Weaver syndrome 2 (i.e., ret-

rognathia with a prominent chin crease or carpal bone age)

(Gibson et al. 2012). Indeed, EZH2 protein acts in the PI3K/

mTOR pathway, which has been associated with growth

defects. This suggests that the pathways through NSD1 and

EZH2 HTMs that contribute to overgrowth disorders can be

different (Tatton-Brown et al. 2011).

Histone methyltransferase (MLL2) and demethylase

(JMJD3) mutations in Kabuki syndrome

Kabuki syndrome 1 (MIM #147920) is an autosomal

dominant intellectual disability syndrome with additional

features, including a highly distinctive recognizable facial

phenotype characterized by long palpebral fissures with

eversion of the lateral third of the lower eyelids, a broad

and depressed nasal tip, large prominent earlobes, scoliosis,

short fifth finger, persistence of fingerpads, radiographic

abnormalities of the vertebrae, hands, and hip joints, and

others (Niikawa et al. 1981). Mutations in the histone

methyltransferase MLL2 gene are a major cause of Kabuki

syndrome (type 1) (Ng et al. 2010; Hannibal et al. 2011),

but mutations in the histone demethylase KDM6A have

also been found in Kabuki syndrome (type 2, MIM

#300867) (Lederer et al. 2012). MLL2 is a lysine H3K4-

specific histone methyltransferase that belongs to the SET1

family of proteins (Dillon et al. 2005), whereas KDM6A

(JMJD3) is a histone demethylase that specifically acts in

mono- di- and trimethylated lysine H3K27 (Hong et al.

2007; Lan et al. 2007). Both enzymes help regulate genes

from the myogenic lineage during embryogenesis (Aziz

et al. 2010; Seenundun et al. 2010). The identification of

mutations in MLL2 and KDM6A suggests a crucial effect

of altered histone methylation profiles on the phenotype of

Kabuki syndrome.

Histone demethylase PHF8 mutations in Siderius

X-linked mental retardation syndrome

Siderius X-linked mental retardation syndrome (MRXSSD,

MIM #300263) is an inherited condition that was first

described in 1999 as a heterogeneous intellectual disability

syndrome associated with cleft lip/cleft palate (Siderius

et al. 1999). Mutations in the PHD finger protein 8 (PHF8)

gene located on the Xp11 chromosome have been identi-

fied in subjects with MRXSSD (Laumonnier et al. 2005;

Abidi et al. 2007; Koivisto et al. 2007). The PHF8 protein

contains a PHD-type domain zinc finger domain and a

Jumanji domain, the latter conferring histone demethylat-

ing catalytic activity with specificity for: mono- and

dimethylated histone H3 at lysine K9 (Yu et al. 2010) and

mono-methyl histone H4 at lysine K20 (Qi et al. 2010).

PHF8 transcript is strongly expressed in the embryonic and

early postnatal steps of brain development (Laumonnier

et al. 2005), implying a connection between PHF8 muta-

tions and the MRXSSD phenotype. Some evidences of the

functional role of PHF8 in different experimental models

exist (Liu et al. 2010; Qi et al. 2010). In zebrafish, it has

been described that PHF8 regulates brain apoptosis and

craniofacial development through the transcriptional gene

regulation (Qi et al. 2010); in Caenorhabditis elegans, a

RNA interference-based functional genomic project iden-

tified PHF8 as a gene involved in controlling cellular

growth and differentiation during embyogenesis (Fernan-

dez et al. 2005); in human HeLa cells, PHF8 deficiency by

siRNA mechanisms leads to a delay in G1/S transition and

its dissociation from chromatin in early mitosis which

demonstrates an active role of PHF8 in the control of cell

cycle (Liu et al. 2010). Interestingly, PHF8 interacts with

another MRXSSD protein, the transcription factor ZNF711

(Kleine-Kohlbrecher et al. 2010). PHF8 and ZNF11 pro-

teins share a set of target genes, being of special relevance

the interaction of both proteins with JARID1C (KDM5C)

(also involved in alterations of the intelectual ability)

(Kleine-Kohlbrecher et al. 2010; Claes et al. 2000). In

addition to MRXSSD syndrome, defects affecting the

chromosomal region of PHF8 (a larger Xp11.22 deletion

that includes the FAM120C and WNK3 genes) have also

been associated with autism (Qiao et al. 2008).

Histone demethylase JARID1C (KDM5C) mutations

in Claes-Jensen X-linked mental retardation syndrome

Mutations in the histone demethylase JARID1C (KDM5C)

were first identified in individuals with X-linked mental

retardation syndrome with additional features of progres-

sive spastic paraplegia, facial hypotonia, aggressive

behavior and strabismus (MIM #300534) (Claes et al.

2000; Jensen et al. 2005). Heterogeneous clinical features

associated with XLMR and mutated KDM5C have been

consistently identified since then (Abidi et al. 2008; Santos-

Reboucas et al. 2011; Ounap et al. 2012). KDM5C gene

encodes a specific histone H3 demethylase at lysine K4

(Tahiliani et al. 2007). It is ubiquitously expressed

although fetal brain tissues have higher KDM5C expres-

sion levels than other tissues (Xu et al. 2008). The tran-

scriptional repressor activity of KDM5C is mediated by the

Hum Genet (2013) 132:359–383 369

123

Re-1 silencing transcription factor (REST) complex

(Tahiliani et al. 2007). Interestingly, knock-out of the

KDM5C complex results in increased trimethylation at

H3K4 and gain of expression of SCN2A and SYN1 neuro-

logical genes, and provides evidence of the contribution of

KDM5C complex to X-linked mental retardation defects

(Tahiliani et al. 2007).

Mutations in chromatin remodelers

Nucleosome positioning and, consequently, DNA accessi-

bility may be controlled by mechanisms that are indepen-

dent of histone-modifying enzymes. Several groups of

protein complexes (‘‘chromatin-remodeling complexes’’)

are known to restructure nucleosomes in an ATP hydro-

lysis-dependent manner. To date, four families of chro-

matin remodelers have been described in eukaryotes: SWI/

SNF, ISWI, NURD/Mi-2/CHD, and INO80/SWR1 (Har-

greaves and Crabtree 2011). The ATPase domain is a

common feature, but the composition of the different

subunits comprising the complex is highly variable. In a

similar manner, each ATPase domain may be targeted to

specific domains (i.e., bromodomain, DNA helicase, etc.).

Together, the binding affinity and the complex composition

confer unique features on chromatin remodelers in a wide

range of biological processes and genomic contexts.

SWI/SNF is one of the best studied chromatin-remod-

eling complexes in human cells and is composed of at least

15–20 subunits, including ATPases, catalytic subunits (i.e.,

SMARCA2 and SMARCA4) and structural components

involved in target recognition or stabilization functions

(i.e., ARID1A, ARID1B and SMARCC1) (Hargreaves and

Crabtree 2011). All members of this complex contain either

SMARCA2 (also known as Brahma protein) or SMARCA4

(also known as BRG1) as a catalytic unit. The two proteins

share 75 % amino acid homology (Santen et al. 2012). The

SWI/SNF complex plays a crucial role in cell differentia-

tion (Ho et al. 2009), cell cycle (Nagl et al. 2007) and DNA

repair (Park et al. 2006). In the human ISWI (imitation

switch) family of chromatin remodelers, the catalytic sub-

unit is represented by SNF2H and SNF2L proteins (Flaus

and Owen-Hughes 2011). ISWI complexes participate in

biological functions such as chromatin assembly, nucleo-

some spacing, DNA replication and activation or repres-

sion of transcriptional regulation (Erdel and Rippe 2011).

The CHD family also contains a SNF2L ATPase domain

together with tandem chromodomains in the N-terminal

region (Murawska and Brehm 2011). CHD complexes are

highly versatile and although they are involved in tran-

scriptional regulation, various CHD regulatory complexes

are involved in the initiation, elongation or termination of

transcription. Furthermore, specific CHD complexes, such

as Mi-2/nuRD, contain HDAC and MBD proteins in the

same complex (Murawska and Brehm 2011). The INO80

subfamily is the most recently identified SWI/SNF family

of chromatin remodelers. Mammalian INO80 complex

comprises the INO80 catalytic unit and Snf2-related CBP

activator protein (SRCAP) and p400 subunits (Morrison

and Shen 2009). The INO80 subfamily is the most evolu-

tionarily conserved of all the chromatin-remodeling com-

plexes due to the high degree of homology of its ATPase

subunit (Morrison and Shen 2009). Apart from regulation

of transcription, the INO80 complex is involved in genome

stability pathways, such as DNA repair, replication, telo-

mere regulation and centromere stability (Ho and Crabtree

2010). In summary, ATP-dependent enzymes that remodel

chromatin are important regulators of chromatin dyna-

mism. Evidence is emerging that alterations in such chro-

matin-remodeling complexes have consequences for

normal development. Some examples of genetic mutations

of chromatin-remodeling complexes in human diseases are

summarized in this section (Table 1).

ATXN7 mutation in Spinocerebellar Ataxia 7

Spinocerebellar Ataxia Type 7 (SCA7; MIM 164500) is an

autosomal dominant inherited neurodegenerative disorder

characterized by progressive cerebellar ataxia, including

dysarthria and dysphagia, and cone-rod and retinal dys-

trophy with progressive central visual loss resulting in

blindness in affected adults. The disease is caused by an

expanded CAG trinucleotide repeat encoding a polygluta-

mine tract in ataxin-7 (ATXN7) gene, from 4 to 35 repeats

in normal gene to a variable expansion of 36–306 repeats in

pathogenic ATXN7 variants (Garden and La Spada 2008).

ATXN7 protein is a transcription factor with important

roles in chromatin regulation through its effect on histone

modification and histone deubiquitination. It is a member

of the transcription coactivator complex STAGA (SPT3/

TAF9/GCN5) with acetyltransferase activity, but may also

be found in the USP22 deubiquitination complex (Sopher

et al. 2011). Although the nuclear expression of ATXN7 is

necessary for transcriptional regulation, a functional role

for cytoplasmic ATXN7 in the regulation of cytoskeletal

dynamics (mediated by its interaction with microtubules)

has recently been proposed (Nakamura et al. 2012).

Defects of the nuclear ATXN7 gene have been correlated

with the SCA7 phenotype (Chen et al. 2012a, b), although

the molecular mechanisms underlying the disease are not

clearly understood and the effect of epigenetic dysregula-

tion of target genes is still a matter of debate. Some results

from yeast and mice indicate that loss of the Gcn5 ace-

tyltransferase function triggered by polyQ-Atxn7, resulting

in chromatin structure changes, could be involved in the

SCA7 phenotype (Yoo et al. 2003; McMahon et al. 2005;

370 Hum Genet (2013) 132:359–383

123

Helmlinger et al. 2006). In contrast, loss of Gcn5 functions

in mice bearing polyQ-Atxn7 accelerates neuronal dys-

function in a mechanism that is independent of gene

expression changes (Chen et al. 2012a, b). Deciphering the

exact causal consequences of ATXN7 dysregulation in

SCA7 disease through its role as nuclear transcription

regulator or cytoplasmic function will need further

research.

ATRX mutations in alpha-thalassemia X-linked mental

retardation syndrome

ATR-X syndrome (MIM #301040) is an X-linked disorder

comprising severe psychomotor retardation, characteristic

facial features, genital abnormalities, and the blood disease

alpha-thalassemia (Gibbons et al. 1995). Mutations in the

ATRX gene located at Xq21.1 and coding for a member of

the SWI/SNF chromatin-remodeling family of proteins

underpin the molecular genetics of the disease (Gibbons

et al. 1995). The X-linked mental retardation-hypotonic

facies syndrome (MIM #309580) and the alpha-thalassemia

myelodysplasia syndrome (MIM #300448) are also asso-

ciated with mutations in the ATRX gene (Abidi et al. 2005;

Gibbons et al. 2003). The N-terminus contains a globular

domain, called ADD (ATRX-DNMT3-DNMT3L) that can

bind to the N-terminal of histone H3 (Argentaro et al.

2007). Indeed, ATRX is known to be required for the

incorporation of the histone H3.3 specifically at telomeric

sequences (Lewis et al. 2010). On the other hand, the

C-terminus contains seven helicase/ATPase domains that

share sequence homology with the SNF2 family of proteins

(Picketts et al. 1996). Through this domain, ATRX shows

in vitro ATP-dependent nucleosome remodeling and DNA

translocase activities (Gibbons et al. 2003). The ATRX

protein is expressed genome-wide, but is enriched at telo-

meric and subtelomeric regions (Law et al. 2010). In this

context, decreased ATRX expression is associated with

altered expression of telomere-associated RNA (Goldberg

et al. 2010) and the DNA-damage response during S-phase

at telomeric regions of pluripotent stem cells (Wong et al.

2010). The mechanisms by which ATRX is recruited to

telomeric regions are not fully understood, although ATRX

binding depends on trimethylation at K9H3 (Kourmouli

et al. 2005). It binds to tandem repeat sequences with

G-rich motifs and has been predicted to form non-B DNA

structures (Law et al. 2010). More importantly, the size of

the tandem repeats located in specific genes influences their

expression (Law et al. 2010), providing a molecular

explanation of how the same mutation at the ATRX gene

can result in different phenotypes. Furthermore, ATRX

mutations have been correlated with alterations in the DNA

methylation patterns of highly repeated sequences,

including rDNA, Y-specific satellite and subtelomeric

repeats (Gibbons et al. 2000). The interplay between

ATRX and the DNA methylation machinery is reinforced

by the discovery that the methyl-binding domain protein

MECP2 targeted the C-terminal helicase domain of ATRX

to heterochromatic foci (Nan et al. 2007). MeCP2 is also

mutated in Rett syndrome, so the finding suggests that

alteration of the MECP2–ATRX interaction leads to path-

ological changes that contribute to the intellectual dis-

ability phenotype observed in both syndromes.

Mutations in ERCC6 and Cockayne syndrome

Cockayne syndrome types A (CSA; MIM #216400) and B

(CSB; MIM #133540) are autosomal recessive disorders

caused by mutation in the ERCC8 and ERCC6 genes,

respectively (Henning et al. 1995; Laugel et al. 2010).

Approximately 62 % patients diagnosed with Cockayne

syndrome carry mutations of the ERCC6 gene (Laugel

et al. 2010). The syndromes are characterized by severe

postnatal growth failure, progressive neurological dys-

function and traits reminiscent of normal aging, such as

visual impairment and sensorineural hearing loss and loss

of adipose tissue (Licht et al. 2003). ERCC (excision

repair cross-complementing) genes are part of the nucle-

otide excision repair (NER) pathway, which are respon-

sible for removing DNA lesions such as UV-induced

DNA damage. ERCC6 is a nuclear protein containing a

SWI/SNF-like ATPase domain, a nucleotide-binding

domain and an ubiquitin-binding domain (Anindya et al.

2010). Apart from NER functions, ERCC6 is also

involved in transcription regulation, chromatin mainte-

nance and remodeling (Newman et al. 2006). At the

transcriptional level, CSB cooperates with the NurD/

CHD4 complex for controlling transcription of rRNA

genes (Xie et al. 2012). In this regard, CHD4/NuRD is

involved in maintaining silenced rRNA genes but in

permissive contexts (‘‘poised’’ for transcription), whereas

CSB mediates the transition from the permissive to the

active state (Xie et al. 2012). CSB-mediated activation

could be due, at least in part, in conjunction with the

CSB–G9a interaction, to an increase in trimethylated

K9H3 and recruitment of Pol-I to chromatin (Yuan et al.

2007). A new chromatin connection for CSB has recently

been proposed (Batenburg et al. 2012). Primary fibro-

blasts derived from a CSB patient had a dysfunctional

telomere structure (Batenburg et al. 2012). CSB knock-

down was accomplished with alterations in TERRA, a

large non-coding telomere repeat-containing RNA,

resulting in alterations of telomere length and integrity

(Batenburg et al. 2012). Finally, a role for CSB in con-

trolling key mitochondrial functions in addition to the

nucleolus function has been proposed (Berquist et al.

2012).

Hum Genet (2013) 132:359–383 371

123

SRCAP mutations and Floating-Harbor syndrome

Floating-Harbor syndrome (FHS; MIM #136140) is a rare

condition characterized by proportionate short stature,

delayed osseous maturation, language deficits and a typical

facial appearance. Mutations in the SNF2-related CBP

activator protein (SRCAP) cause the FHS syndrome (Hood

et al. 2012). SRCAP expression has also been linked to

cancer, whereby it positively modulates PSA antigen

expression and promotes proliferation in prostate cancer

cells (Slupianek et al. 2010) and potentiates Notch-

dependent gene activation (Eissenberg et al. 2005). With

regard to its molecular activity, SCARP catalyzes in vitro

incorporation of the histone variant H2A.Z into chromatin

(Ruhl et al. 2006), a histone with a well-known function in

transcription regulation and cell-cycle progression. As an

example, SRCAP expression in yeast is important for the

deposition of such histone variants in specific promoters

like SP-1, G3BP and FAD synthetase (Wong et al. 2007).

Interestingly, the demethylation effect after 5-Aza-20-deoxycytidine treatment, a drug approved by the US Food

and Drug Administration (FDA) for the treatment of

hematological malignancies, requires the activity of

SCARP to introduce H2A.Z, which facilitates the acqui-

sition of nucleosome-free regions (Yang et al. 2012). On

the other hand, SRCAP is also an interaction partner of the

histone acetyltransferase CBP, meaning that SRCAP-CBP

colocalization may occur at transcriptionally active sites

(Monroy et al. 2001).

CHD7 mutations and CHARGE syndrome

The acronym CHARGE (MIM # 214800) stands for colo-

boma of the eye, heart anomaly, choanal atresia, retarda-

tion of mental and somatic development, genital and/or

urinary abnormalities and ear abnormalities and/or deaf-

ness (Sanlaville et al. 2006). It is an autosomal dominant

condition with genotypic heterogeneity, although most

cases are due to the mutation or deletion of the chromod-

omain helicase DNA-binding domain protein-7 (CHD7), a

member of SNF2-like ATP-dependent chromatin-remod-

eling enzymes (Sanlaville et al. 2006). CHD7 mutations

have been also identified in Kallmann syndrome, a devel-

opmental disorder that shares with CHARGE some phe-

notypic features such as impaired olfaction and

hypogonadism (Kim et al. 2008). Mice with heterozygous

mutations in CHD7 are a good model for studying

CHARGE syndrome, and analyses of mouse mutant phe-

notypes have demonstrated a role in the development and

function of the neuronal system. CHD7 is necessary for

mammalian olfactory tissue development and function

(Layman et al. 2009), proliferation of inner ear neuroblasts

and inner ear morphogenesis in mice (Hurd et al. 2010),

promotion of the formation of multipotent migratory neural

crest that gives rise to craniofacial bones and cartilages,

and the peripheral nervous system, amongst others (Bajpai

et al. 2010). Recent in vitro studies have suggested that

CHD7 may directly regulate BMP4 expression, a protein

involved in cartilage and bone formation, by binding with

an enhancer element downstream of the BMP4 locus (Jiang

et al. 2012). More CHD7 targets have been identified, such

as the CHD7-dependent regulation (in association with

BRG1) of SOX9 and TEIST1 genes in human neural crest

cells (Bajpai et al. 2010). Mechanisms in which CHD7

regulates downstream genes vary in a tissue- and cell-

specific manner and depend on specific binding to meth-

ylated histone H3 lysine 4 in enhancer regions (Schnetz

et al. 2009). Although further research is needed, all these

findings suggest that mutations in CHD7 could transcrip-

tionally deregulate tissue-specific genes and developmental

genes resulting in the CHARGE phenotype.

Mutations in SWI/SNF complex family genes

in Coffin-Siris syndrome and mental retardation

Coffin-Siris syndrome (CSS, MIM #135900) or ‘‘fifth

digit’’ syndrome is a multiple congenital anomaly-mental

retardation syndrome characterized by severe develop-

mental delay, coarse facial features, hirsutism and absent

fifth fingernails, toenails and distal phalanges (Santen et al.

2012). In a recent study, 87 % of patients with CSS carried

a mutation in one or more members of the SWI/SNF family

of genes, which includes SMARCB1, SMARCA4, SMAR-

CA2, SMARCE1, ARID1A and ARID1B (Santen et al. 2012;

Tsurusaki et al. 2012). Interestingly, CSS patients carrying

different genetic mutations of the SWI/SNF chromatin-

remodeling factors gave rise to similar CSS phenotypes

(Tsurusaki et al. 2012), suggesting a general role for these

complexes in coordinating chromatin conformation and

gene expression. Deregulation of the SWI/SNF complexes

is also a common feature of tumorigenesis through its

function in mammalian differentiation, proliferation and

DNA repair (Reisman et al. 2009). However, the link

between SWI/SNF mutations and intellectual disorders is

still unclear. SMARCA2 and SMARCA4 are catalytic

subunits with ATPase activity, while ARID1A and

ARID1B are structural subunits involved in target recog-

nition and protein–protein interactions (Hargreaves and

Crabtree 2011). Both types of subunit are necessary to

regulate the transcription of several genes, such as c-FOS,

vimentin, CD44, cyclins, E-cadherin and important tran-

scription factors that have been functionally linked to SWI/

SNF (Reisman et al. 2009; Santen et al. 2012).

Mutations in SMARCA2 have also been described in

patients with the Nicolaides–Baraitser syndrome (NBS,

MIM # 601358), which is characterized by severe

372 Hum Genet (2013) 132:359–383

123

intellectual disability, early-onset seizures, short stature,

dysmorphic facial features and sparse hair (Van Houdt

et al. 2012; Wolff et al. 2012). Interestingly, Harikrishnan

et al. (2005) found that SMARCA2 associates with MECP2

and regulates FMR1 gene repression in mouse fibroblasts

and human T-lymphoblastic leukemia cells. Methylation at

promoter sites specified the recruitment of MECP2/

SMARCA2; while inhibition of methylation was associated

with complex release (Harikrishnan et al. 2005). These

results highlight an interesting link between epigenetic

marks and ATPase-dependent chromatin remodeling. How

SWI/SNF deregulation produces altered expression pat-

terns of genes associated with the CSS or NBS phenotype

is not clear, although some clues about the role of the

complex in neural differentiations could help interpret

some features (Seo et al. 2005; Lessard et al. 2007).

Mutations in non-coding RNA machinery

Non-coding RNAs, defined as functional RNA molecules

that are not translated into a protein, may also contribute to

the genesis of many human disorders (Table 1) (Esteller

2011). The best characterized ncRNAs in human condi-

tions are microRNAs (miRNAs) (Croce 2009), although

other ncRNAs members are emerging, such as small

nucleolar RNAs (snoRNAs), PIWI-interacting RNAs

(piRNAs), large intergenic non-coding RNAs (lincRNAs),

long non-coding RNAs (lncRNAs) and transcribed ultra-

conserved regions (T-UCRs), among others. If we focus on

miRNAs biogenesis, miRNAs are transcribed as individual

units (named primary miRNA (pri-miRNA)). After pro-

cessing by the Drosha complex, precursor miRNAs (pre-

miRNAs) are exported from the nucleus by the protein

exportin 5 (XPO5). Further processing by Dicer and TAR

RNA-binding protein 2 (TARBP2) generates mature

miRNAs, which are included into the RNA-induced

silencing complex (RISC). Once in this complex, miRNAs

could exert their function through degradation of protein-

coding transcripts or by translational repression. ncRNAs

profiles are frequently disrupted in different types of cancer

and in non-tumoral disorders, such as imprinting disorders,

rheumatoid arthritis, Rett syndrome and Alzheimer’s dis-

ease (Esteller 2011). Widespread alterations of ncRNAs

profiles could be also a consequence of genetic mutations

of the ncRNA-associated machinery. Some recent exam-

ples in this area are discussed in this section.

TARDBP mutations and amyotrophic lateral sclerosis

10 with or without frontotemporal lobar degeneration

Amyotrophic lateral sclerosis 10 (ALS10, MIM #612069)

with or without frontotemporal lobar degeneration (FTLD)

is an autosomal dominant neurodegenerative disorder

characterized by death of the motor neurons in the brain,

brainstem and spinal cord, resulting in fatal paralysis and

respiratory failure with a typical disease course of

1–5 years. Heterozygous mutations in the TAR DNA-

binding protein 43 (TARDBP) on chromosome 1p36,

encoding for the TDP43 protein, are present in a 50 % of

individuals affected by ALS10 (Ling et al. 2010). TARD-

BP is a member of the miRNA machinery. It has a double

effect on miRNA pathway. First, it enhances precursor

miRNA (pre-miRNA) production by both interacting with

the nuclear complex of Drosha or by direct binding to the

primary miRNAs (pri-miRNAs) in the nucleus; and sec-

ondly, it binds to the terminal loops of pre-miRNAs in the

cytoplasm by interaction with the Dicer complex (Kawa-

hara and Mieda-Sato 2012). Under non-pathological con-

ditions, TARDBP is mainly localized inside the nucleus,

but altered cellular distributions, including neuronal cyto-

plasmic, intranuclear inclusions and dystrophic neurites or

glial cytoplasmic locations are found in ALS10 and FTLD

(Arai et al. 2006; Neumann et al. 2006; Lagier-Tourenne

et al. 2010). Mutant TARDBP mice models developed a

similar phenotype than human TARDBP mutation

(Wegorzewska et al. 2009). Interestingly, no cytoplasmic

aggregates were found in mice mutants; suggesting that

other mechanisms rather than toxic cytoplasmic aggrega-

tion are underlying the molecular basis of ALS degenera-

tion (Wegorzewska et al. 2009). Results are in accordance

with a recent paper in which TDP-43 depletion in differ-

entiated Neruo2a results in decreased expression of miR-

132-3p and miR-132-5p. Further research of the specific

involvement of the aforementioned miRNAS (and others to

be explored) will strongly contribute to the understanding

of the pathogenesis of ALS10.

DGCR8 mutations and DiGeorge syndrome

DiGeorge syndrome (MIM #188400) is a complex disorder

characterized by learning disabilities, characteristic facial

appearance, submucous cleft palate, conotruncal heart

defects, thymic hypoplasia or aplasia, neonatal hypocal-

cemia, psychiatric illness and susceptibility to infection

due to a deficit of T cells (Goodship et al. 1998; Shiohama

et al. 2003). DiGeorge syndrome is caused by a 1.5 to 3.0-

Mb hemizygous deletion of chromosome 22q11.2 com-

prising the DiGeorge syndrome critical region gene 8

(DGCR8) (Shiohama et al. 2003) that encodes a double

stranded RNA-binding protein that is essential for miRNA

biogenesis. Specifically, DGCR8 is required in miRNA

maturation for processing pri-miRNAs to release pre-

miRNAs in the nucleus (Han et al. 2006). Genetic modified

mouse models carrying a hemizygous chromosomal defi-

ciency on chromosome 16 that spans a segment syntenic to

Hum Genet (2013) 132:359–383 373

123

the 1.5-Mb 22q11.2 microdeletion showed alterations in

the biogenesis of a set of miRNAs in the brain (Stark et al.

2008). Furthermore, DGCR8 deficiency resulted in altera-

tions of dendritic morphology, impaired sensorimotor gatin

and memory alterations similar to human DiGeorge phe-

notype (Stark et al. 2008). Additionally, it has been also

described that inactivation of a Dgcr8 conditional allele in

neural crest cells results in cardiovascular defects (Chapnik

et al. 2012). Similar results have been found in DGCR8

conditional knock-out mice embryos and knockout vascu-

lar smooth muscle cells (Chen et al. 2012a, b). DGCR8

deficiency was associated with down-regulation of the

miR-17/92 and miR-143/145 clusters in vascular smooth

muscle cells, reduced cell proliferation and increased

apoptosis (Chen et al. 2012a, b). These data provide spe-

cific explanations for cardiovascular and neuronal defects

that could explain, at least in part, the DiGeorge syndrome

phenotype.

DICER mutations in multinodular 1 Goiter disease

Autosomal dominant multinodular Goiter (MNG, MIM

#138800) is a disorder characterized by nodular over-

growth of the thyroid gland. In MNG type 1, some females

may also develop Sertoli–Leydig ovary tumors (Rio Frio

et al. 2011). Heterozygous mutations in DICER, a gene

encoding an RNase III endonuclease essential for micr-

oRNA processing, have recently been linked to MNG

pathogenesis (Rio Frio et al. 2011). Mutations in DICER

are also associated with pleuropulmonary blastoma (Hill

et al. 2009) and play a critical role in normal cardiac

function (Chen et al. 2008). DICER contains two RNase III

domains and a PAZ domain, a module that binds the end of

double-strand RNA (Macrae et al. 2006). Familial MNG

shows clear selective disruption of the PAZ domain, sug-

gesting a potential role of this domain in thyroid devel-

opment (Rio Frio et al. 2011). At the functional level,

lymphoblasts taken from MNG patients showed altered

miRNA compared with control profiles (i.e., LET7A and

miR345), suggesting a dysregulation of gene expression

patterns (Rio Frio et al. 2011). Further determination of the

consequences of specific microRNAs synthesis could be an

important topic for future research into MNG and other

DICER-associated pathologies.

Therapeutic applications of epigenetics

One of the main characteristics of epigenetic mechanisms

is their reversibility, making them potentially powerful

tools for curative pharmacological therapy (Fig. 1). Reac-

tivation of epigenetically silenced genes has been possible

for years by the treatment with DNA demethylation drugs,

such as zebularine or 5-aza-20-deoxycytidine (5-ADC), or

with histone deacetylase (HDAC) inhibitors, including

SAHA (suberoylanilide hydroxamic acid), valproic acid

(VPA) and trichostatin A (TSA). Indeed, some of these

drugs have significant antitumoral activity and the FDA has

approved the use of several of them for treating patients

(Kaminskas et al. 2005; Fiskus et al. 2008; Scuto et al.

2008). This approval has sparked a dramatic increase in the

development and trials of ‘‘epigenetic drugs’’ for treating

cancer and neural diseases (Kazantsev and Thompson

2008; Heyn and Esteller 2012). Although the most

advanced clinical trials are those corresponding to cancer

treatment, interest has grown in the fields of neurological

and neurodegenerative diseases in recent years (Day and

Sweatt 2012). The possibilities have only just begun to be

explored in human patients, but the basis of this therapy

has been confirmed in animal models.

Rubinstein–Taybi syndrome (RSTS) is probably the best

model for studying the therapeutic uses of HDAC inhibi-

tors that could compensate for the deficiency of HAT

activity (CBP mutations). RSTS has been modeled in mice,

and several heterozygous cbp ± mice (homozygous

cbp -/- mutants are embryonic lethal) have been gener-

ated (Alarcon et al. 2004). These models feature deficits in

long-term memory and cognitive impairments reminiscent

of human RTS neural symptoms, confirming the role of

CBP in the etiology of the disease. At the molecular level,

these mice have reduced HAT activity, decreased acetyla-

tion of specific histone proteins and impaired CBP-depen-

dent gene expression (Alarcon et al. 2004). Treatment with

HDAC inhibitors, such as SAHA or TSA, ameliorate defi-

cits in synaptic plasticity and cognition in cbp ± mice

(Hallam and Bourtchouladze 2006; Vecsey et al. 2007) by

enhancing transcriptional expression of specific neuronal

genes. In a similar manner, immortalized human lympho-

cytes derived from patients with RSTS showed reduced

acetylation levels, which primarily affect histones H2A and

H2B, compared to the histone acetylation levels of

immortalized human lymphocytes derived from patients

with Cornelia de Lange syndrome (a neurological disorder)

or healthy controls (Lopez-Atalaya et al. 2012). Interest-

ingly, the acetylation deficits in RSTS cells were rescued by

treatment with TSA (Lopez-Atalaya et al. 2012).

Some therapeutic approaches have been investigated in

Rett syndrome. Mecp2 knock-out mice also exhibit the

neurodevelopmental phenotype characteristic of human

Rett syndrome (Shahbazian et al. 2002). The cognitive

defect of MeCp2-deficient mice can be reverted by MecP2-

induced overexpression in mice (Collins et al. 2004) and in

MeCp2-deficient astrocytes (Lioy et al. 2011), suggesting

that this might well be an effective treatment for Rett

syndrome. Apart from this gene therapy strategy, which is

not really applicable in humans, pharmacological treat-

ments based on epigenetic targets are beginning to be

374 Hum Genet (2013) 132:359–383

123

explored as more feasible therapeutic interventions. Since

MeCP2 binds directly to methylated promoters in associ-

ation with the corepressor complex Sin3 and HDAC,

another therapeutic strategy could be based on targeting

HDAC activity. It has been widely demonstrated that

HDAC inhibitors enhance memory formation and neuronal

postnatal formation in various experimental models

(Kazantsev and Thompson 2008). There are some exam-

ples of the benefits of such treatments in the literature: the

alleviation of motor deficits in mouse models of Hunting-

ton disease by injections of SAHA (Hockly et al. 2003); the

delay of neurodegeneration and tauopathy in a mouse

model of Alzheimer’s disease through the activation of

SIRT1 by means of resveratrol treatments (Kilgore et al.

2010); and the reactivation of the frataxin genes by treat-

ments with benzamide-based HDAC inhibitors in immor-

talized cultured cells from Friedrich’s Ataxia syndrome

patients (Herman et al. 2006).

However, many questions remain unsolved concerning

drug potency, selectivity and permeability. Treatment with

HDAC inhibitors results in the re-expression of a limited

number of memory-associated genes (Vecsey et al. 2007).

This raises a question: are HDAC inhibitors’ treatments

potentially able to restore all types of genes? Or by contrast

do more ‘‘easier reverted’’ genes exist that can be phar-

macologically manipulated? If reversion is conditioned to

any factor (i.e., their genomic context or the acetylation

level) HDAC treatments would be more effective in certain

diseases than others. Multiple epigenetic enzymes could

contribute to a specific mark (especially histone acetyla-

tion) and furthermore, manipulating histone modifications

could also affect DNA methylation. Which is the best

target for inducing specific gene re-expression? Should

systemic or enzyme-targeted drugs be employed? This lack

of specificity means that almost none of the available

epigenetic drugs are isoform-specific and, for instance,

HDAC inhibitors have binding affinities for different

HDACs isoforms (Kilgore et al. 2010). It is expected that

development of subclass-specific inhibitors might contrib-

ute to minimize the side effects of the epigenetic therapy.

On the other hand, the tissue also influences its effective-

ness. DNA demethylating drugs are deoxycytidine analogs

that must be incorporated into the DNA after replication

cycles. As a consequence, the ‘‘speed’’ of base substitution

is strongly influenced by the rate of cell division. It is to be

expected that such DNA demethylating treatments will not

be very effective in neuronal cells with few or no cell

divisions. Further knowledge about the pathogenesis and

identification of the molecular/regulatory pathways that are

altered in the disease will enable the optimal target to be

identified and the therapeutic potential of epigenetic-based

drugs to be exploited. Despite these difficulties, epigenetic-

based therapy may become a successful intervention in the

treatment of human disorders associated with epigenetic

alterations.

Acknowledgments This work was supported by grants from the

European Research Council (Advanced EPINORC), the Fondo de

Investigaciones Sanitarias (PI08-1345 and PI10/02267), the Ministe-

rio de Ciencia e Innovacion (SAF2011-22803), the Fundacio La

Marato de TV3 (111430/31), European COST Action TD09/05, and

the Health Department of the Catalan Government (Generalitat de

Catalunya). M.E. is an Institucio Catalana de Recerca i Estudis

Avancats Research Professor.

References

Aapola U, Liiv I, Peterson P (2002) Imprinting regulator DNMT3L is

a transcriptional repressor associated with histone deacetylase

activity. Nucleic Acids Res 30(16):3602–3608

Abidi FE, Cardoso C, Lossi AM, Lowry RB, Depetris D, Mattei MG,

Lubs HA, Stevenson RE, Fontes M, Chudley AE, Schwartz CE

(2005) Mutation in the 5-prime alternatively spliced region of

the XNP/ATR-X gene causes Chudley-Lowry syndrome. Europ

J Hum Genet 13:176–183

Abidi FE, Miano MG, Murray JC, Schwartz CE (2007) A novel

mutation in the PHF8 gene is associated with X-linked mental

retardation with cleft lip/cleft palate. Clin Genet 72:19–22

Abidi FE, Holloway L, Moore CA, Weaver DD, Simensen RJ,

Stevenson RE, Rogers RC, Schwartz CE (2008) Mutations in

JARID1C are associated with X-linked mental retardation, short

stature and hyperreflexia. J Med Genet 45:787–793

Alarcon JM, Malleret G, Touzani K, Vronskaya S, Ishii S, Kandel ER,

Barco A (2004) Chromatin acetylation, memory, and LTP are

impaired in CBP ± mice: a model for the cognitive defect in

Rubinstein-Taybi syndrome and its amelioration. Neuron

42:947–959

Aldred MA, Sanford ROC, Thomas NS, Barrow MA, Wilson LC,

Brueton LA, Bonaglia MC, Hennekam RCM, Eng C, Dennis

NR, Trembath RC (2004) Molecular analysis of 20 patients with

2q37.3 monosomy: definition of minimum deletion intervals for

key phenotypes. J Med Genet 41:433–439

Amir RE, Van den Veyver IB, Wan M, Tran CQ, Francke U, Zoghbi

HY (1999) Rett syndrome is caused by mutations in X-linked

MECP2, encoding methyl-CpG-binding protein 2. Nat Genet

23(2):185–188

Anindya R, Mari PO, Kristensen U, Kool H, Giglia-Mari G,

Mullenders LH, Fousteri M, Vermeulen W, Egly JM, Svejstrup

JQ (2010) A ubiquitin-binding domain in Cockayne syndrome B

required for transcription-coupled nucleotide excision repair.

Mol Cell 38(5):637–648

Arai T, Hasegawa M, Akiyama H, Ikeda K, Nonaka T, Mori H, Mann

D, Tsuchiya K, Yoshida M, Hashizume Y, Oda T (2006) TDP-43

is a component of ubiquitin-positive tau-negative inclusions in

frontotemporal lobar degeneration and amyotrophic lateral

sclerosis. Biochem Biophys Res Commun 351(3):602–611

Arany Z, Sellers WR, Livingston DM, Eckner R (1994) E1A-

associated p300 and CREB-associated CBP belong to a

conserved family of coactivators. Cell 77(6):799–800

Argentaro A, Yang JC, Chapman L, Kowalczyk MS, Gibbons RJ,

Higgs DR, Neuhaus D, Rhodes D (2007) Structural conse-

quences of disease-causing mutations in the ATRX-DNMT3-

DNMT3L (ADD) domain of the chromatin-associated protein

ATRX. Proc Natl Acad Sci USA 104(29):11939–11944

Arnold MA, Kim Y, Czubryt MP, Phan D, McAnally J, Qi X, Shelton

JM, Richardson JA, Bassel-Duby R, Olson EN (2007) MEF2C

Hum Genet (2013) 132:359–383 375

123

transcription factor controls chondrocyte hypertrophy and bone

development. Dev Cell 12(3):377–389

Aziz A, Liu QC, Dilworth FJ (2010) Regulating a master regulator:

establishing tissue-specific gene expression in skeletal muscle.

Epigenetics 5(8):691–695

Bajpai R, Chen DA, Rada-Iglesias A, Zhang J, Xiong Y, Helms J,

Chang CP, Zhao Y, Swigut T, Wysocka J (2010) CHD7

cooperates with PBAF to control multipotent neural crest

formation. Nature 463(7283):958–962

Ballas N, Lioy DT, Grunseich C, Mandel G (2009) Non-cell

autonomous influence of MeCP2-deficient glia on neuronal

dendritic morphology. Nat Neurosci 12(3):311–317

Bannister AJ, Kouzarides T (2011) Regulation of chromatin by

histone modifications. Cell Res 21(3):381–395

Bartholdi D, Roelfsema JH, Papadia F, Breuning MH, Niedrist D,

Hennekam RC, Schinzel A, Peters DJ (2007) Genetic heteroge-

neity in Rubinstein-Taybi syndrome: delineation of the pheno-

type of the first patients carrying mutations in EP300. J Med

Genet 44(5):327–333

Bartsch O, Labonte J, Albrecht B, Wieczorek D, Lechno S, Zechner

U, Haaf T (2010) Two patients with EP300 mutations and facial

dysmorphism different from the classic Rubinstein-Taybi

syndrome. Am J Med Genet 152A:181–184

Basel-Vanagaite L (2010) Acute lymphoblastic leukemia in Weaver

syndrome. Am J Med Genet 152A:383–386

Batenburg NL, Mitchell TR, Leach DM, Rainbow AJ, Zhu XD (2012)

Cockayne Syndrome group B protein interacts with TRF2 and

regulates telomere length and stability. Nucleic Acids Res

40(19):9661–9674

Baujat G, Rio M, Rossignol S, Sanlaville D, Lyonnet S, Le Merrer M,

Munnich A, Gicquel C, Cormier-Daire V, Colleaux L (2004)

Paradoxical NSD1 mutations in Beckwith-Wiedemann syn-

drome and 11p15 anomalies in Sotos syndrome. Am J Hum

Genet 74:715–720

Berdasco M, Esteller M (2010) Aberrant epigenetic landscape in

cancer: how cellular identity goes awry. Dev Cell 19(5):698–711

Berdasco M, Esteller M (2011) DNA methylation in stem cell renewal

and multipotency. Stem Cell Res Ther 2(5):42

Berdasco M, Ropero S, Setien F, Fraga MF, Lapunzina P, Losson R,

Alaminos M, Cheung NK, Rahman N, Esteller M (2009)

Epigenetic inactivation of the Sotos overgrowth syndrome gene

histone methyltransferase NSD1 in human neuroblastoma and

glioma. Proc Natl Acad Sci USA 106(51):21830–21835

Berquist BR, Canugovi C, Sykora P, Wilson DM 3rd, Bohr VA

(2012) Human Cockayne syndrome B protein reciprocally

communicates with mitochondrial proteins and promotes tran-

scriptional elongation. Nucleic Acids Res 40(17):8392–8405

Bourtchouladze R, Lidge R, Catapano R, Stanley J, Gossweiler S,

Romashko D, Scott R, Tully T (2003) A mouse model of

Rubinstein-Taybi syndrome: defective long-term memory is

ameliorated by inhibitors of phosphodiesterase 4. Proc Natl Acad

Sci USA 100:10518–10522

Brun ME, Lana E, Rivals I, Lefranc G, Sarda P, Claustres M,

Megarbane A, De Sario A (2011) Heterochromatic genes

undergo epigenetic changes and escape silencing in immunode-

ficiency, centromeric instability, facial anomalies (ICF) syn-

drome. PLoS ONE 6(4):e19464

Campeau PM, Kim JC, Lu JT, Schwartzentruber JA, Abdul-Rahman

OA, Schlaubitz S, Murdock DM, Jiang MM, Lammer EJ, Enns

GM, Rhead WJ, Rowland J, Robertson SP, Cormier-Daire V,

Bainbridge MN, Yang XJ, Gingras MC, Gibbs RA, Rosenblatt

DS, Majewski J, Lee BH (2012a) Mutations in KAT6B,

encoding a histone acetyltransferase, cause Genitopatellar syn-

drome. Am J Hum Genet 90(2):282–289

Campeau PM, Lu JT, Dawson BC, Fokkema IF, Robertson SP,

Gibbs RA, Lee BH (2012b) The KAT6B-related disorders

genitopatellar syndrome and Ohdo/SBBYS syndrome have

distinct clinical features reflecting distinct molecular mecha-

nisms. Hum Mutat 33(11):1520–1525

Carney RM, Wolpert CM, Ravan SA, Shahbazian M, Ashley-Koch A,

Cuccaro ML, Vance JM, Pericak-Vance MA (2003) Identifica-

tion of MeCP2 mutations in a series of females with autistic

disorder. Pediat Neurol 28:205–211

Chahrour M, Jung SY, Shaw C, Zhou X, Wong ST, Qin J, Zoghbi HY

(2008) MeCP2, a key contributor to neurological disease,

activates and represses transcription. Science 320(5880):

1224–1229

Champagne N, Bertos NR, Pelletier N, Wang AH, Vezmar M, Yang

Y, Heng HH, Yang XJ (1999) Identification of a human histone

acetyltransferase related to monocytic leukemia zinc finger

protein. J Biol Chem 274:28528–28536

Chan HM, La Thangue NB (2001) p300/CBP proteins: HATs for

transcriptional bridges and scaffolds. J Cell Sci 114:2363–2373

Chang Q, Khare G, Dani V, Nelson S, Jaenisch R (2006) The disease

progression of Mecp2 mutant mice is affected by the level of

BDNF expression. Neuron 49(3):341–348

Chapnik E, Sasson V, Blelloch R, Hornstein E (2012) Dgcr8 controls

neural crest cells survival in cardiovascular development. Dev

Biol 362(1):50–56

Chen B, Cepko CL (2009) HDAC4 regulates neuronal survival in

normal and diseased retinas. Science 323(5911):256–259

Chen JF, Murchison EP, Tang R, Callis TE, Tatsuguchi M, Deng Z,

Rojas M, Hammond SM, Schneider MD, Selzman CH, Meissner

G, Patterson C, Hannon GJ, Wang DZ (2008) Targeted deletion

of Dicer in the heart leads to dilated cardiomyopathy and heart

failure. Proc Natl Acad Sci USA 105(6):2111–2116

Chen H, Gu X, Su IH, Bottino R, Contreras JL, Tarakhovsky A, Kim

SK (2009) Polycomb protein Ezh2 regulates pancreatic beta-cell

Ink4a/Arf expression and regeneration in diabetes mellitus.

Genes Dev 23(8):975–985

Chen YC, Gatchel JR, Lewis RW, Mao CA, Grant PA, Zoghbi HY,

Dent SY (2012a) Gcn5 loss-of-function accelerates cerebellar

and retinal degeneration in a SCA7 mouse model. Hum Mol

Genet 21(2):394–405

Chen Z, Wu J, Yang C, Fan P, Balazs L, Jiao Y, Lu M, Gu W, Li C,

Pfeffer LM, Tigyi G, Yue J (2012b) DiGeorge syndrome critical

region 8 (DGCR8) protein-mediated microRNA biogenesis is

essential for vascular smooth muscle cell development in mice.

J Biol Chem 287(23):19018–19028

Chou RH, Yu YL, Hung MC (2011) The roles of EZH2 in cell lineage

commitment. Am J Transl Res 3(3):243–250

Chrivia JC, Kwok RP, Lamb N, Hagiwara M, Montminy MR,

Goodman RH (1993) Phosphorylated CREB binds specifically to

the nuclear protein CBP. Nature 365:855–859

Claes S, Devriendt K, Van Goethem G, Roelen L, Meireleire J,

Raeymaekers P, Cassiman JJ, Fryns JP (2000) Novel syndromic

form of X-linked complicated spastic paraplegia. Am J Med

Genet 94:1–4

Clayton-Smith J, O’Sullivan J, Daly S, Bhaskar S, Day R, Anderson

B, Voss AK, Thomas T, Biesecker LG, Smith P, Fryer A,

Chandler KE, Kerr B, Tassabehji M, Lynch SA, Krajewska-

Walasek M, McKee S, Smith J, Sweeney E, Mansour S,

Mohammed S, Donnai D, Black G (2011) Whole-exome-

sequencing identifies mutations in histone acetyltransferase gene

KAT6B in individuals with the Say-Barber-Biesecker variant of

Ohdo syndrome. Am J Hum Genet 89:675–681

Collins AL, Levenson JM, Vilaythong AP, Richman R, Armstrong

DL, Noebels JL, David Sweatt J, Zoghbi HY (2004) Mild

overexpression of MeCP2 causes a progressive neurological

disorder in mice. Hum Mol Genet 13(21):2679–2689

Croce CM (2009) Causes and consequences of microRNA dysreg-

ulation in cancer. Nat Rev Genet 10(10):704–714

376 Hum Genet (2013) 132:359–383

123

Das C, Lucia MS, Hansen KC, Tyler JK (2009) CBP/p300-mediated

acetylation of histone H3 on lysine 56. Nature 459(7243):

113–117

Day JJ, Sweatt JD (2012) Epigenetic treatments for cognitive

impairments. Neuropsychopharmacology 37(1):247–260

De Felice C, Signorini C, Leoncini S, Pecorelli A, Durand T, Valacchi

G, Ciccoli L, Hayek J (2012) The role of oxidative stress in Rett

syndrome: an overview. Ann N Y Acad Sci 1259(1):121–135

De Greef JC, Wang J, Balog J, den Dunnen JT, Frants RR,

Straasheijm KR, Aytekin C, van der Burg M, Duprez L, Ferster

A, Gennery AR, Gimelli G, Reisli I, Schuetz C, Schulz A,

Smeets DF, Sznajer Y, Wijmenga C, van Eggermond MC, van

Ostaijen-Ten Dam MM, Lankester AC, van Tol MJ, van den

Elsen PJ, Weemaes CM, van der Maarel SM (2011) Mutations in

ZBTB24 are associated with immunodeficiency, centromeric

instability, and facial anomalies syndrome type 2. Am J Hum

Genet 88(6):796–804

Dillon SC, Zhang X, Trievel RC, Cheng X (2005) The SET-domain

protein superfamily: protein lysine methyltransferases. Genome

Biol 6(8):227

Douglas J, Hanks S, Temple IK, Davies S, Murray A, Upadhyaya M,

Tomkins S, Hughes HE, Cole TR, Rahman N (2003) NSD1

mutations are the major cause of Sotos syndrome and occur in

some cases of Weaver syndrome but are rare in other overgrowth

phenotypes. Am J Hum Genet 72(1):132–143

Ea CK, Hao S, Yeo KS, Baltimore D (2012) EHMT1 binds to NF-jB

p50 and represses gene expression. J Biol Chem 287(37):

31207–31217

Eckner R, Ewen ME, Newsome D, Gerdes M, DeCaprio JA,

Lawrence JB, Livingston DM (1994) Molecular cloning and

functional analysis of the adenovirus E1A-associated 300-kD

protein (p300) reveals a protein with properties of a transcrip-

tional adaptor. Genes Dev 8:869–884

Edbauer D, Neilson JR, Foster KA, Wang CF, Seeburg DP, Batterton

MN, Tada T, Dolan BM, Sharp PA, Sheng M (2010) Regulation

of synaptic structure and function by FMRP-associated microR-

NAs miR-125b and miR-132. Neuron 65(3):373–384

Ehrlich M (2003) The ICF syndrome, a DNA methyltransferase 3B

deficiency and immunodeficiency disease. Clin Immunol

109:17–28

Eissenberg JC, Wong M, Chrivia JC (2005) Human SRCAP and

Drosophila melanogaster DOM are homologs that function in the

notch signaling pathway. Mol Cell Biol 25(15):6559–6569

Erdel F, Rippe K (2011) Chromatin remodelling in mammalian cells

by ISWI-type complexes–where, when and why? FEBS J

278(19):3608–3618

Esteller M (2008) Epigenetics in cancer. N Engl J Med

358(11):1148–1159

Esteller M (2011) Non-coding RNAs in human disease. Nat Rev

Genet 12(12):861–874

Etchegaray JP, Yang X, DeBruyne JP, Peters AH, Weaver DR,

Jenuwein T, Reppert SM (2006) The polycomb group protein

EZH2 is required for mammalian circadian clock function. J Biol

Chem 281(30):21209–21215

Fatemi M, Hermann A, Pradhan S, Jeltsch A (2001) The activity of

the murine DNA methyltransferase Dnmt1 is controlled by

interaction of the catalytic domain with the N-terminal part of

the enzyme leading to an allosteric activation of the enzyme after

binding to methylated DNA. J Mol Biol 309(5):1189–1199

Feng J, Zhou Y, Campbell SL, Le T, Li E, Sweatt JD, Silva AJ, Fan G

(2010) Dnmt1 and Dnmt3a maintain DNA methylation and

regulate synaptic function in adult forebrain neurons. Nat

Neurosci 13(4):423–430

Fernandez AG, Gunsalus KC, Huang J, Chuang LS, Ying N, Liang

HL, Tang C, Schetter AJ, Zegar C, Rual JF, Hill DE, Reinke V,

Vidal M, Piano F (2005) New genes with roles in the C. elegans

embryo revealed using RNAi of ovary-enriched ORFeome

clones. Genome Res 15(2):250–259

Fiskus W, Wang Y, Joshi R, Rao R, Yang Y, Chen J, Kolhe R, Balusu

R, Eaton K, Lee P, Ustun C, Jillella A, Buser CA, Peiper S,

Bhalla K (2008) Cotreatment with vorinostat enhances activity

of MK-0457 (VX-680) against acute and chronic myelogenous

leukemia cells. Clin Cancer Res 14(19):6106–6115

Flaus A, Owen-Hughes T (2011) Mechanisms for ATP-dependent

chromatin remodelling: the means to the end. FEBS J

278(19):3579–3595

Franklin TB, Mansuy IM (2011) The involvement of epigenetic

defects in mental retardation. Neurobiol Learn Mem 96(1):61–67

Froyen G, Bauters M, Voet T, Marynen P (2006) X-linked mental

retardation and epigenetics. J Cell Mol Med 10(4):808–825

Garden GA, La Spada AR (2008) Molecular pathogenesis and cellular

pathology of spinocerebellar ataxia type 7 neurodegeneration.

Cerebellum 7(2):138–149

Giacometti E, Luikenhuis S, Beard C, Jaenisch R (2007) Partial

rescue of MeCP2 deficiency by postnatal activation of MeCP2.

Proc Natl Acad Sci USA 104(6):1931–1936

Gibbons RJ, Picketts DJ, Villard L, Higgs DR (1995) Mutations in a

putative global transcriptional regulator cause X-linked mental

retardation with alpha-thalassemia (ATR-X syndrome). Cell

80:837–845

Gibbons RJ, McDowell TL, Raman S, O’Rourke DM, Garrick D,

Ayyub H, Higgs DR (2000) Mutations in ATRX, encoding a

SWI/SNF-like protein, cause diverse changes in the pattern of

DNA methylation. Nat Genet 24:368–371

Gibbons RJ, Pellagatti A, Garrick D, Wood WG, Malik N, Ayyub H,

Langford C, Boultwood J, Wainscoat JS, Higgs DR (2003)

Identification of acquired somatic mutations in the gene encod-

ing chromatin-remodeling factor ATRX in the alpha-thalassemia

myelodysplasia syndrome (ATMDS). Nat Genet 34:446–449

Gibson WT, Hood RL, Zhan SH, Bulman DE, Fejes AP, Moore R,

Mungall AJ, Eydoux P, Babul-Hirji R, An J, Marra MA, FORGE

Canada Consortium, Chitayat D, Boycott KM, Weaver DD,

Jones SJM (2012) Mutations in EZH2 cause Weaver syndrome.

Am J Hum Genet 90:110–118

Goldberg AD, Banaszynski LA, Noh KM, Lewis PW, Elsaesser SJ,

Stadler S, Dewell S, Law M, Guo X, Li X, Wen D, Chapgier A,

DeKelver RC, Miller JC, Lee YL, Boydston EA, Holmes MC,

Gregory PD, Greally JM, Rafii S, Yang C, Scambler PJ, Garrick

D, Gibbons RJ, Higgs DR, Cristea IM, Urnov FD, Zheng D

(2010) Allis CD (2010) Distinct factors control histone variant

H3.3 localization at specific genomic regions. Cell

140(5):678–691

Goodship J, Cross I, LiLing J, Wren C (1998) A population study of

chromosome 22q11 deletions in infancy. Arch Dis Child

79:348–351

Guy J, Gan J, Selfridge J, Cobb S, Bird A (2007) Reversal of

neurological defects in a mouse model of Rett syndrome.

Science 315(5815):1143–1147

Hallam TM, Bourtchouladze R (2006) Rubinstein-Taybi syndrome:

molecular findings and therapeutic approaches to improve

cognitive dysfunction. Cell Mol Life Sci 63:1725–1735

Han J, Lee Y, Yeom KH, Nam JW, Heo I, Rhee JK, Sohn SY, Cho Y,

Zhang BT, Kim VN (2006) Molecular basis for the recognition

of primary microRNAs by the Drosha-DGCR8 complex. Cell

125(5):887–901

Hannibal MC, Buckingham KJ, Ng SB, Ming JE, Beck AE, McMillin

MJ, Gildersleeve HI, Bigham AW, Tabor HK, Mefford HC,

Cook J, Yoshiura K, Matsumoto T, Matsumoto N, Miyake N,

Tonoki H, Naritomi K, Kaname T, Nagai T, Ohashi H,

Kurosawa K, Hou JW, Ohta T, Liang D, Sudo A, Morris CA,

Banka S, Black GC, Clayton-Smith J, Nickerson DA, Zackai

EH, Shaikh TH, Donnai D, Niikawa N, Shendure J, Bamshad MJ

Hum Genet (2013) 132:359–383 377

123

(2011) Spectrum of MLL2 (ALR) mutations in 110 cases of

Kabuki syndrome. Am J Med Genet 155A:1511–1516

Hargreaves DC, Crabtree GR (2011) ATP-dependent chromatin

remodeling: genetics, genomics and mechanisms. Cell Res

21(3):396–420

Harikrishnan KN, Chow MZ, Baker EK, Pal S, Bassal S, Brasacchio

D, Wang L, Craig JM, Jones PL, Sif S, El-Osta A (2005) Brahma

links the SWI/SNF chromatin-remodeling complex with

MeCP2-dependent transcriptional silencing. Nat Genet 37:

254–264

Helmlinger D, Hardy S, Abou-Sleymane G, Eberlin A, Bowman AB,

Gansmuller A, Picaud S, Zoghbi HY, Trottier Y, Tora L, Devys

D (2006) Glutamine-expanded ataxin-7 alters TFTC/STAGA

recruitment and chromatin structure leading to photoreceptor

dysfunction. PLoS Biol 4(3):e67

Hennekam RCM (2006) Rubinstein-Taybi syndrome. Europ. J Hum

Genet 14:981–985

Henning KA, Li L, Iyer N, McDaniel LD, Reagan MS, Legerski R,

Schultz RA, Stefanini M, Lehmann AR, Mayne LV, Friedberg

EC (1995) The Cockayne syndrome group A gene encodes a WD

repeat protein that interacts with CSB protein and a subunit of

RNA polymerase II TFIIH. Cell 82:555–564

Herman D, Jenssen K, Burnett R, Soragni E, Perlman SL, Gottesfeld

JM (2006) Histone deacetylase inhibitors reverse gene silencing

in Friedreich’s ataxia. Nat Chem Biol 2(10):551–558

Heyn H, Esteller M (2012) DNA methylation profiling in the clinic:

applications and challenges. Nat Rev Genet 13(10):679–692

Heyn H, Vidal E, Sayols S, Sanchez-Mut JV, Moran S, Medina I,

Sandoval J, Simo-Riudalbas L, Szczesna K, Huertas D, Gatto S,

Matarazzo MR, Dopazo J, Esteller M (2012) Whole-genome

bisulfite DNA sequencing of a DNMT3B mutant patient.

Epigenetics 7(6):542–550

Hill DA, Ivanovich J, Priest JR, Gurnett CA, Dehner LP, Desruisseau

D, Jarzembowski JA, Wikenheiser-Brokamp KA, Suarez BK,

Whelan AJ, Williams G, Bracamontes D, Messinger Y, Good-

fellow PJ (2009) DICER1 mutations in familial pleuropulmonary

blastoma. Science 325:965

Ho L, Crabtree GR (2010) Chromatin remodelling during develop-

ment. Nature 463(7280):474–484

Ho L, Jothi R, Ronan JL, Cui K, Zhao K, Crabtree GR (2009) An

embryonic stem cell chromatin remodeling complex, esBAF, is

an essential component of the core pluripotency transcriptional

network. Proc Natl Acad Sci USA 106(13):5187–5191

Hockly E, Richon VM, Woodman B, Smith DL, Zhou X, Rosa E,

Sathasivam K, Ghazi-Noori S, Mahal A, Lowden PA, Steffan JS,

Marsh JL, Thompson LM, Lewis CM, Marks PA, Bates GP

(2003) Suberoylanilide hydroxamic acid, a histone deacetylase

inhibitor, ameliorates motor deficits in a mouse model of

Huntington’s disease. Proc Natl Acad Sci USA 100(4):

2041–2046

Hong S, Cho YW, Yu LR, Yu H, Veenstra TD, Ge K (2007)

Identification of JmjC domain-containing UTX and JMJD3 as

histone H3 lysine 27 demethylases. Proc Natl Acad Sci USA

104(47):18439–18444

Hood RL, Lines MA, Nikkel SM, Schwartzentruber J, Beaulieu C,

Nowaczyk MJ, Allanson J, Kim CA, Wieczorek D, Moilanen JS,

Lacombe D, Gillessen-Kaesbach G, Whiteford ML, Quaio CR,

Gomy I, Bertola DR, Albrecht B, Platzer K, McGillivray G, Zou

R, McLeod DR, Chudley AE, Chodirker BN, Marcadier J,

FORGE Canada Consortium, Majewski J, Bulman DE, White

SM, Boycott KM (2012) Mutations in SRCAP, encoding SNF2-

related CREBBP activator protein, cause Floating-Harbor syn-

drome. Am J Hum Genet 90(2):308–313

Horike S, Cai S, Miyano M, Cheng JF, Kohwi-Shigematsu T (2005)

Loss of silent-chromatin looping and impaired imprinting of

DLX5 in Rett syndrome. Nat Genet 37(1):31–40

Huang N, vom Baur E, Garnier JM, Lerouge T, Vonesch JL, Lutz Y,

Chambon P, Losson R (1998) Two distinct nuclear receptor

interaction domains in NSD1, a novel SET protein that exhibits

characteristics of both corepressors and coactivators. EMBO J

17(12):3398–3412

Hurd EA, Poucher HK, Cheng K, Raphael Y, Martin DM (2010) The

ATP-dependent chromatin remodeling enzyme CHD7 regulates

pro-neural gene expression and neurogenesis in the inner ear.

Development 137(18):3139–3150

Ishibashi T, Thambirajah AA, Ausio J (2008) MeCP2 preferentially

binds to methylated linker DNA in the absence of the terminal

tail of histone H3 and independently of histone acetylation.

FEBS Lett 582(7):1157–1162

Jensen LR, Amende M, Gurok U, Moser B, Gimmel V, Tzschach A,

Janecke AR, Tariverdian G, Chelly J, Fryns JP, Van Esch H,

Kleefstra T, Hamel B, Moraine C, Gecz J, Turner G, Reinhardt

R, Kalscheuer VM, Ropers HH, Lenzner S (2005) Mutations in

the JARID1C gene, which is involved in transcriptional regu-

lation and chromatin remodeling, cause X-linked mental retar-

dation. Am J Hum Genet 76:227–236

Jiang X, Zhou Y, Xian L, Chen W, Wu H, Gao X (2012) The

mutation in Chd7 causes misexpression of Bmp4 and develop-

mental defects in telencephalic midline. Am J Pathol 181(2):

626–641

Jin B, Tao Q, Peng J, Soo HM, Wu W, Ying J, Fields CR, Delmas

AL, Liu X, Qiu J, Robertson KD (2008) DNA methyltransferase

3B (DNMT3B) mutations in ICF syndrome lead to altered

epigenetic modifications and aberrant expression of genes

regulating development, neurogenesis and immune function.

Hum Mol Genet 17(5):690–709

Jin Q, Yu LR, Wang L, Zhang Z, Kasper LH, Lee JE, Wang C,

Brindle PK, Dent SY, Ge K (2011) Distinct roles of GCN5/

PCAF-mediated H3K9ac and CBP/p300-mediated H3K18/27ac

in nuclear receptor transactivation. EMBO J 30(2):249–262

Jones PL, Veenstra GJ, Wade PA, Vermaak D, Kass SU, Landsberger

N, Strouboulis J, Wolffe AP (1998) Methylated DNA and

MeCP2 recruit histone deacetylase to repress transcription. Nat

Genet 19(2):187–191

Josefowicz SZ, Wilson CB, Rudensky AY (2009) Cutting edge: TCR

stimulation is sufficient for induction of Foxp3 expression in the

absence of DNA methyltransferase 1. J Immunol 182(11):

6648–6652

Juan AH, Derfoul A, Feng X, Ryall JG, Dell’Orso S, Pasut A, Zare H,

Simone JM, Rudnicki MA, Sartorelli V (2011) Polycomb EZH2

controls self-renewal and safeguards the transcriptional identity

of skeletal muscle stem cells. Genes Dev 25(8):789–794

Kaminskas E, Farrell A, Abraham S, Baird A, Hsieh LS, Lee SL,

Leighton JK, Patel H, Rahman A, Sridhara R, Wang YC, Pazdur

R (2005) FDA, Approval summary: azacitidine for treatment of

myelodysplastic syndrome subtypes. Clin Cancer Res 11(10):

3604–3608

Kasper LH, Thomas MC, Zambetti GP, Brindle PK (2011) Double

null cells reveal that CBP and p300 are dispensable for p53

targets p21 and Mdm2 but variably required for target genes of

other signaling pathways. Cell Cycle 10(2):212–221

Kawahara Y, Mieda-Sato A (2012) TDP-43 promotes microRNAbiogenesis as a component of the Drosha and Dicer complexes.

Proc Natl Acad Sci USA 109(9):3347–3352

Kazantsev AG, Thompson LM (2008) Therapeutic application of

histone deacetylase inhibitors for central nervous system disor-

ders. Nat Rev Drug Discov 7(10):854–868

Kilgore M, Miller CA, Fass DM, Hennig KM, Haggarty SJ, Sweatt

JD, Rumbaugh G (2010) Inhibitors of class 1 histone

deacetylases reverse contextual memory deficits in a mouse

model of Alzheimer’s disease. Neuropsychopharmacology

35(4):870–880

378 Hum Genet (2013) 132:359–383

123

Kim HG, Kurth I, Lan F, Meliciani I, Wenzel W, Eom SH, Kang GB,

Rosenberger G, Tekin M, Ozata M, Bick DP, Sherins RJ, Walker

SL, Shi Y, Gusella JF, Layman LC (2008) Mutations in CHD7,

encoding a chromatin-remodeling protein, cause idiopathic

hypogonadotropic hypogonadism and Kallmann syndrome. Am

J Hum Genet 83:511–519

Kirmizis A, Bartley SM, Kuzmichev A, Margueron R, Reinberg D,

Green R, Farnham PJ (2004) Silencing of human polycomb

target genes is associated with methylation of histone H3 Lys 27.

Genes Dev 18(13):1592–1605

Kleefstra T, van Zelst-Stams WA, Nillesen WM, Cormier-Daire V,

Houge G, Foulds N, van Dooren M, Willemsen MH, Pfundt R,

Turner A, Wilson M, McGaughran J, Rauch A, Zenker M, Adam

MP, Innes M, Davies C, Lopez AG, Casalone R, Weber A,

Brueton LA, Navarro AD, Bralo MP, Venselaar H, Stegmann

SP, Yntema HG, van Bokhoven H, Brunner HG (2009) Further

clinical and molecular delineation of the 9q subtelomeric

deletion syndrome supports a major contribution of EHMT1

haploinsufficiency to the core phenotype. J Med Genet 46:

598–606

Kleefstra T, Brunner HG, Amiel J, Oudakker AR, Nillesen WM,

Magee A, Genevieve D, Cormier-Daire V, van Esch H, Fryns JP,

Hamel BC, Sistermans EA, de Vries BB, van Bokhoven H

(2006) Loss-of-function mutations in euchromatin histone

methyl transferase 1 (EHMT1) cause the 9q34 subtelomeric

deletion syndrome. Am J Hum Genet 79:370–377

Kleefstra T, Kramer JM, Neveling K, Willemsen MH, Koemans TS,

Vissers LE, Wissink-Lindhout W, Fenckova M, van den Akker

WM, Kasri NN, Nillesen WM, Prescott T, Clark RD, Devriendt

K, van Reeuwijk J, de Brouwer AP, Gilissen C, Zhou H, Brunner

HG, Veltman JA, Schenck A, van Bokhoven H (2012) Disrup-

tion of an EHMT1-Associated Chromatin-Modification Module

Causes Intellectual Disability. Am J Hum Genet 91(1):73–82

Klein CJ, Botuyan MV, Wu Y, Ward CJ, Nicholson GA, Hammans S,

Hojo K, Yamanishi H, Karpf AR, Wallace DC, Simon M,

Lander C, Boardman LA, Cunningham JM, Smith GE, Litchy

WJ, Boes B, Atkinson EJ, Middha S, Dyck B, Parisi JE, Mer G,

Smith DI, Dyck PJ (2011) Mutations in DNMT1 cause

hereditary sensory neuropathy with dementia and hearing loss.

Nat Genet 43:595–600

Kleine-Kohlbrecher D, Christensen J, Vandamme J, Abarrategui I,

Bak M, Tommerup N, Shi X, Gozani O, Rappsilber J, Salcini

AE, Helin K (2010) A Functional Link between the Histone

Demethylase PHF8 and the Transcription Factor ZNF711 in

X-Linked Mental Retardation. Mol Cell 38(2–2):165–178

Koivisto AM, Ala-Mello S, Lemmela S, Komu HA, Rautio J, Jarvela

I (2007) Screening of mutations in the PHF8 gene and

identification of a novel mutation in a Finnish family with

XLMR and cleft lip/cleft palate. Clin Genet 72:145–149

Korzus E, Rosenfeld MG, Mayford M (2004) CBP histone acetyl-

transferase activity is a critical component of memory consol-

idation. Neuron 42:961–972

Kourmouli N, Sun YM, van der Sar S, Singh PB, Brown JP (2005)

Epigenetic regulation of mammalian pericentric heterochromatin

in vivo by HP1. Biochem Biophys Res Commun 337(3):901–907

Kraft M, Cirstea IC, Voss AK, Thomas T, Goehring I, Sheikh BN,

Gordon L, Scott H, Smyth GK, Ahmadian MR, Trautmann U,

Zenker M, Tartaglia M, Ekici A, Reis A, Dorr HG, Rauch A,

Thiel CT (2011) Disruption of the histone acetyltransferase

MYST4 leads to a Noonan syndrome-like phenotype and

hyperactivated MAPK signaling in humans and mice. J Clin

Invest 121(9):3479–3491

Kramer JM, van Bokhoven H (2009) Genetic and epigenetic defects

in mental retardation. Int J Biochem Cell Biol 41(1):96–107

Kramer JM, Kochinke K, Oortveld MA, Marks H, Kramer D, de Jong

EK, Asztalos Z, Westwood JT, Stunnenberg HG, Sokolowski

MB, Keleman K, Zhou H, van Bokhoven H, Schenck A (2011)

Epigenetic regulation of learning and memory by Drosophila

EHMT/G9a. PLoS Biol 9(1):e1000569

Kurotaki N, Harada N, Yoshiura K, Sugano S, Niikawa N, Matsumoto

N (2001) Molecular characterization of NSD1, a human

homologue of the mouse Nsd1 gene. Gene 279(2):197–204

Kurotaki N, Imaizumi K, Harada N, Masuno M, Kondoh T, Nagai T,

Ohashi H, Naritomi K, Tsukahara M, Makita Y, Sugimoto T,

Sonoda T, Hasegawa T, Chinen Y, Tomita Ha HA, Kinoshita A,

Mizuguchi T, Yoshiura Ki K, Ohta T, Kishino T, Fukushima Y,

Niikawa N, Matsumoto N (2002) Haploinsufficiency of NSD1

causes Sotos syndrome. Nat Genet 30(4):365–366

Lacombe D, Saura R, Taine L, Battin J (1992) Confirmation of

assignment of a locus for Rubinstein-Taybi syndrome gene to

16p13.3. Am J Med Gent 44:126–128

Lagier-Tourenne C, Polymenidou M, Cleveland DW (2010) TDP-43

and FUS/TLS: emerging roles in RNA processing and neurode-

generation. Hum Mol Genet 19(R1):R46–R64

Lan F, Bayliss PE, Rinn JL, Whetstine JR, Wang JK, Chen S, Iwase

S, Alpatov R, Issaeva I, Canaani E, Roberts TM, Chang HY, Shi

Y (2007) A histone H3 lysine 27 demethylase regulates animal

posterior development. Nature 449(7163):689–694

Lana E, Megarbane A, Tourriere H, Sarda P, Lefranc G, Claustres M,

De Sario A (2012) DNA replication is altered in Immunodefi-

ciency Centromeric instability Facial anomalies (ICF) cells

carrying DNMT3B mutations. Eur J Hum Genet 20(10):

1044–1050

Lapunzina P (2005) Risk of tumorigenesis in overgrowth syndromes:

a comprehensive review. Am J Med Genet C Semin Med Genet

137C(1):53–71

Laugel V, Dalloz C, Durand M, Sauvanaud F, Kristensen U, Vincent

MC, Pasquier L, Odent S, Cormier-Daire V, Gener B, Tobias ES,

Tolmie JL, Martin-Coignard D, Drouin-Garraud V, Heron D,

Journel H, Raffo E, Vigneron J, Lyonnet S, Murday V, Gubser-

Mercati D, Funalot B, Brueton L, Sanchez Del Pozo J, Munoz E,

Gennery AR, Salih M, Noruzinia M, Prescott K, Ramos L, Stark

Z, Fieggen K, Chabrol B, Sarda P, Edery P, Bloch-Zupan A,

Fawcett H, Pham D, Egly JM, Lehmann AR, Sarasin A, Dollfus

H (2010) Mutation update for the CSB/ERCC6 and CSA/ERCC8

genes involved in Cockayne syndrome. Hum Mutat 31(2):

113–126

Laumonnier F, Holbert S, Ronce N, Faravelli F, Lenzner S, Schwartz

CE, Lespinasse J, Van Esch H, Lacombe D, Goizet C, Phan-

Dinh Tuy F, van Bokhoven H, Fryns JP, Chelly J, Ropers HH,

Moraine C, Hamel BC, Briault S (2005) Mutations in PHF8 are

associated with X linked mental retardation and cleft lip/cleft

palate. J Med Genet 42:780–786

Law MJ, Lower KM, Voon HP, Hughes JR, Garrick D, Viprakasit V,

Mitson M, De Gobbi M, Marra M, Morris A, Abbott A, Wilder

SP, Taylor S, Santos GM, Cross J, Ayyub H, Jones S, Ragoussis

J, Rhodes D, Dunham I, Higgs DR, Gibbons RJ (2010) ATR-X

syndrome protein targets tandem repeats and influences allele-

specific expression in a size-dependent manner. Cell 143:

367–378

Layman WS, McEwen DP, Beyer LA, Lalani SR, Fernbach SD, Oh E,

Swaroop A, Hegg CC, Raphael Y, Martens JR, Martin DM

(2009) Defects in neural stem cell proliferation and olfaction in

Chd7 deficient mice indicate a mechanism for hyposmia in

human CHARGE syndrome. Hum Mol Genet 18(11):1909–1923

Lederer D, Grisart B, Digilio MC, Benoit V, Crespin M, Ghariani SC,

Maystadt I, Dallapiccola B, Verellen-Dumoulin C (2012)

Deletion of KDM6A, a histone demethylase interacting with

MLL2, in three patients with Kabuki syndrome. Am J Hum

Genet 90:119–124

Lessard J, Wu JI, Ranish JA, Wan M, Winslow MM, Staahl BT, Wu

H, Aebersold R, Graef IA, Crabtree GR (2007) An essential

Hum Genet (2013) 132:359–383 379

123

switch in subunit composition of a chromatin remodeling

complex during neural development. Neuron 55(2):201–215

Lewis PW, Elsaesser SJ, Noh KM, Stadler SC, Allis CD (2010) Daxx

is an H3.3-specific histone chaperone and cooperates with

ATRX in replication-independent chromatin assembly at telo-

meres. Proc Natl Acad Sci USA 107(32):14075–14080

Li H, Zhong X, Chau KF, Williams EC, Chang Q (2011) Loss of

activity-induced phosphorylation of MeCP2 enhances synapto-

genesis. LTP and spatial memory. Nat Neurosci 14(8):

1001–1008

Licht CL, Stevnsner T, Bohr VA (2003) Cockayne syndrome group B

cellular and biochemical functions. Am J Hum Genet 73:

1217–1239

Ling SC, Albuquerque CP, Han JS, Lagier-Tourenne C, Tokunaga S,

Zhou H, Cleveland DW (2010) ALS-associated mutations in

TDP-43 increase its stability and promote TDP-43 complexes

with FUS/TLS. Proc Natl Acad Sci USA 107(30):13318–13323

Lioy DT, Garg SK, Monaghan CE, Raber J, Foust KD, Kaspar BK,

Hirrlinger PG, Kirchhoff F, Bissonnette JM, Ballas N, Mandel G

(2011) A role for glia in the progression of Rett’s syndrome.

Nature 475(7357):497–500

Liu L, van Groen T, Kadish I, Tollefsbol TO (2009) DNA

methylation impacts on learning and memory in aging. Neuro-

biol Aging 30(4):549–560

Liu W, Tanasa B, Tyurina OV, Zhou TY, Gassmann R, Liu WT, Ohgi

KA, Benner C, Garcia-Bassets I, Aggarwal AK, Desai A,

Dorrestein PC, Glass CK, Rosenfeld MG (2010) PHF8 mediates

histone H4 lysine 20 demethylation events involved in cell cycle

progression. Nature 466(7305):508–512

Lopez-Atalaya JP, Gervasini C, Mottadelli F, Spena S, Piccione M,

Scarano G, Selicorni A, Barco A, Larizza L (2012) Histone

acetylation deficits in lymphoblastoid cell lines from patients

with Rubinstein-Taybi syndrome. J Med Genet 49(1):66–74

Lu T, Jackson MW, Wang B, Yang M, Chance MR, Miyagi M,

Gudkov AV, Stark GR (2010) Regulation of NF-kappaB by

NSD1/FBXL11-dependent reversible lysine methylation of p65.

Proc Natl Acad Sci USA 107(1):46–51

Macrae IJ, Zhou K, Li F, Repic A, Brooks AN, Cande WZ, Adams

PD, Doudna JA (2006) Structural basis for double-stranded RNA

processing by Dicer. Science 311(5758):195–198

Mayo S, Garin I, Monfort S, Rosello M, Orellana C, Oltra S, Zazo C,

de Naclares GP, Martınez F (2012) Hypomethylation of the

KCNQ1OT1 imprinting center of chromosome 11 associated to

Sotos-like features. J Hum Genet 57(2):153–156

McMahon SJ, Pray-Grant MG, Schieltz D, Yates JR 3rd, Grant PA

(2005) Polyglutamine-expanded spinocerebellar ataxia-7 protein

disrupts normal SAGA and SLIK histone acetyltransferase

activity. Proc Natl Acad Sci USA 102(24):8478–8482

McManus KJ, Hendzel MJ (2003) Quantitative analysis of CBP- and

P300-induced histone acetylations in vivo using native chroma-

tin. Mol Cell Biol 23(21):7611–7627

Merson TD, Dixon MP, Collin C, Rietze RL, Bartlett PF, Thomas T,

Voss AK (2006) The transcriptional coactivator Querkopf

controls adult neurogenesis. J Neurosci 26(44):11359–11370

Miska EA, Langley E, Wolf D, Karlsson C, Pines J, Kouzarides T

(2001) Differential localization of HDAC4 orchestrates muscle

differentiation. Nucleic Acids Res 29(16):3439–3447

Monroy MA, Ruhl DD, Xu X, Granner DK, Yaciuk P, Chrivia JC

(2001) Regulation of cAMP-responsive element-binding protein-

mediated transcription by the SNF2/SWI-related protein.

SRCAP. J Biol Chem 276(44):40721–40726

Moretti P, Zoghbi HY (2006) MeCP2 dysfunction in Rett syndrome

and related disorders. Curr Opin Genet Dev 16(3):276–281

Morris B, Etoubleau C, Bourthoumieu S, Reynaud-Perrine S, Laroche

C, Lebbar A, Yardin C, Elsea SH (2012) Dose dependent

expression of HDAC4 causes variable expressivity in a novel

inherited case of brachydactyly mental retardation syndrome.

Am J Med Genet A 158A(8):2015–2020

Morrison AJ, Shen X (2009) Chromatin remodelling beyond

transcription: the INO80 and SWR1 complexes. Nat Rev Mol

Cell Biol 10(6):373–384

Muotri AR, Marchetto MC, Coufal NG, Oefner R, Yeo G, Nakashima

K, Gage FH (2010) L1 retrotransposition in neurons is modu-

lated by MeCP2. Nature 468(7322):443–446

Murawska M, Brehm A (2011) CHD chromatin remodelers and the

transcription cycle. Transcription 2(6):244–253

Na ES, Nelson ED, Kavalali ET, Monteggia LM (2012) The Impact

of MeCP2 Loss- or Gain-of-Function on Synaptic Plasticity.

Neuropsychopharmacology. doi:10.1038/npp.2012.116

Nagl NG Jr, Wang X, Patsialou A, Van Scoy M, Moran E (2007)

Distinct mammalian SWI/SNF chromatin remodeling complexes

with opposing roles in cell-cycle control. EMBO J 26(3):

752–763

Nakamura Y, Tagawa K, Oka T, Sasabe T, Ito H, Shiwaku H, La

Spada AR, Okazawa H (2012) Ataxin-7 associates with micro-

tubules and stabilizes the cytoskeletal network. Hum Mol Genet

21(5):1099–1110

Nan X, Hou J, Maclean A, Nasir J, Lafuente MJ, Shu X, Kriaucionis

S, Bird A (2007) Interaction between chromatin proteins MECP2

and ATRX is disrupted by mutations that cause inherited mental

retardation. Proc Nat Acad Sci USA 104:2709–2714

Neumann M, Sampathu DM, Kwong LK, Truax AC, Micsenyi MC,

Chou TT, Bruce J, Schuck T, Grossman M, Clark CM,

McCluskey LF, Miller BL, Masliah E, Mackenzie IR, Feldman

H, Feiden W, Kretzschmar HA, Trojanowski JQ, Lee VM (2006)

Ubiquitinated TDP-43 in frontotemporal lobar degeneration and

amyotrophic lateral sclerosis. Science 314(5796):130–133

Newman JC, Bailey AD, Weiner AM (2006) Cockayne syndrome

group B protein (CSB) plays a general role in chromatin

maintenance and remodeling. Proc Natl Acad Sci USA

103(25):9613–9618

Ng SB, Bigham AW, Buckingham KJ, Hannibal MC, McMillin MJ,

Gildersleeve HI, Beck AE, Tabor HK, Cooper GM, Mefford HC,

Lee C, Turner EH, Smith JD, Rieder MJ, Yoshiura K,

Matsumoto N, Ohta T, Niikawa N, Nickerson DA, Bamshad

MJ, Shendure J (2010) Exome sequencing identifies MLL2

mutations as a cause of Kabuki syndrome. Nat Genet 42:

790–793

Nguyen MV, Du F, Felice CA, Shan X, Nigam A, Mandel G,

Robinson JK, Ballas N (2012) MeCP2 Is Critical for Maintaining

Mature Neuronal Networks and Global Brain Anatomy during

Late Stages of Postnatal Brain Development and in the Mature

Adult Brain. J Neurosci 32(29):10021–10034

Nielsen AL, Jørgensen P, Lerouge T, Cervino M, Chambon P, Losson

R (2004) Nizp1, a novel multitype zinc finger protein that

interacts with the NSD1 histone lysine methyltransferase through

a unique C2HR motif. Mol Cell Biol 24(12):5184–5196

Niikawa N, Matsuura N, Fukushima Y, Ohsawa T, Kajii T (1981)

Kabuki make-up syndrome: a syndrome of mental retardation,

unusual facies, large and protruding ears, and postnatal growth

deficiency. J Pediat 99:565–569

Ogawa H, Ishiguro K, Gaubatz S, Livingston DM, Nakatani Y (2002)

A complex with chromatin modifiers that occupies E2F- and

Myc-responsive genes in G0 cells. Science 296(5570):

1132–1136

Ogryzko VV, Schiltz RL, Russanova V, Howard BH, Nakatani Y

(1996) The transcriptional coactivators p300 and CBP are

histone acetyltransferases. Cell 87:953–959

Ounap K, Puusepp-Benazzouz H, Peters M, Vaher U, Rein R, Proos

A, Field M, Reimand T (2012) A novel c.2T [ C mutation of the

KDM5C/JARID1C gene in one large family with X-linked

intellectual disability. Eur J Med Genet 55(3):178–184

380 Hum Genet (2013) 132:359–383

123

Park JH, Park EJ, Lee HS, Kim SJ, Hur SK, Imbalzano AN, Kwon J

(2006) Mammalian SWI/SNF complexes facilitate DNA double-

strand break repair by promoting gamma-H2AX induction.

EMBO J 25(17):3986–3997

Pasillas MP, Shah M, Kamps MP (2011) NSD1 PHD domains bind

methylated H3K4 and H3K9 using interactions disrupted by

point mutations in human Sotos syndrome. Hum Mutat

32(3):292–298

Petrij F, Giles RH, Dauwerse HG, Saris JJ, Hennekam RC, Masuno

M, Tommerup N, van Ommen GJ, Goodman RH, Peters DJ

(1995) Rubinstein-Taybi syndrome caused by mutations in the

transcriptional co-activator CBP. Nature 376(6538):348–351

Picketts DJ, Higgs DR, Bachoo S, Blake DJ, Quarrell OW, Gibbons

RJ (1996) ATRX encodes a novel member of the SNF2 family of

proteins: mutations point to a common mechanism underlying

the ATR-X syndrome. Hum Molec Genet 5:1899–1907

Portela A, Esteller M (2010) Epigenetic modifications and human

disease. Nat Biotechnol 28(10):1057–1068

Qi HH, Sarkissian M, Hu GQ, Wang Z, Bhattacharjee A, Gordon DB,

Gonzales M, Lan F, Ongusaha PP, Huarte M, Yaghi NK, Lim H,

Garcia BA, Brizuela L, Zhao K, Roberts TM, Shi Y (2010)

Histone H4K20/H3K9 demethylase PHF8 regulates zebrafish

brain and craniofacial development. Nature 466(7305):503–507

Qiao Y, Liu X, Harvard C, Hildebrand MJ, Rajcan-Separovic E,

Holden JJ, Lewis ME (2008) Autism-associated familial mic-

rodeletion of Xp11.22. Clin Genet 74(2):134–144

Rahman N (2005) Mechanisms predisposing to childhood overgrowth

and cancer. Curr Opin Genet Dev 15(3):227–233

Rajan I, Savelieva KV, Ye GL, Wang CY, Malbari MM, Friddle C,

Lanthorn TH, Zhang W (2009) Loss of the putative catalytic

domain of HDAC4 leads to reduced thermal nociception and

seizures while allowing normal bone development. PLoS ONE

4(8):e6612

Ravn K, Nielsen JB, Uldall P, Hansen FJ, Schwartz M (2003) No

correlation between phenotype and genotype in boys with a

truncating MECP2 mutation. J Med Genet 40(1):e5

Rayasam GV, Wendling O, Angrand PO, Mark M, Niederreither K,

Song L, Lerouge T, Hager GL, Chambon P, Losson R (2003)

NSD1 is essential for early post-implantation development and

has a catalytically active SET domain. EMBO J 22(12):

3153–3163

Reardon W (2002) Genitopatellar syndrome: a recognizable pheno-

type. Am J Med Genet 111:313–315

Reik W (2007) Stability and flexibility of epigenetic gene regulation

in mammalian development. Nature 447(7143):425–432

Reisman D, Glaros S, Thompson EA (2009) The SWI/SNF complex

and cancer. Oncogene 28(14):1653–1668

Rio Frio T, Bahubeshi A, Kanellopoulou C, Hamel N, Niedziela M,

Sabbaghian N, Pouchet C, Gilbert L, O’Brien PK, Serfas K,

Broderick P, Houlston RS, Lesueur F, Bonora E, Muljo S,

Schimke RN, Bouron-Dal Soglio D, Arseneau J, Schultz KA,

Priest JR, Nguyen VH, Harach HR, Livingston DM, Foulkes

WD, Tischkowitz M (2011) DICER1 mutations in familial

multinodular goiter with and without ovarian Sertoli-Leydig cell

tumors. JAMA 305:68–77

Roelfsema JH, White SJ, Ariyurek Y, Bartholdi D, Niedrist D,

Papadia F, Bacino CA, den Dunnen JT, van Ommen GJ,

Breuning MH, Hennekam RC, Peters DJ (2005) Genetic

heterogeneity in Rubinstein-Taybi syndrome: mutations in both

the CBP and EP300 genes cause disease. Am J Hum Genet

76(4):572–580

Ruhl DD, Jin J, Cai Y, Swanson S, Florens L, Washburn MP,

Conaway RC, Conaway JW, Chrivia JC (2006) Purification of a

human SRCAP complex that remodels chromatin by incorpo-

rating the histone variant H2A.Z into nucleosomes. Biochemistry

45:5671–5677

Samaco RC, Nagarajan RP, Braunschweig D (2004) LaSalle JM

(2004) Multiple pathways regulate MeCP2 expression in normal

brain development and exhibit defects in autism-spectrum

disorders. Hum Mol Genet 13(6):629–639

Samaco RC, Hogart A, LaSalle JM (2005) Epigenetic overlap in

autism-spectrum neurodevelopmental disorders: MECP2 defi-

ciency causes reduced expression of UBE3A and GABRB3.

Hum Mol Genet 14(4):483–492

Sanlaville D, Etchevers HC, Gonzales M, Martinovic J, Clement-Ziza

M, Delezoide AL, Aubry MC, Pelet A, Chemouny S, Cruaud C,

Audollent S, Esculpavit C, Goudefroye G, Ozilou C, Fredouille

C, Joye N, Morichon-Delvallez N, Dumez Y, Weissenbach J,

Munnich A, Amiel J, Encha-Razavi F, Lyonnet S, Vekemans M,

Attie-Bitach T (2006) Phenotypic spectrum of CHARGE

syndrome in fetuses with CHD7 truncating mutations correlates

with expression during human development. J Med Genet

43:211–217

Santen GW, Aten E, Sun Y, Almomani R, Gilissen C, Nielsen M,

Kant SG, Snoeck IN, Peeters EA, Hilhorst-Hofstee Y, Wessels

MW, den Hollander NS, Ruivenkamp CA, van Ommen GJ,

Breuning MH, den Dunnen JT, van Haeringen A, Kriek M

(2012) Mutations in SWI/SNF chromatin remodeling complex

gene ARID1B cause Coffin-Siris syndrome. Nat Genet

44:379–380

Santos-Reboucas CB, Fintelman-Rodrigues N, Jensen LR, Kuss AW,

Ribeiro MG, Campos M Jr, Santos JM, Pimentel MM (2011) A

novel nonsense mutation in KDM5C/JARID1C gene causing

intellectual disability, short stature and speech delay. Neurosci

Lett 498:67–71

Schnetz MP, Bartels CF, Shastri K, Balasubramanian D, Zentner GE,

Balaji R, Zhang X, Song L, Wang Z, Laframboise T, Crawford

GE, Scacheri PC (2009) Genomic distribution of CHD7 on

chromatin tracks H3K4 methylation patterns. Genome Res

19(4):590–601

Scuto A, Kirschbaum M, Kowolik C, Kretzner L, Juhasz A, Atadja P,

Pullarkat V, Bhatia R, Forman S, Yen Y, Jove R (2008) The

novel histone deacetylase inhibitor, LBH589, induces expression

of DNA damage response genes and apoptosis in Ph- acute

lymphoblastic leukemia cells. Blood 111(10):5093–5100

Seenundun S, Rampalli S, Liu QC, Aziz A, Palii C, Hong S, Blais A,

Brand M, Ge K, Dilworth FJ (2010) UTX mediates demethyl-

ation of H3K27me3 at muscle-specific genes during myogenesis.

EMBO J 29(8):1401–1411

Seo S, Richardson GA, Kroll KL (2005) The SWI/SNF chromatin

remodeling protein Brg1 is required for vertebrate neurogenesis

and mediates transactivation of Ngn and NeuroD. Development

132(1):105–115

Shahbazian M, Young J, Yuva-Paylor L, Spencer C, Antalffy B,

Noebels J, Armstrong D, Paylor R, Zoghbi H (2002) Mice with

truncated MeCP2 recapitulate many Rett syndrome features and

display hyperacetylation of histone H3. Neuron 35:243–254

Shiohama A, Sasaki T, Noda S, Minoshima S, Shimizu N (2003)

Molecular cloning and expression analysis of a novel gene

DGCR8 located in the DiGeorge syndrome chromosomal region.

Biochem Biophys Res Commun 304:184–190

Siderius LE, Hamel BC, van Bokhoven H, de Jager F, van den

Helm B, Kremer H, Heineman-de Boer JA, Ropers HH,

Mariman EC (1999) X-linked mental retardation associated

with cleft lip/palate maps to Xp11.3-q21.3. Am J Med Genet

85:216–220

Simpson MA, Deshpande C, Dafou D, Vissers LE, Woollard WJ,

Holder SE, Gillessen-Kaesbach G, Derks R, White SM, Cohen-

Snuijf R, Kant SG, Hoefsloot LH, Reardon W, Brunner HG,

Bongers EM, Trembath RC (2012) De novo mutations of the

gene encoding the histone acetyltransferase KAT6B cause

Genitopatellar syndrome. Am J Hum Genet 90(2):290–294

Hum Genet (2013) 132:359–383 381

123

Slupianek A, Yerrum S, Safadi FF, Monroy MA (2010) The

chromatin remodeling factor SRCAP modulates expression of

prostate specific antigen and cellular proliferation in prostate

cancer cells. J Cell Physiol 224(2):369–375

Song J, Teplova M, Ishibe-Murakami S, Patel DJ (2012) Structure-

based mechanistic insights into DNMT1-mediated maintenance

DNA methylation. Science 335(6069):709–712

Sopher BL, Ladd PD, Pineda VV, Libby RT, Sunkin SM, Hurley JB,

Thienes CP, Gaasterland T, Filippova GN, La Spada AR (2011)

CTCF regulates ataxin-7 expression through promotion of a

convergently transcribed, antisense noncoding RNA. Neuron

70:1071–1084

Stark KL, Xu B, Bagchi A, Lai WS, Liu H, Hsu R, Wan X, Pavlidis P,

Mills AA, Karayiorgou M, Gogos JA (2008) Altered brain

microRNA biogenesis contributes to phenotypic deficits in a

22q11-deletion mouse model. Nat Genet 40:751–760

Tahiliani M, Mei P, Fang R, Leonor T, Rutenberg M, Shimizu F, Li J,

Rao A, Shi Y (2007) The histone H3K4 demethylase SMCX

links REST target genes to X-linked mental retardation. Nature

447:601–605

Tao J, Hu K, Chang Q, Wu H, Sherman NE, Martinowich K, Klose

RJ, Schanen C, Jaenisch R, Wang W, Sun YE (2009)

Phosphorylation of MeCP2 at Serine 80 regulates its chromatin

association and neurological function. Proc Natl Acad Sci USA

106(12):4882–4887

Tatton-Brown K, Rahman N (2007) Sotos syndrome. Eur J Hum

Genet 15(3):264–271

Tatton-Brown K, Hanks S, Ruark E, Zachariou A, Duarte Sdel V,

Ramsay E, Snape K, Murray A, Perdeaux ER, Seal S, Loveday

C, Banka S, Clericuzio C, Flinter F, Magee A, McConnell V,

Patton M, Raith W, Rankin J, Splitt M, Strenger V, Taylor C,

Wheeler P, Temple KI, Cole T; Childhood Overgrowth Collab-

oration, Douglas J, Rahman N (2011) Germline mutations in the

oncogene EZH2 cause Weaver syndrome and increased human

height. Oncotarget 2(12):1127-1133

Thomas T, Voss AK, Chowdhury K, Gruss P (2000) Querkopf, a

MYST family histone acetyltransferase, is required for

normal cerebral cortex development. Development 127(12):

2537–2548

Tropea D, Giacometti E, Wilson NR, Beard C, McCurry C, Fu DD,

Flannery R, Jaenisch R, Sur M (2009) Partial reversal of Rett

Syndrome-like symptoms in MeCP2 mutant mice. Proc Natl

Acad Sci USA106(6):2029-2034

Tsai AC, Dossett CJ, Walton CS, Cramer AE, Eng PA, Nowakowska

BA, Pursley AN, Stankiewicz P, Wiszniewska J, Cheung SW

(2011) Exon deletions of the EP300 and CREBBP genes in two

children with Rubinstein-Taybi syndrome detected by aCGH.

Europ J Hum Genet 19:43–49

Tsurusaki Y, Okamoto N, Ohashi H, Kosho T, Imai Y, Hibi-Ko Y,

Kaname T, Naritomi K, Kawame H, Wakui K, Fukushima Y,

Homma T, Kato M, Hiraki Y, Yamagata T, Yano S, Mizuno S,

Sakazume S, Ishii T, Nagai T, Shiina M, Ogata K, Ohta T,

Niikawa N, Miyatake S, Okada I, Mizuguchi T, Doi H, Saitsu H,

Miyake N, Matsumoto N (2012) Mutations affecting components

of the SWI/SNF complex cause Coffin-Siris syndrome. Nat

Genet 44:376-378

Urdinguio RG, Lopez-Serra L, Lopez-Nieva P, Alaminos M, Diaz-

Uriarte R, Fernandez AF, Esteller M (2008) Mecp2-null mice

provide new neuronal targets for Rett syndrome. PLoS ONE

3(11):e3669

Urdinguio RG, Sanchez-Mut JV, Esteller M (2009) Epigenetic

mechanisms in neurological diseases: genes, syndromes, and

therapies. Lancet Neurol 8(11):1056–1072

Valinluck V, Tsai HH, Rogstad DK, Burdzy A, Bird A, Sowers LC

(2004) Oxidative damage to methyl-CpG sequences inhibits the

binding of the methyl-CpG binding domain (MBD) of methyl-

CpG binding protein 2 (MeCP2). Nucleic Acids Res

32(14):4100–4108

Van Esch H, Bauters M, Ignatius J, Jansen M, Raynaud M, Hollanders

K, Lugtenberg D, Bienvenu T, Jensen LR, Gecz J, Moraine C,

Marynen P, Fryns JP, Froyen G (2005) Duplication of the

MECP2 region is a frequent cause of severe mental retardation

and progressive neurological symptoms in males. Am J Hum

Genet 77(3):442–453

Van Houdt JK, Nowakowska BA, Sousa SB, van Schaik BD,

Seuntjens E, Avonce N, Sifrim A, Abdul-Rahman OA, van den

Boogaard MJ, Bottani A, Castori M, Cormier-Daire V, Deardorff

MA, Filges I, Fryer A, Fryns JP, Gana S, Garavelli L, Gillessen-

Kaesbach G, Hall BD, Horn D, Huylebroeck D, Klapecki J,

Krajewska-Walasek M, Kuechler A, Lines MA, Maas S,

Macdermot KD, McKee S, Magee A, de Man SA, Moreau Y,

Morice-Picard F, Obersztyn E, Pilch J, Rosser E, Shannon N,

Stolte-Dijkstra I, Van Dijck P, Vilain C, Vogels A, Wakeling E,

Wieczorek D, Wilson L, Zuffardi O, van Kampen AH, Devriendt

K, Hennekam R, Vermeesch JR (2012) Heterozygous missense

mutations in SMARCA2 cause Nicolaides-Baraitser syndrome.

Nat Genet 44:445–449

Vecsey CG, Hawk JD, Lattal KM, Stein JM, Fabian SA, Attner MA,

Cabrera SM, McDonough CB, Brindle PK, Abel T, Wood MA

(2007) Histone deacetylase inhibitors enhance memory and

synaptic plasticity via CREB: CBP-dependent transcriptional

activation. J Neurosci 27:6128–6140

Villard L (2007) MECP2 mutations in males. J Med Genet 44:

417–423

Viosca J, Lopez-Ayala JP, Olivares R, Eckner R, Barco A (2010)

Syndromic features and mild cognitive impairment in mice with

genetic reduction on p300 activity: differential contribution of

p300 and CBP to Rubinstein-Taybi syndrome etiology. Neuro-

biol Dis 37:186–194

Wang J, Weaver IC, Gauthier-Fisher A, Wang H, He L, Yeomans J,

Wondisford F, Kaplan DR, Miller FD (2010) CBP histone

acetyltransferase activity regulates embryonic neural differenti-

ation in the normal and Rubinstein-Taybi syndrome brain. Dev

Cell 18(1):114–125

Watson P, Black G, Ramsden S, Barrow M, Super M, Kerr B,

Clayton-Smith J (2001) Angelman syndrome phenotype associ-

ated with mutations in MECP2, a gene encoding a methyl CpG

binding protein. J Med Genet 38:224–228

Weaver DD, Graham CB, Thomas IT, Smith DW (1974) A new

overgrowth syndrome with accelerated skeletal maturation,

unusual facies, and camptodactyly. J Pediat 84:547–552

Wegorzewska I, Bell S, Cairns NJ, Miller TM, Baloh RH (2009)

TDP-43 mutant transgenic mice develop features of ALS and

frontotemporal lobar degeneration. Proc Natl Acad Sci USA

106(44):18809–18814

Willemsen MH, Vulto-van Silfhout AT, Nillesen WM, Wissink-

Lindhout WM, van Bokhoven H, Philip N, Berry-Kravis EM,

Kini U, van Ravenswaaij-Arts CM, Delle Chiaie B, Innes AM,

Houge G, Kosonen T, Cremer K, Fannemel M, Stray-Pedersen

A, Reardon W, Ignatius J, Lachlan K, Mircher C, Helderman van

den Enden PT, Mastebroek M, Cohn-Hokke PE, Yntema HG,

Drunat S, Kleefstra T (2012) Update on Kleefstra Syndrome.

Mol Syndromol 2(3–5):202–212

Williams SR, Aldred MA, Der Kaloustian VM, Halal F, Gowans G,

McLeod DR, Zondag S, Toriello HV, Magenis RE, Elsea SH

(2010) Haploinsufficiency of HDAC4 causes brachydactyly

mental retardation syndrome, with brachydactyly type E, devel-

opmental delays, and behavioral problems. Am J Hum Genet

87:219–228

Winkelmann J, Lin L, Schormair B, Kornum BR, Faraco J, Plazzi G,

Melberg A, Cornelio F, Urban AE, Pizza F, Poli F, Grubert F,

Wieland T, Graf E, Hallmayer J, Strom TM, Mignot E (2012)

382 Hum Genet (2013) 132:359–383

123

Mutations in DNMT1 cause autosomal dominant cerebellar

ataxia, deafness and narcolepsy. Hum Mol Genet 21(10):

2205–2210

Wolff D, Endele S, Azzarello-Burri S, Hoyer J, Zweier M, Schanze I,

Schmitt B, Rauch A, Reis A, Zweier C (2012) In-frame deletion

and missense mutations of the C-terminal helicase domain of

SMARCA2 in three patients with Nicolaides-Baraitser syn-

drome. Molec Syndromol 2:237–244

Wong MM, Cox L, Chrivia J (2007) The chromatin remodeling

protein SRCAP is critical for deposition of the histone variant

H2A.Z at promoters. J Biol Chem 282:26132–26139

Wong LH, McGhie JD, Sim M, Anderson MA, Ahn S, Hannan RD,

George AJ, Morgan KA, Mann JR, Choo KH (2010) ATRX

interacts with H3.3 in maintaining telomere structural integrity

in pluripotent embryonic stem cells. Genome Res 20(3):351–360

Wyngaarden LA, Delgado-Olguin P, Su IH, Bruneau BG, Hopyan S

(2011) Ezh2 regulates anteroposterior axis specification and

proximodistal axis elongation in the developing limb. Develop-

ment 138(17):3759–3767

Xie W, Ling T, Zhou Y, Feng W, Zhu Q, Stunnenberg HG, Grummt I,

Tao W (2012) The chromatin remodeling complex NuRD

establishes the poised state of rRNA genes characterized by

bivalent histone modifications and altered nucleosome positions.

Proc Natl Acad Sci USA 109(21):8161–8166

Xu GL, Bestor TH, Bourc’his D, Hsieh CL, Tommerup N, Bugge M,

Hulten M, Qu X, Russo JJ, Viegas-Pequignot E (1999)

Chromosome instability and immunodeficiency syndrome

caused by mutations in a DNA methyltransferase gene. Nature

402:187–191

Xu J, Deng X, Disteche CM (2008) Sex-specific expression of the

X-linked histone demethylase gene Jarid1c in brain. PLoS ONE

3(7):e2553

Yang X, Noushmehr H, Han H, Andreu-Vieyra C, Liang G, Jones PA

(2012) Gene reactivation by 5-aza-20-deoxycytidine-induced

demethylation requires SRCAP-mediated H2A.Z insertion to

establish nucleosome depleted regions. PLoS Genet 8(3):e1002604

Yasui DH, Peddada S, Bieda MC, Vallero RO, Hogart A, Nagarajan

RP, Thatcher KN, Farnham PJ, Lasalle JM (2007) Integrated

epigenomic analyses of neuronal MeCP2 reveal a role for long-

range interaction with active genes. Proc Natl Acad Sci USA

104(49):19416–19421

Yazdani M, Deogracias R, Guy J, Poot RA, Bird A, Barde YA (2010)

Disease modeling using embryonic stem cells: MeCP2 regulates

nuclear size and RNA synthesis in neurons. Stem Cells

30(10):2128–2139

Yehezkel S, Segev Y, Viegas-Pequignot E, Skorecki K, Selig S

(2008) Hypomethylation of subtelomeric regions in ICF syn-

drome is associated with abnormally short telomeres and

enhanced transcription from telomeric regions. Hum Mol Genet

17(18):2776–2789

Yoneda Y, Saitsu H, Touyama M, Makita Y, Miyamoto A, Hamada

K, Kurotaki N, Tomita H, Nishiyama K, Tsurusaki Y, Doi H,

Miyake N, Ogata K, Naritomi K, Matsumoto N (2012) Missense

mutations in the DNA-binding/dimerization domain of NFIX

cause Sotos-like features. J Hum Genet 57(3):207-211

Yoo SY, Pennesi ME, Weeber EJ, Xu B, Atkinson R, Chen S,

Armstrong DL, Wu SM, Sweatt JD, Zoghbi HY (2003) SCA7

knockin mice model human SCA7 and reveal gradual accumu-

lation of mutant ataxin-7 in neurons and abnormalities in short-

term plasticity. Neuron 37(3):383–401

Young JI, Hong EP, Castle JC, Crespo-Barreto J, Bowman AB, Rose

MF, Kang D, Richman R, Johnson JM, Berget S, Zoghbi HY

(2005) Regulation of RNA splicing by the methylation-depen-

dent transcriptional repressor methyl-CpG binding protein 2.

Proc Natl Acad Sci USA102(49):17551-17558

Yu L, Wang Y, Huang S, Wang J, Deng Z, Zhang Q, Wu W, Zhang

X, Liu Z, Gong W, Chen Z (2010) Structural insights into a

novel histone demethylase PHF8. Cell Res 20(2):166–173

Yuan X, Feng W, Imhof A, Grummt I, Zhou Y (2007) Activation of

RNA polymerase I transcription by cockayne syndrome group B

protein and histone methyltransferase G9a. Mol Cell

27(4):585–595

Zachariah RM, Rastegar M (2012) Linking epigenetics to human

disease and Rett syndrome: the emerging novel and challenging

concepts in MeCP2 research. Neural Plast 2012:415825

Zhou Z, Hong EJ, Cohen S, Zhao WN, Ho HY, Schmidt L, Chen WG,

Lin Y, Savner E, Griffith EC, Hu L, Steen JA, Weitz CJ,

Greenberg ME (2006) Brain-specific phosphorylation of MeCP2

regulates activity-dependent Bdnf transcription, dendritic

growth, and spine maturation. Neuron 52(2):255–269

Hum Genet (2013) 132:359–383 383

123