epigenetics: an overview - skemman · 1 epigenetics: an overview cristina hernández rollán 12...

54
Epigenetics: an overview Cristina Hernández Rollán Líf- og umhverfisvísindadeild Háskóli Íslands 2014

Upload: lamquynh

Post on 11-May-2018

213 views

Category:

Documents


1 download

TRANSCRIPT

Epigenetics: an overview

Cristina Hernández Rollán

Líf- og umhverfisvísindadeild

Háskóli Íslands 2014

1

Epigenetics: an overview

Cristina Hernández Rollán

12 eininga ritgerð sem er hluti af LÍF219F Námskeið til meistaraprófs í líffræði

Leiðbeinandi

Zophonías O. Jónsson (Ph.D)

Líf- og umhverfisvísindadeild

Háskóli Íslands

2014

2

Epigenetics: an overview

12 eininga ritgerð sem er hluti af LÍF219F Námskeið til meistaraprófs í líffræði

Höfundarréttur © 2014 Cristina Hernández Rollán

Öll réttindi áskilin

Líf- og umhverfisvísindadeild

Háskóli Íslands

Askja, Sturlugatu 7

101, Reykjavik

Iceland

Sími: 525 4000

Skráningarupplýsingar:

Cristina Hernández Rollán, 2014, Epigenetics: an overview, LÍF219F Námskeið til

meistaraprófs í líffræði, Líf- og umhverfisvísindadeild, Háskóli Íslands.

Reykjavík, Janúar 2014

3

TABLE OF CONTENTS

1. INTRODUCTION ............................................................................................................................................... 5

2. CHROMATIN STRUCTURE ........................................................................................................................... 8

3. DNA METHYLATION ................................................................................................................................... 10

3.1. DNA methyltransferases in mammals:................................................................................. 12

3.2. CpG islands ........................................................................................................................... 13

4. HISTONE MODIFICATIONS ...................................................................................................................... 15

5. NON-CODING RNAs AS EPIGENETICS REGULATORS .................................................................... 18

6. THE EPIGENETIC CODE ............................................................................................................................. 20

7. X-CHROMOSOME INACTIVATION IN MAMMALS ........................................................................... 21

8. STEM CELLS .................................................................................................................................................... 23

9. IMPRINTED GENES ...................................................................................................................................... 24

9.1. Prader Willi syndrome .......................................................................................................... 26

9.2. Angelman syndrome ............................................................................................................ 27

10. SOME OTHER SYNDROMES CAUSED BY EPIGENETIC MISFUNCTION .................................. 28

10.1. Rett syndrome ...................................................................................................................... 28

10.2. Fragile-X syndrome .............................................................................................................. 28

11. MODEL SYSTEMS TO STUDY EPIGENETICIS .................................................................................... 29

12. PLANT EPIGENETICS .................................................................................................................................. 30

12.1. Arabidopsis thaliana ............................................................................................................ 30

13. INSECT EPIGENETICS ................................................................................................................................. 31

13.1. The Polycomb system ........................................................................................................... 32

14. YEAST AND EPIGENETICS ........................................................................................................................ 33

15. CLINICAL APPLICATIONS OF EPIGENETICS ..................................................................................... 34

15.1. Mitochondria in epigenetics ................................................................................................. 34

15.2. Bacterial and virus infections ............................................................................................... 34

16. CANCER AND EPIGENETICS ..................................................................................................................... 37

16.1. Epigenetic therapy ............................................................................................................... 37

17. EPIGENETIC RESEARCH METHODOLOGY ......................................................................................... 38

17.1. High resolution tiling oligonucleotides arrays: ..................................................................... 38

17.2. Shotgun bisulfite sequencing: .............................................................................................. 38

17.3. Digestion with restriction endonucleases: ........................................................................... 38

17.4. Chromatin immunoprecipitation followed by CHIP-sequencing: ......................................... 39

17.5. MeDIP and next generation sequencing: ............................................................................. 39

17.6. Cobra assays (Combined Bisulfite Restricted Assay): ........................................................... 39

18. CONCLUDING REMARKS ........................................................................................................................... 40

19. REFERENCES .................................................................................................................................................. 42

4

LIST OF FIGURES

Figure 1: Waddingston’s Epigenetics Landscape………………………………………………………………… ……………………5

Figure 2: Representation of the chromatin organization in the nucleus. ..................................................... 9

Figure 3: CpG methylation reaction............................................................................................................ 11

Figure 4: Scheme of de novo and maintenance methylation in mammals................................................. 12

Figure 5: Representation of the total percentage of methylation seen within gene bodies against gene expression .......................................................................................................................................... 14

Figure 6: Nucleosome modifications .......................................................................................................... 15

Figure 7: Representation of chromatin regulation during transcription .................................................... 17

Figure 8: Illustration showing some examples of non-coding RNA roles in histone modifications in different kingdoms…………………………………………………………………………………………………………………………………19

Figure 9: Histone modifications at the aminotails can define short-term and long-term output .............. 21

Figure 10: X-chromosome inactivation ....................................................................................................... 22

Figure 11: Imprinting clusters in human and mouse genomes ................................................................. 251

Figure 12: Picture showing facial phenotypic characteristics from Prader Willis syndrome ...................... 26

Figure 13: Picture showing the typical facial characteristics of Angelman syndrome ............................... 27

Figure 14: Picture of Arabidopsis thaliana ................................................................................................. 30

Figure 15: Schematic description of the increase methylation experience by infected lymphocytes T by the HIV virus……………………………………………………………………………………………………………………………………………….. 35

Figure 16: Schematic representation of some approaches used for the detection of epigenetic marks……………………………………………………………………………………………………………………………………………………. 40

LIST OF TABLES

Table 1. Human DNA methyltransferases…………...………...........……………………………………..……………………....9

Table 2. Histone modifications associated with transcription………………………...……………………………………..12

Table 3. Bacterial species, bacterial factors and effects……………………………………………………………………..….32

5

1.INTRODUCTION

The publication of the human genome in Science and Nature on February 16th, 2001

in many ways brought with it more questions than answers. We are now aware of how

the human genome is organized, but we have yet to understand how all the genomic

components interrelate to drive gene expression, inheritance, regulation, etc. DNA

sequences alone give us very incomplete knowledge about functions of the genome or

how it directs cell development.

Cells respond in different ways to external stimuli, they gain and retain responses to

these changes as they develop and are able to transfer this information as changes that

occur within the chromatin to the daughter cells (Ashe et al., 2012). Responses that are

potentially inheritable without altering of the DNA sequence are called epigenetic

mechanisms. Epigenetic modifications can be further classified into three major groups:

DNA methylation and hydroxymethylation, posttranslational modifications of histone

tails and regulation of gene expression by non-coding RNAs (ncRNAs) (Gonzalo 2010;

Minocherhomji et al., 2012).

Many external factors such as temperature, viral infections, bacteria and diet for

example, can affect the chromatin organization and structure, driving the cells to

respond to them by epigenetic changes (Abraham et al., 2012; Bierne et al., 2012) and

impacting the phenotype without disruption of the genotype.

The term epigenetic was first introduced by Conrad Waddington in the 1940s.

Although he did not believe that some external factors could be inherited without

changes in the DNA sequences, he was aware of the interactions of the environment

with inheritance. He coined the term “epigenetic landscape”.

Figure 1: Waddingston’s Epigenetics Landscape. The figure represents the complex

process of cellular development where the ball represents the cell and the slopes (lines)

represent metaphorically the possible pathways to follow. Every possibility will give

different outcomes. Figure adapted from Waddington, 1957 adapted from Goldberg et

al., 2007.

6

The discovery of the epigenome and how the epigenetic changes can be

phenotypically inherited over generations have given rise to a new area of study in

biology. It is now clear that epigenetic changes influence cell fate and development, and

studying them will help understanding more complex areas of biology such as

embryogenesis and development. In addition, the study of epigenetics can also

contribute understanding of many diseases such as some types of cancer, immune

disorders, even bacterial and viral infections that have eluded old fashioned approaches

of understanding genetics and inheritance (Rodríguez et al., 2003; Gal-Yam et al., 2007;

Paschos and Allday, 2010; Bierne et al., 2012).

The father of epigenetics Conrad Waddington proposed the name epigenetics based

on the Aristotelian theory of epigenesis in his book “On the generation of animals”. He

believed that environmental changes can gradually influence hereditability in an animal.

Waddington choose the term “epi” that means over, upon or above and “genetics” to

infer that genes take part in this process. As illustrated in figure 1, he tried to visualize

and explain that there is no simple relationship between the cell and gene expression

and the possible outcomes. Epigenetics is understood as the result of many cascades of

interactions throughout the process that can lead to many possible outcomes.

Consequently, Conrad Waddington defined epigenetics as “the branch of biology which

studies the causal interactions between genes and their products, which bring the

phenotype into being” (Waddington, 1947). For example, in nature it can be appreciated

the fact that having the same genotype does not always lead to the same outcome. For

example, most of the cells in the human body share the same genotype but clearly do

not all have the same phenotype (e.g. skin cells are quite different from the lungs cells

and they maintain the same phenotype throughout divisions).

Some old experiments in frogs (Briggs and King, 1952) and some more recent in

mammals (Wilmut et al., 1997 and Wakayama et al., 1998), demonstrated the basic

principle that the nucleus of a differentiated cell contains the complete genetic

information for a new cell/organism and can, at least in theory, be reprogrammed from

scratch (for review see McCarrey 2011). The phenotypic difference observed in

multicellular organisms is explained by the term plasticity. Plasticity in this cellular

context is understood as the ability of a cell to differentiate or develop into every cell

type of that organism (a particular characteristic of stem cells) (Price et al., 2003).

Plasticity is easily appreciated in colonies of honeybees where the development of the

queen bee is quite different from the development of the worker bees (although they

share exactly the same genetic information) and it is solely dependent on the way they

have been fed at larvae stage. This difference in phenotype is a plasticity example

brought on by environmental factors and mediated through epigenetics rather than

differences in genotype.

Modern usage of the term epigenetics differs significantly from Waddington’s

original definition. In 2001 it was redefined in Science as “The study of changes in gene

function that are mitotically and/or meiotically heritable and that do not entail a change

in DNA sequence” (Wu et al., 2001). Nevertheless, the definition and term are still

nowadays under continuous consideration and discussion (Reviewed in Jablonka and

Lamb 2002).

7

An obvious deduction from Waddington’s theory that has since interested the

scientific community is the possible involvement of epigenetics in evolution. Do

epigenetics have any theoretical implications for the neo-Darwinian theory of

evolution? Given the fact that epigenetic modifications occurring from environmental

adaptations could be inherited between generations (Ashe et al., 2012) maybe Lamark

was not so wrong after all.

The scientific community disagrees when it comes to epigenetic inheritability. It is

clear that epigenetic changes are decisive for differentiation in an organism, but for

many specialists it is consider unlikely that this heritability of phenotype is frequently

transmitted in a transgenerational manner. Nevertheless, this statement is not true for

some species such Caenorhabditis elegans (Henderson & Jacobsen, 2007; Ashe et al.,

2012), and it is accepted nowadays that heritable epigenetic variation can lead to

evolutionary changes (Jablonka and Lamb, 2002; McCarrey 2011; Ashe et al., 2012;

Chong and Whitelaw, 2004).

It has also been identified that when the chromatin modifications are inherited, they

form an epigenetic memory (Bird 2007; Paschos and Allday 2010). Furthermore,

epigenetics is not only important or crucial for differentiation of tissues and organs,

since the epigenetic mechanisms have also been found in unicellular organisms. Here,

epigenetics play a different role where the environment is a major issue in epigenetics

inheritance (Jablonka and Lamb, 2002; Ashe et al., 2012). Therefore, the environmental

component also has to be taken into account in evolution and adaptation; as Lamark

once stated, “environment is a stimulator in addition to a selector of variation and

specialization”. There are some examples where organisms such Caenorhabditis

elegans for instance, can under certain conditions inherit epigenetic modifications

through several generations at a selective advantage (see Ashe et al., 2012; Jablonka et

al., 1995).

A study from the University of Michigan in 2013 showed that tadpoles grow larger

tails when they are under stress because of the nearby presence of predators, such as

dragonfly larvae. By these means, they achieve an advantage over tadpoles with smaller

tails, since they can swim faster and hence, they are more difficult prey for the

dragonfly. It has been proposed that this is due to plasticity responses to the

environment that allows this body change (Maher et al., 2013).

Another example comes from a study by Chandler and Stam in 2004; they identified

cases where paramutation (explained later) is affected by environmental temperature.

The change in temperature alters the transitions between the active and inactive forms

of an allele (Mikula, 1995). Thus, paramutation and paramutation-like events supply a

mechanism in which gene expression can change rapidly due to adaptation.

8

2.CHROMATIN STRUCTURE

Understanding how the chromatin is organized and knowing how proteins interact

with the DNA to form this peculiar structure is vital to the study and comprehension of

the mechanisms behind epigenetics. In eukaryotes, chromosomal DNA is organized

inside the nucleus of the cell. Condensed DNA and associated proteins constitute the

chromatin (Bernstein et al., 2007). The organization of chromatin takes place at several

levels. The first stage consists of approximately 147 base pairs of double stranded DNA

wrapping approximately 1.65 times around the core of a histone octamer. This

represents the basic level of organization. The histones that form the octamer are two

copies of each H2A, H2B, H3 and H4. The next level of organization involves the

nucleosomes aligning along the DNA and a fifth type of histone (H1 and isoforms)

which serves as a linker between them. H1, H1-isoforms, RNAs and other non-histones

proteins assist in a higher level of condensation forming the 30nm fiber. Scaffold

proteins together with chromatin remodeling complexes maintain the chromatin

structure and allow dynamic movements from euchromatin (easily accessible DNA) to

heterochromatin (DNA that is difficult to access) (Bierne et al., 2012). Euchromatin is

the loosely packed chromatin where active genes are found transcribed, whereas the

heterochromatin contains highly compact DNA in two different varieties, facultative

and constitutive heterochromatin. Note that the core histones (H2A, H2B, H3 and H4)

are the histones found in the S-phase of the cell cycle and for a long time they were

considered the only core histones (Kornberg and Lorch, 1999). Later studies revealed

that many other histone variants are found outside the S-phase in various organisms and

are introduced into the chromatin in a DNA-replication independent manner (review in

Kamakaka & Biggins, 2005). They are referred as non-canonical histones. Differences

between the canonical and non-canonical histones are found in the histone tails or

between key amino acid residues (Bönisch & Hake, 2012).

Histones are characterized by long C or N-terminal tails subject to several post-

translational modifications which in turn are connected with biological functions. Post-

translational modifications in histone tails are diverse and often the same type of

chemical modification can result in different responses. Such observation indicates that

cross-talk must exist between histone modifications (see figure 6). Cross-talk is defined

as the influence of post-translational modifications on the same or separate histones tails

that can result in different signaling cascades leading to chromatin changes. For

instance, methylation on H3-K9 seems to cause inhibition of acetylation of residues

K14, K18, and K23 in the H3 tail, thereby preventing transcription (Fischle et al.,

2003).

9

Figure 2: Representation of the chromatin organization in the nucleus. The DNA is

condensate into chromatin, firstly organized into nucleosomes (11nm diameter) with

histones H2A, H2B, H3 and H4. The second level of organization is set when the

histone H1 binds nucleosomes into the fiber of 30nm of diameter. Higher level of

condensation is set by the scaffold proteins that grant different levels of condensations

or loops to allow or repress transcription. Figure from CliffsNotes, Structure of

Chromatin, 2013.

The chromatin not only functions to keep the DNA tight up together inside the

nucleus, but also to regulate gene expression and regulate DNA processes such as

transcription, replication, recombination, DNA repair, DNA control, formation of the

chromosomes, mitosis, meiosis, etc.

As explained before, the definition of epigenetics includes every mechanism that can

control gene expression in a potentially inherited way without a change in the DNA

sequence. Thus, DNA can be modified without disruption of its sequence by epigenetic

marks that alter the chromatin structure and the affinity of proteins to chromatin

(Miranda, & Jones 2007). But, how do these changes work? What makes them

potentially inheritable? To answer these questions, we need to look at the modifications

that the chromatin and DNA undergo in order to determine inheritable changes in gene

expression.

10

3.DNA METHYLATION

Epigenetic marks at DNA level include methylation and hydroxymethylation of the

cytosine bases. For instance, in mammals, cytosines preceding guanines (CpG) are

known to be susceptible to methylation and almost all CpGs in the genome are found

methylated (Bird, 2002; Goll and Bestor, 2005). However, only 3% of the cytosines in

the human genome are methylated. In addition to mCpG a small fraction is found as

mCpNpG, where N can be any nucleotide (Hum Mol Gen, 4th

Edition).

Methylation of the 5 C of the cytosines has a crucial role for the survival and the

correct function of the cell and is found in almost all vertebrates. However, cytosines

that are methylated are chemically unstable and more prone to deamination.

Deamination of methylated cytosines results in conversion to the nucleotide thymine

which is a normal base not recognized by the DNA repair machinery and consequently,

over time, CpG sequences in highly methylated regions are converted to TpG unless the

presence of C is selected for. Spontaneous deamination of non-methylated cytosines

gives rise to uracil that is an abnormal base in DNA and is recognized and efficiently

removed by the DNA repair machinery (see Weiner and Toth, 2011; Hum Mol Gen, 4th

Edition). The conversion of CpG dinucleotides to TpG dinucleotides has been analysed

by the Human Genome Sequencing Consortium which have concluded that CpG

dinucleotides in the human genome only occur at 21% of the expected frequency

(International Human Genome Sequencing Consortium, et al. 2001).

Methylated DNA has been demonstrated to attract methyl-DNA-binding proteins that

in turn recruit chromatin modifying complexes which condensate the chromatin around

the methylated DNA silencing the region to a repressive state (Bird and Wolffe, 1999;

Urnov and Wolffe, 2001).

11

Figure 3: CpG methylation reaction. A) Representation of a CpG islands where the

cytosine is preceded by a guanine base in a DNA sequence. B) Methylation reaction of

CpG islands. The reaction is catalyzed by a DNA methyltransferases (DNMTs) that use

the methyl group from S-adenosyl methionine (SAM-CH3) and transfer it to the cytosine

base. Figure from Maxwell et al., 2009.

Cytosine methylation (along with the derived variant hydroxymethylation) is the

only covalent DNA modification known in mammals and provides heritable epigenetic

information that is not encoded in the DNA sequence (Bernstein et al., 2007). Passage

of heritable information through cell division is possible because the CpG methylation

is not reset but copied to the nascent DNA strands.

Covalent modifications of the protein component of chromatin can be classified

according to the function of the enzymes that carry out such modification. There are

basically two main functions, adding or removing covalent modifications. Examples of

enzymes that add the covalent modifications are histone methyltansferases (HMTs),

histone acetyltransferases, kinases, etc. On the other hand, enzymes that remove these

modifications are histone demethylases (HDMs), histone deacetylases (HDAC), and

phosphatases (see Bierne et al., 2012). Histone deacetylases catalyse deacetylation of

acetylated lysine residues in histone tails. They are divided into HDAC class I, II and III

where class I and II are Zinc-dependent enzymes and class III is NAD-dependent

(Paschos and Allday 2010).

12

3.1. DNA methyltransferases in mammals:

In humans, DNA methyltransferases are classified into de novo methyltransferases

and maintenance methyltransferases. The de novo DNA methyltransferases in humans

are named DNMT3A and DNMT3B and the single maintenance DNA

methyltransferase is known as DNMT1 (Bestor, 2000; and Okano et al., 1999).

De novo DNA methyltransferases carry out the methylation reaction by the transfer

of a methyl group (-CH3) from S adenosyl-L-methionine to the carbon 5 of the cytosine

(Doerfler, 1983) (see figure 4).

There is another methyltransferase in humans known as DNMT3L that allows

targeting to specific sequences. Finally, a related enzyme named TRDMT1 has similar

structure to the other DNA methyltransferase and was therefore classified as such and

named DNMT2. It has however shown to be RNA-methylating enzyme and has

therefore been renamed (Human Molecular Genetics, Garland Science, 4th

Edition).

Figure 4: A schematic figure of de novo and maintenance methylation in mammals.

DNMT3a and DNMT3b introduce methyl groups to DNA in cytosine bases preceding

guanines. DNMT1 maintain the state of methylation when replication occurs. It follows

the pattern of the original strand and copies it to the daughter strand. Picture adapted

from Reik and Walter 2001.

13

Maintenance of the CpG methylation is only possible thanks to the enzyme DNMT1

recognizes the newly synthesized DNA strand (recently methylated) and methylates

according to the pattern on the parent strand. Recent studies have shown that DNMT

methyltransferases are brought to specific location of the genome through interactions

with proteins such as the polymerase clamp PCNA, or transcription factors suggested to

play a role in repression of transcription (see Rodríguez et al., 2004).

Table 1: Human DNA methyltransferases. Adapted from Human Molecular Genetics.

Enzyme OMIM

no.

Major functions Associated proteins

DNMT1 126375 Maintenance methylase PCNA, HP1, methyl-DNA-

binding proteins

DNMT2 602478 Methylation of cytosine-38 in tRNA, no

DNA methylation activity

DNMT3A 602769 De novo methylase Histone methyltransferases;

histone deacetylases

DNMT3B 602900 De novo methylase Histone methyltransferases;

histone deacetylases

DNMT3L 606588 Binds to chromatin with unmethylated H3K4

and stimulates activity of DNMT3A/3B

DNMT3A, DNMT3B; histone

deacetylases

3.2. CpG islands

Even though the frequency of the CpG dinucleotides in the genome is low, there are

regions of CpG dinucleotides that are rarely or never methylated. These regions, named

CpG islands, are rich in CG content and have CpG frequency that would be expected by

chance, unlike the rest of the genome. Furthermore, they also contain about 7% of all

CpG dinucleotides in the genome (see Bernstein et al., 2007).

Gardiner-Garden and Frommer defined CpG islands as “regions of at least 200 bases

with a (C+G) content of at least 50% and a ratio of observed CpG frequency to be

expected of at least 0.6” (Gardiner-Garden and Frommer, 1987). Methylation is found

in gene bodies, introns and exons as well as in intergenic sequences with the CpG

dinucleotide being the least common dinucleotide (Hum Mol Gen, 4th

Edition) while on

the other hand, CpG islands are found at gene promoters at high frequency, without

methylation, which allows gene expression. It has also been found that CpG islands are

differently methylated depending on the developmental stage of the organism that they

are found in (Fazzari and Greally, 2004). For instance, globin genes, where the CpG

islands are demethylated in all tissues of the adult organism are found silenced during

embryonic development (Bird et al., 1987). Additionally, differential CpG island

methylation is also present in imprinted genes.

Interestingly, DNA methylation is found to be more frequent within gene-bodies than

in promoter regions, and somewhat surprisingly, methylation in gene-bodies appears to

be positively correlated with gene expression in contrast to the repressive effect that

promoter methylation has on transcription. The reasons for this paradox have yet to be

14

resolved, and contrary to what was firstly thought, human genome-wide methylation

studies have shown that genes with middle level of expression exhibit the highest levels

of methylation. However, genes with low and high gene expression levels show low

methylation patterns as illustrated by the bell-shaped plots in figure 5 (Jjingo et al.,

2012).

Figure 5: Representation of the total percentage of methylation seen within gene bodies

against gene expression. Graph A shows a human lymphoblastoid cell line (GM12878).

Graph B is from a human myelogenous leukemia line (K562) and Graph C is from a

liver hepatocellular cell line (HepG2). A positive correlation between gene expression

and percentage of methylation exists until a maximum where methylation starts to

decrease (bell-shape correlation). Adapted from Jjingo et al., 2012.

15

4.HISTONE MODIFICATIONS

In addition to DNA cytosine methylation, histones can be epigenetically modified.

Histone modifications are defined as the second level of epigenetic marks that are

achieved by covalent modifications. Like DNA methylation these modifications can be

inherited through successive cell-divisions.

Chemical histone modifications occur at the amino-terminal tails of histone proteins.

Histones are modified by specific enzymes and modifications include: acetylation of

lysines, methylation of lysines and arginines, phosphorylation of serines and threonines,

ADP ribosylation of lysines, ubiquitination of lysines, and sumoylation of lysines (see

figure 4). This list is however by no means exhaustive and it is not unlikely that yet

undescribed types of covalent histone modifications remain to be discovered.

Figure 6: Nucleosome modifications. All eight histones of the nucleosome have N-

terminal tails exposed on the nucleosomal surface. The figure shows the sequences of

human H3 and H4 N-terminal histone tails, in single letter amino acid code and the

potential modifications. Red K stands for acetylatable lysines; sites of potential

phosphorylation are represented in purple and methylation sites in blue. In addition, the

interplay between different post-translational modifications is shown. Acetylated (Ac),

Methylated (Me), Phosphorylated (P) residues are represented by those letters. Positive

and negative effects are indicated. From Zhang and Reinberg, 2001.

Epigenetic modification at histone tales gives the nucleosome surface chemical

information and “memory” (Turner, 2007 and Li et al., 2007). Such modifications have

been observed when studying the role of the chromatin in transcription (Reviewed in Li

et al., 2007). For example, acetylation of histone 3 and 4 or di- or tri-methylation of

H3K4 is known as euchromatin modifications because they are implicated with sites of

active transcription. On the other hand, heterochromatin domains are characterised by

16

methylation on H3K9 and H3K27 residues which are correlated with transcription

repression (Strahl and Allis, 2000). Although it is often suggested as a “common rule”

that acetylation opens up the chromatin for transcription and that methylation does the

contrary (repression of transcription), this is by no means a universal truth. For instance,

acetylation of histone H2A at position K119 results in repression of transcription while

methylation of H3 K4 results in activation of transcription. For more detail see table 2.

The majority of the modifications are found in regions within the upstream region, the

core promoter, the 5’end of the open reading fragment (ORF) and the 3’end of the ORF,

and this location is of great importance for transcription (Li et al., 2007).

Table 2: Histone modifications associated with transcription. Li et al., 2007.

Ever since the discovery of histone modifications, scientists have been wondering

whether these modifications have specific functions and what those functions are. This

has led to the introduction of a new term, the epigenetic code. It is known that all

histone modifications, except for methylation, result in a change in the overall charge of

the nucleosomes, leading to changes in inter- and intranucleosomal interactions (Li et

al., 2007). Histones that have been acetylated are easier to remove from the DNA in

vivo (reviewed in Li et al., 2007). Furthermore, it has been proposed that such

17

modifications could have an effect on the higher order chromatin structures. For

example, acetylation of H4K16 is responsible for inhibiting the formation of 30nm

chromatin fibbers (Shogren-Knaak et al., 2006).

One possible model of how histone modifications can interfere with transcription

initiation was suggested by Li et al., 2007.

Figure 7: Schematic representation of chromatin regulation during transcription. In the

silent promoter nucleosomes are found flanking a ~200bp NFR (nucleosome free

region). Prior to UAS targeting, co-activation recruitment occurs (e.g. recruitment of

SAGA and mediator). At this point, histones are found to be acetylated at promoter

regions mobilizing the nucleosomes in order to prepare the chromatin for transcription.

In the model on the left, called histone eviction, the acetylated histones and the

chromatin remodeling complexes collaborate to remove the HTZ-1 histone variant from

chromatin. Thus, the promoter is now free to engage the GTFs and RNA Pol II, and

with the help of SAGA and mediator, enable the formation of a pre-initiation complex

(PIC). The partial PIC model on the right shows the assemblage of the PIC without loss

of HTZ-1. In this case, the binding of Pol II and the transcription factor TFIIH are

responsible for histone eviction and therefore the assembly of PIC at the promoter (Li et

al., 2007).

18

5.NON-CODING RNAs AS EPIGENETICS REGULATORS

The third class of epigenetic modifications is mediated by non-coding RNAs.

Growing evidence supports the role of non-coding RNA and RNA-binding proteins in

chromatin development, modifications and epigenetic control.

Non-coding RNAs are generated from every part of the genome, and they have been

underestimated for years. It has not been until recently that whole genome approaches

using high-throughput transcript sequencing have shown that almost all the genome (85

to 90%) is actually transcribed, contrary to what was previously believed (Human

Molecular Genetics). This led to the discovery that non-coding RNAs can be transcribed

from introns, exons, long intergenic regions, and enhancer regions (Mattick and

Makunin 2005). Moreover, they play a fundamental role in the survival of the cell by

mediating gene silencing repressing the mobilization of transposable elements, as well

as for defence mechanisms against virus attack (Matzke et al., 2000). Several kinds of

non-coding RNAs have been identified to have a relationship with chromatin

modifications and regulation. Examples of them are microRNA (miRNA), piwi-binding

RNA (piRNA), endogenous short interfering RNA (endo-siRNA) and long non coding

regulatory RNA. Their major functions include gene regulation in development,

repression of transposon activity, gene imprinting, X-inactivation and antisense

regulation (Hum Mol Gen, 4th

Edition).

Mechanisms of silencing mediated by RNAi can be grouped into two different

categories:

PTGS: when RNAi leads to post-transcriptional gene silencing.

TGS: when RNAi leads to transcriptional gene silencing, including mechanisms that

epigenetically modify chromatin (see Tomari and Zamore 2005).

RNAi-like mechanisms are often found to co-operate in gene silencing, dosage

compensation, DNA elimination in protists, and centromere silencing (Reviewed in

Bernstein and Allis 2005).

19

Figure 8: Illustration showing some examples of non-coding RNA roles in histone

modifications in different kingdoms. From Bernstein and Allis 2005.

Several different examples of how RNAs play an important role in genetic regulation

in the four different kingdoms of life are shown. An example from the kingdom of

Fungi is the yeast S. pombe where siRNA influence methylation of H3K9 and H3K27 to

repress transcription. The fission yeast silences genes in centromeric and telomeric

regions as well as in the mat loci using RNAi dependent mechanisms (see Forsburg,

2005). Representing the Protist kingdom is Tetrahymena thermophila which in the

vegetative state (upper cell) contains a micronucleus (MIC). The micronucleus is

transcriptionally inactive (silent) and a macronucleus (MAC) which is transcriptionally

active (represented in green). When conjugation occurs, represented in the lower panel,

the MIC divides to form the new MAC and the old MAC is marked with H3K9me and

K27me for destruction by DNA an elimination process initiated by siRNA. In the Plant

kingdom siRNAs regulates heterochromatic knobs (Arabidopsis, maize) by H3K9me

and in the Animal kingdom the mechanisms of dosage compensation in both flies and

mammals involve regulatory RNAs (roX and Xist, respectively). Whereas mammals

silence one of the X chromosomes in females, flies up regulate expression of genes on

of the male X chromosome by two fold. In flies the process is control by the roX RNAs

and the chromosome is marked by H416ac and H3K4me while the centromere that is

transcriptionally inactive is marked by H3K9me. In mammals, the silencing of the X-

chromosome is achieved by the Xist RNAs and the chromosome is marked by

chromatin modifications such as H3K9me, H3K29me and H4K20me (for more details

see figure 10).

Recent publications have also identified a combined role of the spliceosome

(complex of snRNA and proteins which function is to remove introns) and chromatin

where some chromatin modifications influence spliceosomal functions. For instance,

20

HDAC (histone deacytalase) inhibition modifies the splicing patterns of around 700

human genes. Moreover, the spliceosome has been seen to interact with chromatin

regulators such HP1 or MRG15 (reviewed in Bierne et al., 2012). Likewise, it has been

revealed that in the genome of Arabidopsis thaliana there is a strong correlation

between DNA methylation, DNA repeats, non-coding RNA and H3K9 methylation in

the heterochromatin knobs (see figure 8) (reviewed in Bernstein and Allis, 2005).

Amplification of siRNA responses to dsDNA in the worm C. elegans by RNA

dependent RNA polymerase leads to a strong response that can persist for at least two

generations. Additionally, Ashe et al., 2012 have reported a transgenerational epigenetic

memory generated by piRNAs in C. elegans. piRNA or piwi-protein-interacting RNAs

are different from siRNAs in size and biogenesis. Their major role is to avoid

transposition of retrotransposons and they have been found in several animals (Hum

Mol Gen, 4th

Edition; Ashe et al., 2012). A co-operation between chromatin factors and

RNAi which is involved in passing on the signal has been identified. piRNAs can

mediate stable long-term silencing for at least 20 generations, depending solely on the

nuclear RNAi and chromatin silencing pathways. The exact mechanisms involved in

transgenerational inheritance are still unclear.

6.THE EPIGENETIC CODE

Ever since the discovery of epigenetics, scientists have sought to understand the

mechanisms involved in the chromatin changes that lead to changes in gene expression.

The notion of an “epigenetic code” was originally coined as the code that hypothetically

defines the specific modifications that occur in the histones of the nucleosomes as a

consequence of external factors and how these changes are in principle, inherited

through a series of cell divisions. Unlike the genetic code, it is apparent that the

epigenetic code is both tissue and cell specific, and differs significantly between

organisms. In 2007 an issue about the epigenetics code was published in Nature Cell

Biology “Defining an epigenetic code” where the author Brian M. Turner tried to settle

the bases of a possible epigenetics code. He established that the primarily the code

needs to be arbitrary, meaning that the code does not imply a direct relation between the

sign and the outcome (e.g. red traffic light means stop and there is not direct

relationship between the red light and the fact that we have to stop the car).

For example, as Crick already stated: “The genetic code describes the way in which a

sequence of twenty or more things is determined by a sequence of four things of a

different type” (Crick, 1963). Something similar could be used to determine the

epigenetic code, although it has turned out to be difficult to decide what the epigenetics

code would actually code for. It would be more useful and correct to talk about an

epigenetic code when possible inheritable modifications are involved in long term

effects. Turner proposes the following definition: “The epigenetic code describes the

way in which the potential for expression of genes in a particular cell type is specified

by chromatin modifications put in place at an earlier stage of differentiation” (Turner,

21

2007). Contrary to the genetic code, the epigenetic code has many possible numbers of

signs (more than one outcome is possible) and the number of outcomes is likely to be a

large one (see figure 1). Hence, for the code to be meaningful, it must be combinatorial

(Turner, 2007). In addition, many other factors such as non-coding RNAs can alter the

chromatin status which in turn can affect the epigenetic code (Bernstein and Allis,

2005). Thus, the context of the cell and the DNA surroundings matter and has to be

taken in consideration when deciphering the epigenetic code (Weiner and Toth 2011).

Bearing this complexity in mind it is not surprising that little progress has been made in

deciphering the code thus far, and to many experts the goal seems unattainable. It is at

least clear that a universal epigenetic code, similar to the genetic code does not exist.

Figure 9: Environmental factors or conditions cause histone modifications at the amino

acid tails that can define short-term and long-term output. Depending on the stimuli, the

result can be a short-term responses or heritable traits. From Turner, 2007.

7.X-CHROMOSOME INACTIVATION IN MAMMALS

The X-chromosome inactivation is a common phenomenon in mammalian

organisms. It occurs as a consequence of the need to balance gene expression levels

between females and males due to the different sex chromosomes. While females carry

two copies of the same sex chromosome (XX), males only carry one (XY) (Lyon,

1961). The process of silencing is the best example of epigenetic regulation through

DNA methylation along an entire chromosome that subsequently gets converted into

heterochromatin along most of its length (Heard et al., 1997) forming a structure called

the Barr body. As mentioned above, non-coding RNAs also play an important role in

22

this process (see figure 8). Shortly after implantation of the nascent embryo, the X

chromosome to be silenced is chosen at random in each cell. The silencing is then

passed on to the progeny of that cell. Notably, the X chromosome silencing in human is

not entirely complete as some genes escape from the inactivation and are express to

some degree (Carrel et al., 1999). X-inactivation starts with the expression of the XIST

gene (X inactive specific transcript). XIST is localized in the middle of the inactivation

centre of the chromosome X (XIC, X inactivation centre in Xq13). The XIST gene is in

a demethylated state when the X chromosome is silenced but is methylated on active X

chromosomes (see Rodríguez et al., 2004). There are several features that differentiate

the inactive X chromosome from the active one. Among them are different distribution

of histone variants, delayed replication in S-phase and different covalent histone

modifications. Furthermore, the inactivated human X chromosome contains two

different kinds of non-overlapping heterochromatin. One of these two domains in the Xi

is characterized by the presence of Xi-specific transcript RNA, histone variant

macroH2A and histone H3 trimethylated at lysine 27 (H3TrimK27) while the other

domain includes histone H3 trimethylated at lysine 9 (H3TrimK9), heterochromatin

protein 1 and histone H4 trimethylated at lysine 20 (H4TrimK20) (see Chadwick and

Willard 2004).

Figure 10: X-chromosome inactivation. Chadwick and Willard proposed this scheme

for the specialized packaging of the inactivated X chromosome into the Barr body. The

two different types of heterochromatin that can be found in the X inactivated

chromosome are distinguished by macroH2A, H3TrimK27 (represented in red) or XIST

(yellow) and H3TrimK9, H3TrimK20 (represented in green) and HP1 (blue). When the

chromosome is packaged into the Barr body at metaphase, the XIC (X inactivation

centre) synthesizes XIST RNA that associates with macroH2A and H3TrimK27 in cis

along the chromosome. On the other hand, HP1 associates with H3TrimK9 and

H3TrimK20. Figure from Chadwick and Willard 2004.

23

8.STEM CELLS

Epigenetic mechanisms are responsible for the maintenance of pluripotency in stem

cells through cell divisions. This is achieved by histone modifications, Polycomb group

proteins, non-coding RNAs and chromatin remodeling complexes leading to the

expression of specific regulatory genes (Lunyak and Rosenfeld 2008).

Stem cells are defined cells that are able to divide “endlessly” conserving the same

phenotype and give rise to differentiated somatic cells (Plews et al., 2012). Stem cells

can be classified into three groups, embryonic, postnatal, and induced pluripotent stem

cell (iPS). The embryonic cells are considered to be pluripotent, meaning that they can

differentiate into every cell of the organism (excluding placenta and adjacent tissues).

Postnatal are cells found in some organs in the body and contribute to regeneration and

renewal of that particular organ or tissue, whereas induced pluripotent stem cells are

stem cells derived from a non-pluripotent cell by experimental manipulation (see

Williams et al., 2011). Pluripotency of somatic cells decrease as they differentiate

caused by gene silencing due to specific expression of transcription factors (TFs),

chromatin modifications, DNA methylation, and histone modifications (Lunyak and

Rosenfeld 2008). The TFs most important for pluripotency are Oct-4, Nanog, Sox-2 and

some others such Stat3, Tbx3, FoxD3, Myc, p53. Their common function is to repress

transcription of genes that lead to differentiation (reviewed in Lunyak and Rosenfeld

2009; Szutorisz and Dillon 2005). Cells from the blastula are characterized by

unmethylated DNA and genes involved in pluripotency at early stages of embryonic

development are secure from chromatin remodeling complexes and therefore from

silencing, while the rest of the genome is known to be subject to compaction and

relaxation of chromatin (Williams el al., 2011).

Differentiated cells can already be re-programmed into a stem cell state and the next

step is to be able to direct these cells toward specific fates or phenotypes. The use of

adult cells instead of embryonic stem cells is also less ethically controversial.

Understanding and being able to manipulate the epigenetic mechanisms involved in the

maintenance of cellular pluripotency will lead to important medical advantages.

24

9.IMPRINTED GENES

In mammal cells, autosomal genes are present as alleles, two copies of the same

gene, one inherited paternally and the other one inherited from the mother. Both are

normally evenly expressed in the cell. Nonetheless, there is a small subset of genes that

does not follow this rule. These genes which are transcribed unequally depending on

their paternal origin are known as imprinted genes (Davies et al., 2005). This is due to

epigenetic mechanisms that control their silenced or transcribed state. Defects in

imprinting are linked to several known syndromes (Reik and Walter, 2001; Davies et

al., 2005).

The methylation pattern of the vast majority of genes is reprogrammed shortly after

fertilization of the embryo. In other words the methylation patterns present in the egg

and sperm are deleted (Kafri et al., 1992; Morgan et al., 2005). However, some genes

retain a specific methylation pattern of alleles that is conserved and passes through to

the embryo. These genes are known as imprinted genes. In the first stages of germ cells

development imprinted genes undergo total erasure of the methylation patterns.

Methylation marks are then reset during spermatogenesis or oogenesis so that both

alleles become identically imprinted (Reik and Walter, 2001). Some imprinted

methylation patterns can be maintained throughout development and tissues or they can

change through development and acquire specific methylation patterns in specific

tissues (Shemer et al., 1997). Nevertheless, how these epigenetically imprinted marks

are linked with methylation remains a mystery. It has been shown that most imprinted

genes are found in clusters (around 80% of all known) and unusually contain regions

rich in CpG (Paulsen et al., 2000). In addition, no histone or chromatin modifications

that are independent of methylation seem to be involved in the development of

imprinted genes (Reik and Walter, 2001).

25

Figure 11: Imprinted clusters on human chromosome 15 and the orthologous regions

on mouse chromosome 7. Maternally expressed genes are shown in red while paternally

expressed genes are shown in black and imprinted genes with intermediate expression

patterns are represented in green. a) Represents the Beckwith–Wiedemann (BWS)

cluster and b) Prader-Willi syndrome/Angelman (PWS/AS) syndrome cluster (15q11-

q13. Adapted from Reik and Walter 2001.

As explained above, imprinting can be understood as the expression of a determined

gene in a parent-of-origin. Methylation of imprinted genes has been studied in CpG

islands of three differentially imprinted genes, the H19, the IGF-2 and the IGF2-

receptor (for review see Obata et al., 1998). These three genes present a specific

methylation state according to maternal or paternal inheritance. Using knock-out mice

for the receptor IGF-2 gene it was demonstrated that when the mice inherited the non-

functional IGF-2r maternally, they displayed a larger body size, 20-30% more than the

control mice, and died at birth (Lau et al, 1994). However, when the paternal allele was

deleted, the embryo developed normally to adult state (Kono, 1996). Many experiments

have been performed after the discovery that DNA methylation is required to ensure

proper fetal development. Before 2010, the number of known imprinted genes was

around 50 in the mouse, most of them having a counterpart homologue in the human. It

was around 2010 when a science paper signed by Gregg et al. claimed to have found

1300 possible loci by performing a genome-wide characterization in the mouse brain.

The method used relied on whole-transcriptome sequencing (mRNA-seq) of SNPs and

robust statistical methods to determine false negatives (Gregg et al. 2010). Studies by

two separate groups suggest that results obtained by genome-wide high-throughput

26

screens must be carefully studied and alternative methods are necessary to validate data

(DeVeale et al., 2012; Okae et al., 2012).

As stated previously, the study of genomic imprinting is of great importance because

it affects many diseases, gene aberrations and syndromes. One of the best studied cases

involves the two well-known syndromes, Prader Willi and Angelman.

9.1. Prader Willi syndrome

Prader Willi is a syndrome caused by the loss of function of imprinted genes on

chromosome 15 (region 15q11-q13), that are imprinted and expressed paternally. The

loss of function can be due to deletion or silencing of the genes involved. The syndrome

affects about 1 in 10,000 to 30,000 people worldwide (Bittel and Butler, 2005).

Figure 12: Picture showing facial phenotypic characteristics from Prader Willi

syndrome. Clinical symptoms include hypotonia (severe weak muscle tone), feeding

difficulties at early infancy, short stature, undeveloped genitalia (mostly sterile) and

delayed language and learning skills, and motor development. In childhood patients

show an insatiable appetite that if not controlled leads in the majority of the cases to

morbid obesity. PWS is also characterized by a peculiar behavior phenotype;

individuals suffer temper tantrums, stubbornness, manipulative behavior, and

obsessive-compulsive characteristics (Driscoll et al., 1998).

The genetics behind this syndrome are explained by genomic imprinting and usually

caused by deletion of a segment of chromosome 15. In around 70% of cases the

syndrome is caused by the deletion of a segment of the paternal chromosome 15 in each

cell. Sometimes the cause is the inheritance of two copies of the chromosome 15 from

the mother (maternal uniparental disomy) (25% of the cases). More rarely the Prader

Willi syndrome can be due to chromosomal translocations or silencing of the region

(inactivation of the genes) (see Driscoll et al., 1998). In those cases where the syndrome

is caused by silencing of the gene, the paternal chromosome contains a maternal pattern

of imprinting. However, it has been claimed that other epigenetic phenomena than

27

silencing can be involved in the development of the syndrome (see Rodríguez et al.,

2004).

9.2. Angelman syndrome

The genetics behind this syndrome show that it is due to the loss of function of the

UBE3A gene located within the same 15q11.2-q13 chromosomal region that is involved

in Prader Willi syndrome. Both alleles paternally and maternally inherited alleles are

active in almost every tissue of the body except in the brain where only the allele

inherited from the mother is active. The fact that only one copy is actively expressed in

the brain means that Angelman syndrome will develop if the active copy is defective.

The loss of function is predominantly caused by deletion of the gene (in 70% of the

cases). In about 11% of cases, the syndrome is caused by a mutation in the maternal

UBE3A copy. A small minority of cases are caused by paternal uniparental disomy or

chromosomal translocations but curiously the causes of 10-15% of cases still remain

unknown and other genes or chromosomes are likely to be involved (Genetics home

reference, 2011).

Angelman syndrome is present every 1 in 12,000 to 20,000 people, a lower incidence

than Prader Willi syndrome, although both syndromes are caused by defects in the same

chromosomal region (see figure 11). Like Prader Willi, Angelman syndrome is a clear

example of genomic imprinting (Dagli and Williams, 1998).

Figure 13: Picture showing the typical facial characteristics of Angelman syndrome.

Individuals possess coarse facial features (1 and 2). Pictures A to F clearly demonstrate

the distinct happy, excitable behavior, characteristic of Angelman syndrome, with

frequent smiling and laughing including uplifting of arms. Some other characteristics

include ataxia (hard and unbalanced movements), epilepsy, delay in development,

severe retardation, hyperactivity and water fascination, Source: psychnet-uk (1 and 2)

and Dagli and Williams, 1998 (A to F).

1 2

28

10. OTHER SYNDROMES CAUSED

BY EPIGENETIC MISFUNCTION

10.1. Rett syndrome

Rett syndrome is a neurological defect caused by mutations in Methyl-CpG-binding

protein 2, (MeCP2) which is encoded on the X chromosome (Xq28). It affects mainly

women, with a prevalence of 1 in 10 000. Due to the random nature of X inactivation

50% of cells in the body of heterozygous females heterozygous express only the

mutated allele. It has recently been reported that MeCP2 binds strongly to 5-

Hydoxymethylcytosine (5-hmC), an epigenetic mark predominantly found in the brain,

as well as to conventional methylcytosine (5-mC). MeCP2 binding to 5-hmC results in

transcription of highly expressed genes. The Rett syndrome mutation MeCP2, R133C

has been demonstrated to specifically inhibit binding of MeCP2 to 5-hmC while binding

to 5-mC still occurs. This likely explains why symptoms are neurological (Mellén et al.,

2012).

10.2. Fragile-X syndrome

Fragile X syndrome is associated with the FRAXA region on the X-chromosome

(fragile site, X chromosome, A site - Xq27.3). It is due to a massive trinucleotide repeat

amplification within the gene FMR1 (fragile X mental retardation-). The 5’ region of

the first exon contains a CGG trinucleotide repeat which is between 7-60 fold in normal

individuals. Affected individuals possess repetitions that are larger than 230 repeats and

usually hyper-methylated. Therefore, the gene experiences silencing and the product is

missing.

Alleles containing between 60 and 230 CGG repetitions are considered permutated,

i.e. the gene is normally transcribed due to normal methylation profiles but highly

unstable during female meiosis and can easily give cause amplification and subsequent

methylation in the next generation (reviewed in Jin and Warren 2000).

29

11. MODEL SYSTEMS TO STUDY EPIGENETICIS

Currently, there are many examples and model species available in laboratories to

study epigenetics. Some of them present better advantages over others depending on the

features of interest. From unicellular fungi, plants, mammals, etc, it has been shown

possible to study epigenetics marks one way or another.

For instance, yeast has been widely used to study chromatin structure and function.

Yet they do not present methylation in their genome, the fission yeast

Schizosaccharomyces pombe presents siRNA regulation on histone modifications. In

turn, Saccharomyces cerevisiae also known as the budding yeast has been best studied

for the annotation of the chromosome structure, cell cycle, telomere silencing and

mating type switching (mat loci) (Forsburg 2005; Nakayama et al., 2001). Neurospora

crassa is the fungus chosen to study methylation patterns in the DNA given the fact that

yeast lack DNA methylation profiles (Allis et al., 2007).

Caenorhabditis elegans is another key model organism well suited for the study of

epigenetics. The worm has a fairly short generation time (around 3 days), is easy to

manipulate and maintain (feeds on bacteria, e.g. E.coli) and easy to control and follow

under experimental conditions. It has been used to study epigenetic-mediated cellular

differentiation, dosage compensation of X-chromosomes, and RNAi, epigenetic

inheritance, etc (see Ashe et al., 2012).

The protozoan Tetrahymena thermophilia is another example of an organism use for

studies of epigenetics. T. thermophilia contains a micro nucleus, MIC that is inactive

during vegetative life and is only active during reproduction (for more details see figure

8) (Bernstein and Allis 2005; Allis et al., 2007).

The fruit fly Drosophila melanogaster has been the classical genetic model for

epigenetic investigation. The epigenetic mechanisms studied in Drosophila include the

PEV or position variegation effect, the Polycomb group (PcG) and the Trithorax group

(TrxG) that will be explained later. One of the disadvantages of Drosophila is that

cytosine methylation is virtually absent and hence, DNA methylation cannot be studied

in this model organism (Hum Mol Gen, 4th Edition; Kraus and Reuter, 2011). More

recently the fly has been used to study transgenerational epigenetic inheritance

(reviewed in Chong and Whitelaw, 2004)

In the plant kingdom Arabidopsis thaliana is the most intensely studied plant by far.

Paramutations that were firstly discovered in maize (see below) are currently studied in

Arabidopsis (Chong and Whitelaw, 2004). It also serves as a sophisticated tool for the

study of RNAi pathways, DNA methylation, environmental studies, histone

modifications and many more (Henderson and Jacobsen 2007).

30

12. PLANT EPIGENETICS

One of the best examples of epigenetics inheritance is found in plants. Some plants

exhibit a phenomenon called paramutation that does not follow Mendelian rules of

inheritance. Paramutation is defined as the interaction between two alleles of the same

locus where one of them changes the gene expression of the other, making this change

heritable without involving mutation (Chong and Whitelaw, 2004).

An example of paramutation involves a silenced allele that acts in trans on a

homologous allele generating heritable and stable inheritance, then this new silenced

allele can also induce a silenced state in other alleles as paramutations (Stam and

Scheid, 2005).

Paramutation was first identified in the maize (Zea mays) where it plays an important

role determining pigmentation. Paramutation has also been discovered in polyploid

plants where this feature is of great importance for plant specialization (Scheid and

Paszkowski 2003).

12.1. Arabidopsis thaliana

Figure 14: Picture of Arabidopsis thaliana. A. thaliana is one of the best model plants

for studying epigenetics. Available and small sequenced genome allows forward and

reverse genetics.

Arabidopsis thaliana was the first polyploid plant (tetraploid) where paramutation-

like effects have been reported (Scheid and Paszkowski 2003). Scheid and Paszkowski

identified that diploid A. thaliana homozygous for a hygromycin phosphotransferase

transgene (HPT) maintain uniform hygromyzin resistance over generations while

31

crossing with tetraploid plants resulted in plants with reduced HPT activity. This change

was discovered to be due to epigenetic silencing activated by changes in ploidy.

Another well studied epigenetic regulatory system in A. thaliana is the Polycomb

system FLC. FLC stands for Flowering Locus C and it is responsible for flowering. It is

an example of how the environment triggers the epigenetic switch for flowering which

depends on how the cold or warm the winter was (Kim and Sung, 2013). During early

development of the plant, FLC is maintained repressed preventing flowering. In

addition, its expression can also be found silenced if the adult plant has been exposed to

long periods of cold weather, leading to a late flowering. The FLC is found to be

regulated solely by the Polycomb system that mediates histone methylation and not by

DNA methylation (Sung and Amasino, 2005).

DNA methylation is also widespread in plant genomes both on Cytosines preceding

Guanines (CpG) and in Cytosines not directly preceding Guanines (CpNpG) identified

in both symmetric and asymmetric sequence context (Vaughn et al., 2007). It is known

that the cytosine patterns in eukaryote genomes are inherited from mother cell to

daughter cell in a process involving methyltransferases enzymes. Methyltransferases are

known to act symmetrically on CG and CNG nucleotides. Nevertheless, the process is

not entirely perfect and possible misses leadings to loss of methylation are corrected by

de novo methyltransferases (explained previously), although the mechanisms of action

are still unknown (Bender, 2004). In mammalian cells maintenance and de novo

methyltranferase mutants are lethal owing to the broad number of genes that are miss

regulated, being this issue a strong evidence of how important imprinted genes are for

mammalian genomes (Klose and Bird 2006). However, while mammals do not allow

methylation mutation studies, plants like Arabidopsis can in fact tolerate them, creating

a great advantage for methylation genetic analysis.

13. INSECT EPIGENETICS

Phenotypic plasticity is a mechanism where epigenetic modifications allow

organisms to adapt their phenotypes to a changing environment without genetic change.

A good example of genome plasticity is found in the group of Hymenoptera which

includes social insects such as ants, bees and wasp, all of them showing behavioural

plasticity. Social insects are frequently found organized in complex levels of

organization (castes) presenting different phenotypes that are derived from the same

genetic input (e.g. worker bees are sterile but not the genetically identical queen bee).

This is mainly due to environmental reasons that affect the genomic plasticity either by

inferring in alternative splicing, posttranslational modifications, transcription regulation

or epigenetic modifications (see Weiner and Toth, 2011). DNA methylation has been

proposed to be the main epigenetic modification to cause the development of castes,

although some other epigenetic modification might have to be taken into consideration

(Drewell et al., 2012). DNA methylation is important in this organism to suppress

mobile elements, transposons being the most methylated regions found within gene

32

bodies. It has been suggested that DNA methylation in social insects affects caste

determination and knock-down experiments of DNA methyltransferases have lent

support to this hypothesis (see Weiner and Toth, 2011). For example, dmnt3 knockdown

in worker bees at larval stage will produce adult honeybees presenting queen-like

features (larger size, larger ovaries) and methylation patterns characteristics of the

queen (see Weiner and Toth, 2011).

Nevertheless, not all insect show DNA methylation patters as strong as social insects.

For instance, one of the best studied model organisms in epigenetics is the fruit fly

Drosophila melanogaster which presents very little amount of DNA methylation

throughout its genome. Thus, it has lost almost all of its DNA methyltransferases

(Krauss and Reuter, 2011). However, Drosophila melanogaster is highly useful to study

other epigenetic features such as chromatin modifications, plasticity of the nervous

system, RNA silencing, and for the study of the Polycomb and Trithorax systems.

13.1. The Polycomb system

The Polycomb system is a group of proteins responsible for chromatin remodeling.

One of the best studied examples involved Polycomb regulation are the Hox genes. The

Hox genes modulate the chromatin structure through development in multicellular

organisms where methylation of cytosines regulates gene expression in in cooperation

with the Polycomb complex to achieve development of an organism (Terranova et al.,

2006). Even though a lot of work has been done in Drosophila on the study of the

polycomb complex and Trithorax system, it is still unclear exactly how these systems

are regulated.

In humans, the contribution of the Polycomb complex to development is poorly

understood. Lee and colleagues found that the Polycomb Repressive Complex 2

(PRC2) subunit SUZ12 in human stem cells is spread over 200 genes containing

trimethylated nucleosomes at lysine 27 (transcription repressor). Among the genes

found were stem cell transcription factors like Oct4, Sox2 and Nanog (see Lunyak and

Rosenfeld 2009; Szutorisz and Dillon 2005). PRC2 targeted genes are mostly activated

upon stem cell differentiation. The authors conclude that the repression by H3K27

methylation is important to maintain pluripotency of ES cells (Lee et al., 2006).

Moreover, Polycomb group proteins have been reported to be associated with aberrant

methylation patterns in cancer because they interact directly with DNMTs directing

them to chromatin with Polycomb target genes being more prone to be abnormally

methylated than others (see Paschos and Allday 2010). Mitotic inheritance of gene-

expression specific profiles is also catalyzed by the Polycomb group and Trithorax-

group complexes by methylation of histone H3K26 and H3K4. Suggesting a close

cooperation between the Polycomb group and methylated histones in the chromatin that

could guide them back to destination sites after cell division, although the mechanisms

involved are still unknown (Bernstein et al., 2007).

33

14. YEAST AND EPIGENETICS

Saccharomyces cerevisiae and Schizosaccharomyces pombe are the two eukaryotic

unicellular organisms belonging to the fungi kingdom that are most used in laboratories

to answer biological and molecular questions. Although neither of the two yeast

methylate DNA, they serve as great tools for the study of epigenetics. Examples of

epigenetics processes studied in yeast include the RNA interference (RNAi) machinery

that is important in the fission yeast Sc. pombe but conspicuously absent in S cerevisiae.

This system makes the fission yeast especially useful because it has single-copy genes

for components of the RNAi pathway (argonaute, dicer and RNA-dependent RNA

polymerase (RdRP)) (Martienssen et al., 2005).

Another epigenetic mechanism studied is the position effect variegation (PEV). It is a

rare event observed in genes that are located close to boundaries between euchromatin

and heterochromatin. These genes can represent a mosaic pattern of gene expression as

a result of silencing or activation. One example is found in Sc. pombe, where the gene

CLR4 (a SET domain containing histone methyltransferase enzyme 6) homologue of

the Drosophila gene Su(var)3-9 (suppressor of variegation effect) is responsible for

silencing of the mating type by methylation of the lysine residues at position 9 in

histone 3 (H3K9) (Nakayama et al., 2001).

Yeast lack DNA methylation, instead they maintain the heterochromatic state over

cell divisions through hypoacetylated histones and SIR proteins (in the case of

S.cerevisiae) and between the interplay of H3K9 methylated histones and the Swi6

chromodomain (in the case of S. pombe), both to conserve the heterochromatic state

during cell division (Grewal and Moazed, 2003; Nakayama et al., 2001).

Furthermore, in S. pombe, RNAi are found to act in transcriptional gene silencing

silencing of repetitive regions especially through H3K9 methylation (Bernstein and

Allis 2005). RNAi-like mechanisms can be seen acting on repetitive sequence as the

centromeres, the mating type loci and at telomeres repetitions (see figure 8) (Bernstein

and Allis 2005).

However, S. cerevisiae has been widely chosen over S. pombe for the study of

chromatin remodeling complexes, telomere silencing, etc. For instance, in S. cerevisiae

the telomere position effect has been elucidated. The telomere position effect is an

epigenetic mechanism which silences all the genes close to telomeres and is dependent

on SIR proteins. Besides telomere silencing the mating-type loci (HML and HMR) and

tandem rDNA arrays, are silenced by conversion to heterochromatin (Loney et al.,

2009). Furthermore, S. cerevisiae has been used to study histone modifications like

acetylation, phosphorylation, etc.

Other phenomena originally identified in budding yeast include the role of

chromatin complexes like the INO80 which mobilizes nucleosomes throughout the

DNA and where it has also been found that the Rvb proteins are necessary for the

functional integrity of the complex by recruiting the Arp5p (Jónsson et al., 2004).

34

Studies in the field of epigenetics using yeast as model organisms have yielded

countless discoveries, making them very suitable tools for assisting basic molecular

genetics and cellular approaches.

15. CLINICAL APPLICATIONS OF

EPIGENETICS

15.1. Mitochondria in epigenetics

Being derived from a prokaryote, the mitochondrial DNA is not associated with

histones and lacks introns (except for a few self-splicing ones). However DNA

methyltransferases have been identified in mammalian mitochondria. These include

mitochondrial DNA methyltransferase 1 (mtDNMT1) which gives rise to both 5-mC

and 5-hmC (Shock et al., 2011). As already explained above, 5-mC stands for

methylation at carbon 5 of cytosine residues and 5-hmC stands for hydroxymethylation

(Kriaucionis and Heintz, 2009). Defective mitochondria will cause dysfunction in redox

state and in S-adenosyl methionine (SAM-CH3) production. SAM-CH3 is precursor that

donates a methyl group to methylate cytosines (see figure 3). Such dysfunction can

cause changes in methylation profiles in the nuclear genome leading to genomic

instability and pathologies (Minocherhomji et al., 2012).

15.2. Bacterial and virus infections

Recent publications have provided evidence that bacteria and viruses can promote

infection through epigenetics means. They can alter chromatin modifications, DNA

methylation and affect the activity of chromatin remodelling complexes in the host cells

for their own benefit.

One of the best examples of how viruses can influence host responses to infection is

found in viruses called latent (e.g. herpesviruses) that remain “silent” within genome of

the host. Other examples include the Hepatitis B virus, which has been linked to

abnormal DNA methylation of the host genome by inducing high levels of

methyltransferase (DNMT1, DNMT3A and DNMT3B) expression (reviewed in

Paschos and Allday, 2010).

The well-known HIV virus that infects CD4+ T cells has also been reported to

mediate epigenetic response by increasing DNA methylation in the infected

lymphocyte. As a result, general repression of transcription is localized at the gamma

interferon (IFN-γ) promoter leading to reduced IFN-γ expression (Mikovits et al, 1998).

35

Figure 15: Schematic description of the increase methylation in T lymphocytes

infected by the HIV virus. CD4+ T cells from the same donor were used in this study

where one set was infected with HIV-1 and the other was uninfected. Note the peak of

maximum methyltranferase activity at day 7 for the HIV-1 infected cells compared to

the control sample (healthy cells). Figure from Mikovits et al, 1998.

Other pathogens have also been linked different kinds of epigenetics responses such

as chromatin remodeling complexes, the Polycomb protein group (e.g. Paramecium

bursaria), etc. Bacteria can also disrupt the epigenetic environment of the infected host

cell. Bacteria mostly produce signalling cascades (phosphorylation/acetylation and

dephosphorilation/deacetylation) in the host cell though their components (metabolites),

bringing by activation of transcriptional factors, controlling chromatin remodeling

complexes, DNA methylation, etc. (Reviewed in Bierne et al., 2012). Furthermore, an

increasing number of bacteria having SET domains that have linked functions with

chromatin remodeling like in Chlamydophila pneumoniae (Paschos and Allday, 2010)

have been described.

Some of the main effects caused by bacterial infections are illustrated in table 3 from

Bierne et al., 2012.

36

Table 3: Bacterial species, bacterial factors and effects caused. Bierne et al., 2012.

As many studies have linked virus infection with up-regulation or down-regulation of

methylation patterns of the host cell they infect, it is of great importance to continue

research on this topic in order to understand the role of epigenetics in pathogenesis. For

instance, the use of DNMT inhibitors could be a major advantage in the treatment of

many of these diseases. The same way, the full knowledge of how bacteria influence

and disrupt the epigenome of a cell can be useful to find better treatments for infectious

diseases.

37

16. CANCER AND EPIGENETICS

It is a well-known fact that mutations are one of the bases of cancer development.

However, recently the fact that underlying epigenetic modifications also play an

important role in the maintenance and development of cancer has become apparent.

DNA methylation is essential for mammalian cells, and is necessary for many

important processes such as telomere length regulation, establishment of gene

expression, development, embryogenesis, X chromosome inactivation, cell

differentiation, etc. (Reviewed in Bernstein et al., 2007). As explained in the previous

sections, 5-methylcytosine is prone to deamination to thymine because of its chemical

instability, leading to point mutations that can result in genetic disorders such as cancer.

It is well-known that DNA methylation profiles change drastically in cancer. Changes

observed in cancerous cells include hypomethylation (loss of methylation in sequences

that are usually methylated) and hypermethylation of sequences usually unmethylated,

particularly CpG islands (Rodriguez et al., 2003; Wilson et al., 2007; Gal-Yam et al.,

2007) whereby methylation causes permanent silencing. Furthermore, high activity of

methyltransferases has been observed in several cancer types suggesting a role in the

proliferation of malignant cells. In particular, de novo methyltranferase activity which

produces permanent methylation is propagated during mitosis to descendent cells (Bird,

2002).DNA methylation changes in cancer seem to be even more prevalent than

mutations, and have been identified even in precancerous cells (Rodríguez et al., 2004).

16.1. Epigenetic therapy

Nowadays, new drugs targeting epigenetic modifications are being developed and

some of them are even approved for clinical trials. Some examples are DNA

methyltransferases inhibitors like 5-aza cytidine and 5-aza 2’ deoxyazacytine. 5-aza

cytidine works by “kidnapping” methyltransferases at nucleoside incorporation sites

(Gal-Yam et al., 2007). These two drugs, 5-aza cytidine and 5-aza 2’ deoxyzacitine

have been recently approved by the U. S. Food and Drug Administration and are

commonly used in patients with myelodysplactic syndromes (MDS).

In other types of cancer, aberrant epigenetic histone modifications are implicated,

and therefore, inhibitors of histone deacetylases are used instead, such as vorinostat in

cutaneous T-cell lymphoma. Remarkably, vorinostat induces cell cycle-arrest, and/or

apoptosis of transformed cells in vitro and in vivo (see Pratap et al., 2010). Pratap and

colleagues showed that vorinostat treatment in mice with breast and prostate cancer

cells reduces tumour growth effectively.

38

17. EPIGENETIC RESEARCH METHODOLOGY

Comparison of the epigenetically modified regions to their unmodified counterparts

can be done by common molecular biology techniques such as chromatin

immunoprecipitation, protein-protein interaction assays, protein-chromatin interaction

assays, etc. or by Next-Generation whole genome sequencing approaches.

Modifications include DNA methylation, histone modifications, RNA, nucleosome, etc.

The following sections name a few examples of approaches that are used to study

genome-wide DNA methylation patterns:

17.1. High resolution tiling oligonucleotides arrays:

Tiling oligonucleotides are a set of overlapping oligonucleotide probes that represent

a contiguous specific known region in the DNA. They are used to characterize genomic

regions bound by specific factors or modifications. Another methodology is the WGA

(whole genome array) where array-based hybridization is used to scan the complete

genome of an organism (Gal-Yam et al., 2007).

17.2. Shotgun bisulfite sequencing:

This method is used to profile the methylation of cytosines in genomic DNA. The

assay consists on treating the DNA with sodium bisulfite or metabisulfite (Na2SO3 or

Na2S2O5) which will deaminate cytosine to uracil leaving the 5-methylcytosine

unchanged. For identifying the bases that have changed, PCR is performed and

compared with the sequence before and after treatment with Na2SO3. Because uracil is

read as thymine during the PCR process it is possible to identify which cytosines where

methylated. There are several methods used to reveal the CpG dinucleotide methylation

status in the genome, and one of them is by sequencing after the PCR, another method

is to digest the DNA with restriction enzymes or by arrays containing specific

oligonucleotides that only match the unchanged sequence or the treated one can also be

used (see Human Molecular Genetics, Garland Science 4th Edition).

17.3. Digestion with restriction endonucleases:

Some restriction enzymes are methylation sensitive and cut the DNA at specific

sequences depending on the methylation state, for example the enzyme TaqI recognizes

39

the cut-site sequence only when cytosine is unmethylated. Another example is the

restriction enzyme McrBC that has been used to discriminate symmetric CpG and

CpNpG methylation versus asymmetric methylation in Arabidopsis in combination with

tiling microarrays that can also profile polymorphisms (Vaughn et al., 2007). Other

restriction enzymes used to identify methylation include HpaII which cuts unmethylated

CCGG and MspI that indistinctly cuts CCGG methylated or not allowing the

comparison of different size restriction fragments (Hum Mol Gen, 4th Edition).

17.4. Chromatin immunoprecipitation followed by CHIP-sequencing:

This approach can be used to detect histone modifications by using antibodies

specific for the various kinds of modified histones. First, the specific modified histone,

cross-linked to its cognate DNA is identified by chromatin immunoprecipitation (ChIP).

After this treatment, the samples are sonicated to reach ideal DNA sizes. Finally,

oligonucleotides adapters are added to the DNA in order to allow whole genome DNA

sequencing. This method also helps quantifying the amount of specific histone

modifications at given sites (Illumina, Inc. 2013).

17.5. MeDIP and next generation sequencing:

MeDIP stands for methylated DNA immunoprecipitation. This assay uses a specific

antibody raised against 5-methylcitosine (5mC). The purified immunoprecipitated

methylated DNA can then be identified and quantified by next generation sequencing

or, PCR / microarray hybridization approaches (Weber et al., 2005).

17.6. Cobra assays (Combined Bisulfite Restricted Assay):

The cobra assay is a quantitative molecular technique that employs restriction

enzyme digestions of sodium bisulfite-treated DNA (as described above) to discover the

DNA methylation levels at specific gene loci in small quantities of genomic DNA (see

Xiong and Laird, 1997).

Many more methodologies can be used coupled to the previous approaches in order

to study epigenetics utilizing PCR, DNA sequencing, quantitative PCR, Southern and

Northern blotting, HPLC, primer extension, MALDI-TOF MS, etc. (Esteller, 2008;

Suzuki, 2008).

40

Figure 16: Schematic representation of some approaches used for the detection of

epigenetic marks. DNA methylation can be identified by Bisulfite conversion, using

specific restriction enzymes to cut methylated or non-methylated DNA or by

immunoprecipitation using antibodies to 5’mC. In addition, quantification of genes

activated after treatment with 5-aza-CdR (a demethylating agent) can be done by qPCR.

Histone modifications are unraveled by ChIP (chromatin immunoprecipitation) using

modification-specific antibodies Picture taken from Gal-Yam et al., 2007.

18. CONCLUDING REMARKS

Epigenetics is a relatively new and complex branch of biology and scientist have just

started scratching the tip of the iceberg with respect to understanding it. It is still not

clear how most epigenetic marks are inherited over generations or what are all the

factors controlling the epigenetic modifications are. Furthermore, a whole new global

network has been opened with the discovery of the non-coding RNAs in the regulation

of epigenetics that in combination with DNA methylation and histone modifications

make possible the development, differentiation and adaptation of the cell to its

environment.

Better understanding of the molecular basis of epigenetics in cell differentiation,

gene expression and regulation will be beneficial for the identification, diagnosis and

treatment of syndromes, illness, cancer, infections as well as explaining syndromes that

are not explained by classical Mendelian inheritance.

Although the idea of an epigenetics code is enticing, the scientific community is

currently far away from unravelling the basis of the code due to the broad number of

combinations of the modifications that are involved.

Nevertheless, the increasing improvements of the new molecular methodologies are

currently allowing great discoveries in this new field and will help to unravel the

mysteries of epigenetics, hopefully within the near future.

41

42

19. REFERENCES

Abraham, A.L., Nagarajan, M., Veyrieras J.B., Bottin H.H., Steinmetz, L.M., Yvert,

G.G. Genetic modifiers of chromatin acetylation antagonize the reprogramming of

epi-polymorphisms. PLoS Genet. 8, 1002958 (2012)

Allis, C.D., Jenuwein, T., Reinberg, D. “Epigenetics”, M. Caparros, Cold Spring

Harbor Laboratory Press. 1, 23-62 (2007)

Ashe, A., Sapetschnig, A., Weick, E.M., Mitchell, J., Bagijn, M.P., Cording, A.C.,

Doebley, A.L., Goldstein, L.D., Lehrbach, N.J., Le, Pen, J., Pintacuda, G.,

Sakaguchi, A., Sarkies, P., Ahmed, S., and Miska, E.A. piRNAs can trigger a

multigenerational epigenetic memory in the germline of C. elegans. Cell. 150, 88-99

(2012)

Bender, J. DNA methylation and epigenetics. Annu Rev Plant Biol. 55, 41–68 (2004)

Bernstein, A., & Allis, D. RNA meets chromatin. Genes & Dev. 19, 65-1655 (2005)

Bernstein, B.E., Meissner, A., and Lander, E.S. The Mammalian Epigenome. Cell. 128,

69-681 (2007)

Bestor, T.H. The DNA methyltransferases of mammals. Hum Mol Genet. 9, 2395-402

(2000)

Bierne, H., Hamon, M., and Cossart, P. Epigenetics and Bacterial Infections. Cold

Spring Harb Perspect Med. 2, 0102722012 (2012)

Bird, A.P., & Wolffe, A.P. Methylation-induced repression – belts, braces, and

chromatin. Cel.l 99, 451-454 (1999)

Bird, A.P., Taggart, M.H., Nicholls, R.D., Higgs, D.R. None-methylated CpG-rich

islands at the human a globin locus: implications for evolution of the a globin

pseudogene. EMBO J. 6, 999-1004 (1987)

Bird, A. DNA methylation patterns and epigenetic memory. Genes Dev. 16, 6–21.

(2002)

Bird, A. Perceptions of epigenetics. Nature. 447, 396–398. (2007)

43

Bittel, D.C., Butler, M.G. Prader-Willi syndrome: clinical genetics, cytogenetics and

molecular biology. Expert Rev Mol Med. 14, 1-20. (2005)

Bönisch, C., & Hake, S.B. Hake. Histone H2A variants in nucleosomes and chromatin:

more or less stable? Nucleic Acids Res. 40, 10719–10741. (2012)

Briggs, R., King, T.J. Transplantation of living nuclei from blastula cells into

enucleated frogs’eggs. Proc. Natl. Acad. Sci. 38, 455-463. (1952)

Carrel, L., Cottle, A.A., Goglin, K.C. & Willard, H.E. A first-generation X-inactivation

profile of the human X chromosome. Proc. Natl. Acad. Sci. 96, 14440-14444. (1999)

Chadwick, B.P. and Willard, H.F. Multiple spatially distinct types of facultative

heterochromatin on the human inactive X chromosome. PNAS. 101, 17450-17455.

(2004)

Chandler, V.L., & Stam, M. Chromatin conversations: mechanisms and implications of

paramutation. Nat Genet. 5, 532-544. (2004)

Chong, S., & Whitelaw, E. Epigenetic germline inheritance. Current Opinion in

Genetics & Development. 14, 692-696. (2004)

Crick, F.H. On the genetic code. Science. 139, 461-464. (1963)

Davies, W., Isles, A.R., Wilkinson, L.S. Imprinted gene expression in the brain.

Neurosci Biobehav Rev. 29, 421-430. (2005)

DeVeale, B., van der Kooy, D., Babak, T. Critical evaluation of imprinted gene

expression by RNA-Seq: a new perspective. PLoS Genet. 8, 1002600. (2012)

Doerfler W. DNA methylation and gene activity. Annu Rev Biochem. 52, 93-124.

(1983)

Dragich, J., Houwink-Manville, I., Schaenen, C. Rett syndrome: a surprising result of a

mutation in MECP2. Hum Mol Genet. 9, 2365-75. (2000)

Drewell, A.R., Lo, N., Oxley, R.P., Oldroyd P.B. Kin conflict in insect societies: a new

epigenetic perspective. Trends in Ecology & Evolution. 27, 367-373. (2012)

Driscoll, J.D., Miller, J.L., Schwartz, S., and Cassidy, S.B. Prader-Willi Syndrome

Synonyms: PWS, Prader-Labhart-Willi Syndrome. Bookshelf ID: BK1330PMID:

20301505. Initial Posting: October 6, 1998; Last Update: October 11, (2012)

44

Esteller, M. Epigenetics in Cancer. N Engl J Med. 358, 1148-1159. (2008)

Fazzari, M.J., & Greally, J.M. Epigenomics: beyond CpG islands. Nat. Rev. Genet. 5,

446-455. (2004)

Fischle, W., Wang, Y., Allis, C.D. Histone and chromatin cross-talk. Curr Opin Cell

Biol. 15, 172-83. (2003)

Forsburg, S.L. The Yeasts Saccharomyces cerevisiae and Schizosaccharomyces pombe:

Models for Cell Biology Research. Gravitational and Space Biology. 18. (2005)

Fraga, M.F., Ballestar, E., Paz, M.F., Ropero, S., Setien, F., Ballestar, M.L., Heine-

Suñer, D., Cigudosa, J.C., Urioste, M., Benitez, J., Boix-Chornet, M., Sanchez-

Aguilera, A., Ling, C., Carlsson, E., Poulsen, P., Vaag, A., Stephan, Z., Spector,

T.D., Wu, Y.Z., Plass, C., Esteller, M. Epigenetic differences arise during the

lifetime of monozygotic twins. Proc Natl Acad Sci. 26, 10604-9. (2005)

Gal-Yam, E.N., Saito, Y., Egger, G., and Jones, P.A. Cancer Epigenetics:

Modifications, Screening, and Therapy. Annu. Rev. Med. 59, 267-80. (2007)

Gardiner-Garden, M., Frommer, M. CpG islands in vertebrate genomes. J Mol Biol. Jul

20, 261-82. (1987)

Goldberg, A.D., Allis, D.C., and Bernstein, E. Epigenetics: A Landscape Takes Shape.

Cell. 128, 635-638. (2007)

Goll, M.G., Bestor, T.H. Eukaryotic cytosine methyltransferases. Annu Rev Biochem.

74, 481-514. (2005)

Gonzalo, S. Epigenetic alterations in aging. J Appl Physiol. 109, 586–597. (2010)

Gregg, C., Zhang, J., Weissbourd, B., Luo, S., Schroth, G.P., Haig, D., Dulac, C. High-

resolution analysis of parent-of-origin allelic expression in the mouse brain. Science.

329, 643–648. (2010)

Grewal, S.I., Moazed, D. Heterochromatin and epigenetic control of gene expression.

Science. 301, 798–802. (2003)

Hall, I.M., Shankaranarayana, G.D., Noma, K., Ayoub, N., Cohen, A., and Grewal, S.I.

Establishment and maintenance of a heterochromatin domain. Science. 297, 2232-

2237. (2002)

45

Heard, E., Clerc, P., & Avner, P. X-chromosome inactivation in mammals. Annu Rev

Genet. 31, 571-610. (1997)

Henderson, I.R., & Jacobsen, S.E. Epigenetic inheritance in plants. Nature. 447, 418-

424. (2007)

International Human Genome Sequencing Consortium, et al. "Initial sequencing and

analysis of the human genome". Nature. 409, 860–921. (2001)

Jablonka, E., et al. The adaptive advantage of phenotypic memory in changing

environments. Phil. Trans. R. Soc. B. 350, 133–141. (1995)

Jablonka, E., & Lamb, M.J. The changing concept of epigenetics. Ann. N. Y. Acad. Sci.

981, 82-96. (2002).

Jjingo, D., Conley, A.B., Yi, S.V., Lunyak, V.V., Jordan, I.K. On the presence and role

of human gene-body DNA methylation. Oncotarget. 3, 462-74. (2012)

Jin, P., & Warren, S.T. Understanding the molecular basis of fragile X syndrome. Hum

Mol Genet 9, 901-8. (2000)

Jónsson, Z.O., Jha, S., Wohlschlegel, J.A., and Dutta, A. Rvb1p/Rvb2p recruit Arp5p

and assemble a functional Ino80 chromatin remodeling complex. Mol Cell. 16, 465-

77. (2004)

Kamakaka, R.T., & Biggins, S. Histone variants: deviants? Genes Dev. 1, 295-310.

(2005)

Kafri, T., Ariel, M., Brandeis, M., Shemer, R., Urven, L., McCarrey, J., Cedar, H.,

Razin, A., Developemental pattern of gene-specific DNA methylation in the mouse

embryo and germ line. Genes Dev. 6, 705-714. (1992)

Krauss, V., & Reuter, G. DNA methylation in Drosophila--a critical evaluation. Prog

Mol Biol Transl Sci. 101,177-91. (2011)

Kim, D.H. & Sung, S. Coordination of the Vernalization Response through a VIN3 and

FLC Gene Family Regulatory Network in Arabidopsis. PLANT CELL. 25, 454-469.

(2013)

Klose, R.J., & Bird, A.P. Genomic DNA methylation: The mark and its mediators.

Trends Biochem Sci. 31, 89–97. (2006)

46

Kono, T., Obata, Y., Yoshimizu, T., Nakahara, T. and Carroll, J. Epigenetic

modifications during oocyte growth correlates with extended parthenogenetic

development in the mouse. Nature Genet. 13, 91-94. (1996)

Kornberg, R.D., & Lorch, Y. Twenty-Five Years of the Nucleosome, Fundamental

Particle of the Eukaryote Chromosome. Cell. 98, 285–294 . (1999)

Kriaucionis, S., Heintz, N. The nuclear DNA base 5-hydroxylmethylcytosine is present

in Purkinje neurons and the brain. Science. 324, 929-30. (2009)

Lau, M.M., Stewart, C.E., Liu, Z., Bhatt, H., Rotwein, P., and Stewart, C.L. Loss of the

imprinted IGF2/cation independent mannose 6-phosphate receptor results in fetal

overgrowth and perinatal lethality. Genes Dev. 8, 2953-2963. (1994)

Lee, TI., Jenner, R.G., Boyer, L.A., Guenther, M.G., Levine, S.S., Kumar, R.M.,

Chevalier, B., Johnstone, S.E., Cole, M.F., Isono, K., Koseki, H., Fuchikami, T.,

Abe, K., Murray, H.L., Zucker, J.P., Yuan, B., Bell, G.W., Herbolsheimer, E.,

Hannett, N.M., Sun, K., Odom, D.T., Otte, A.P., Volkert, T.L., Bartel, D.P., Melton,

D.A., Gifford, D,K., Jaenisch, R., Young, R.A. Control of developmental regulators

by Polycomb in human embryonic stem cells. Cell. 125, 301-13. (2006)

Li, B., Carey, M., and Workman, J.L. The role of Chromatin during transcription. Cell.

128, 707-719. (2007)

Loney, E.R., Inglis, P.W., Sharp, S., Pryde, F., Kent, N.A., Mellor, J., & Louis, E.J.

Repressive and non-repressive chromatin at native telomeres in Saccharomyces

cerevisiae. Epigenetics & Chromatin. 2, 18. (2009)

Lunyak, V.A., & Rosenfeld, M.G. Epigenetic regulation of stem cell fate. Human

Molecular Genetics. 17, 28-36. (2008)

Lyon, M. F., Gene action in the X chromosome of the mouse (Mus musculus L.)

Nature. 190, 372–373. (1961)

Maher, M.J., Werner, E.E., and Denver, R.J. Stress hormones mediate predator-induced

phenotypic plasticity in amphibian tadpoles. Proc Biol Sci. 6, 1758. (2013)

Martienssen, R.A., Zaratiegui, M., & Goto, D.B. RNA interference and heterochromatin

in the fission yeast Schizosaccharomyces pombe. Trends Genet. 21, 450-6. (2005)

Mattick, J.S., & Makunin I.V. Small regulatory RNA in mammals. Hum Mol Genet. 14,

121-132. (2005)

47

Matzke, M.A., Mette, M.F. & Matzke, A.J. Transgene silencing by the host genome

defense: Implications for the evolution of epigenetic control mechanisms in plants

and vertebrates. Plant Mol. Biol. 43, 401-415. (2000)

Maxwell, A., McCudden, C.R., Wians, F.H., and Willis, M.S. Recent Advances in the

Detection of Prostate Cancer Using Epigenetic Markers in Commonly Collected

Laboratory Samples. Lab Med. 40, 171-78. (2009)

McCarrey, J.R. The epigenome as a target for heritable environmental disruptions of

cellular function. Molecular and cellular endocrinology. 354, 9-15. (2011)

Mellén, M., Ayata, P., Dewell, S., Kriaucionis, S., and Heintz, N. MeCP2 binds to

5hmC enriched within active genes and accessible chromatin in the nervous system.

Cell. 7, 1417-30. (2012)

Mikovits, J.A., Young, H.A., Vertino, P., Issa, J.P., Pitha, P.M., Turcoski-Corrales, S.,

Taub, D.D., Petrow, C.L., Baylin, S.B., Ruscetti, F.W. Infection with human

immunodeficiency virus type 1 upregulates DNA methyltransferase, resulting in de

novo methylation of the gamma interferon (IFN-gamma) promoter and subsequent

downregulation of IFN-gamma production. Mol Cell Biol. 18, 5166-77. (1998)

Mikula, B.C. Enviromental programming of heritable epigenetic changes in paramutant

r-gene expression using temperature and light at a specific stage of early

development in maize seedlings. Genetics. 140, 1379-1387. (1995)

Minocherhomji, S., Tollefsbol, T.O., & Singh, K.K. Mitochondrial regulation of

epigenetics and its role in human diseases. Epigenetics. 7, 326-334. (2012)

Miranda, T.B,. & Jones, P.A. DNA methylation: The nuts and bolts of repression. J.

Cell. Physiol. 213, 384–390. (2007)

Morgan, H.D., Santos, F., Green, K., Dean, W., Reik, W., Epigenetic reprogramming in

mammals. Hum. Mol. Genet. 14, R47-58. (2005)

Nakayama, J., Rice, J. C., Strahl, B. D., Allis, C. D., and Grewal, S. I. Role of histone

H3 lysine 9 methylation in epigenetic control of heterochromatin assembly. Science.

292, 110-113. (2001)

Nano, N., & Houry, W.A. Chaperone-like activity of the AAA+ proteins Rvb1 and

Rvb2 in the assembly of various complexes. Phil. Trans. R. Soc. B. 368, 20110399.

(2013)

48

Okae, H., Hiura, H., Nishida, Y., Funayama, R., Tanaka, S., Chiba, H., Yaegashi, N.,

Nakayama, K., Sasaki, H., Arima, T. Re-investigation and RNA sequencing-based

identification of genes with placenta-specific imprinted expression. Hum Mol Genet.

21, 548–558. (2012)

Obata, Y., Kaneko-Ishino, T., Koide, T., Takai, Y., Ueda, T., Domeki, I., Shiroishi, T.,

Ishino, F., Kono, T. Disruption of primary imprinting during oocyte growth leads to

the modified expression of imprinted genes during embryogenesis. Development.

125, 1553-1560. (1998)

Okano, M., Bell, D.W., Haber, D.A., Li, E. DNA methyltransferases Dnmt3a and

Dnmt3b are Essentials for the novo methylation and mammalian development. Cell.

99, 247-57. (1999)

Paulsen, M., El-Maarri, O., Engemann, S., Strödicke, M., Franck, O., Davies, K.,

Reinhardt, R., Reik, W., Walter J. Sequence conservation and variability of

imprinting in the Beckwith–Wiedemann syndrome gene cluster in human and mouse.

Hum. Mol. Genet. 9, 1829–1841. (2000)

Paschos, K., & Allday, M.J. Epigenetic reprogramming of host genes in viral and

microbial pathogenesis. Trends in Microbiology. 18, 10. (2010)

Plews, J.R., Gu, M., Longaker, M.T., and Wu, J.C. Large animal induced pluripotent

stem cells as pre-clinical models for studying human disease. J Cell Mol Med. 6,

1196-202. (2012)

Pratap, J., Akech, J., Wixted, J.J., Szabo, G., Hussain, S., McGee-Lawrence, M.E., Li,

X., Bedard, K., Dhillon, R.J., van Wijnen, A.J., Stein, J.L., Stein, G.S., Westendorf,

J.J., Lian, J.B. The histone deacetylase inhibitor, vorinostat, reduces tumor growth at

the metastatic bone site and associated osteolysis, but promotes normal bone loss.

Mol Cancer Ther. 9, 3210-20. (2010)

Price, T.D., Qvarnström, A., and Irwin, D.E. The role of phenotypic plasticity in driving

genetic evolution. Proc. R. Soc. 270, 1433-1440. (2003)

Reik, W., & Walter, J., Genomic imprinting: paternal influence on the genome. Nat.

Rev. Genet. 2, 21-32. (2001)

Rodríguez, M.D., Téllez, N.A., López, M., Cervantes, A., DNA methylation: an

epigentic process of medical importance. Rev Invest Clín. 56, 56-71. (2004)

49

Scheid, M., Afscar, O.K., and Paszkowski J: Formation of stable epialleles and their

paramutation-like interaction in tetraploid Arabidopsis thaliana. Nat Genet. 34, 450-

454. (2003)

Shemer, R., Birger, Y., Riggs, A. D. & Razin, A. Structure of the imprinted mouse

snrpn gene and establishment of its parental-specific methylation pattern. Proc. Natl

Acad. Sci. USA. 94, 10267–10272. (1997)

Shock, L.S., Thakkar, P.V., Peterson, E.J., Moran, R.G., Taylor, S.M. DNA

methyltransferase 1, cytosine methylation and cytosine hydroxymethylation in

mammalian mitochondria. Proceedings of the National Academy of Sciences. 108,

3630-5. (2011)

Shogren-Knaak, M., Ishii, H., Sun, J., Pazin, M.J., Davie, J.R., Peterson, C.L., Histone

H4-K16 Acetylation Controls Chromatin Structure and Protein Interactions. Science.

311, 844-847. (2006)

Sung, S. & Amasino, R.M. Remembering winter: Toward a molecular understanding of

vernalization. Annu Rev Plant Biol. 56, 491–508. (2005)

Suzuki, M.M., & A. Bird. DNA methylation landscapes: provocative insights from

epigenomics. Nat. Rev. Genet. 9, 465-476. (2008)

Stam, M., & Scheid, M. O. Paramutation: an encounter leaving a lasting impression.

Trends Plant Sci. 10, 283-290. (2005)

Stöger, R. et al. Maternal-specific methylation of the imprinted mouse Igf2r locus

identifies the expressed locus as carrying the imprinting signal. Cell. 73, 61–71.

(1993)

Strahl, B.D., & Allis, C.D. The language of covalent histone modifications. Nature.

403, 41-45. (2000)

Szutorisz, H., & Dillon, N. The epigenetic basis for embryonic stem cell pluripotency.

BioEssays. 27, 1286-1293. (2005)

Terranova, R., Agherbi, H., Boned, A., Meresse, S., Djabali, M. Histone and DNA

methylation defects at Hox genes in mice expressing a SET domaintruncated form of

Mll. Proc Natl Acad Sci. 103, 6629–6634. (2006)

Tomari, Y., & Zamore, P.D. Perspective: Machines for RNAi. Genes & Dev. 19, 517-

529. (2005)

50

Turner, B.M. Defining an epigenetic code. Nature Cell Biology. 9, 2-6. (2007).

Urnov, F.D., Wolffe, A.P. Chromatin remodeling and transcriptional activation: the cast

(in order of appearance). Oncogene. 20, 2991-3006. (2001)

Vaughn, M.W., Tanurdžić, M., Lippman, Z., Jiang, H., Carrasquillo, R., Rabinowicz,

P.D., Dedhia, N., McCombie, W.R., Agier, N., Bulski, A., Colot, V., Doerge, R.W.,

and Martienssen, R. Epigenetic Natural Variation in Arabidopsis thaliana. PLoS Biol.

5, 174. (2007)

Waddington, C.H. The basic ideas of biology. In: Towards a Theoretical Biology

Prolegomena. Edinburgh University Press. 1, 1-32. (1968)

Wakayama, T., Perry, A.C., Zuccotti, M., Johnson, K.R., Yanagimachi, R. Fullterm

development of mice from enucleated oocytes injected with cumulus cell nuclei.

Nature. 394, 369-374. (1998)

Weber, M., Davies, J.J., Wittig, D., Oakeley, E.J., Haase, M., Lam, W.L., Schübeler, D.

"Chromosome-wide and promoter-specific analyses identify sites of differential

DNA methylation in normal and transformed human cells". Nat. Genet. 37, 853–62.

(2005)

Weiner, S.A., Toth A.L. Epigenetics in social insects: a new direction for understanding

the evolution of castes. Genet Res Int, 609810. (2012)

Williams, M.D., Mitchell, G.M., and Hardikar, A.A. Stem Cells: Epigentic Basis of

Differentiation. The Open Stem Cell Journal. 3, 28-33. (2011)

Wilmut, I., Schieke, A.E., McWhir, J., Kind, A.J., Campbell, K.H. Viable offspring

derived from fetal and adult mammalian cells. Nature. 385, 810-813. (1997)

Wilson, A.S., Power, B.E., Molloy, P.L. DNA hypomethylation and human diseases.

Biochim Biophys Acta. 1775, 138-52. (2007)

Wu, C. T. & J. R. Morris. Genes, genetics, and epigenetics: a correspondence. Science.

293, 1103-1105. (2001)

Xiong, Z., & Laird, P.W. COBRA: a sensitive and quantitative DNA methylation assay

Nucl. Acids Res. 25, 532-2534. (1997)

51

Zhang, Y., Reinberg, D. Transcription regulation by histone methylation: interplay

between different covalent modifications of the core histone tails. Genes & Dev. 15,

2343-2360. (2001)

Books:

Strachan, T., and Read, A. Human Molecular Genetics, Garland Science, 4th Edition.

Published on April 02, 2010.

Prader-Willi Syndrome Synonyms: PWS, Prader-Labhart-Willi Syndrome. Bookshelf

ID: NBK1330PMID: 20301505. Initial Posting: October 6, 1998; Last Update:

October 11, 2012

Dagli, A.I., & Williams, C.A. Angelman Syndrome. Initial Posting: September 15,

1998; Last Update: June 16, 2011. Accessed at

http://www.ncbi.nlm.nih.gov/books/NBK1144/.

52