You are on page 1of 14

NEW RESEARCH HORIZON Review

Epigenetic regulation of endometrium


during the menstrual cycle
S.K. Munro
1,2
, C.M. Farquhar
3
, M.D. Mitchell
1,2,4
,
and A.P. Ponnampalam
1,*
1
The Liggins Institute, The University of Auckland, Private Bag 92019, Auckland 1142, New Zealand
2
The National Research Centre for Growth
and Development, c/- The Liggins Institute, The University of Auckland, Private Bag 92019, Auckland 1142, New Zealand
3
Department of
Obstetrics and Gynaecology, Faculty of Medical and Health Sciences, The University of Auckland, Private Bag 92019, Auckland 1142,
New Zealand
4
University of Queensland Centre for Clinical Research, RBWH Campus, Herston, Brisbane, QLD 4029, Australia
*Correspondence address. E-mail: a.ponnampalam@auckland.ac.nz
Submitted on November 3, 2009; resubmitted on December 22, 2009; accepted on February 2, 2010
abstract: The endometrium undergoes morphological and functional changes during the menstrual cycle which are essential for uterine
receptivity. These changes are driven by estrogen and progesterone and involve the ne control of many different genesseveral of which
have been identied as being epigenetically regulated. Epigenetic modication may therefore inuence the functional changes in the endo-
metrium required for successful implantation. There is, however, only limited information on epigenetic regulation in endometrium. We
review the potential role of epigenetic regulation of key processes during the menstrual cycle and present our own ndings following a pre-
liminary study into global acetylation levels in the human endometrium. A changing epigenetic state is associated with the differentiation of
stem cells into different lineages and thus may be involved in endometrial regeneration. Histone acetylation is implicated in the vascular endo-
thelial growth factor pathway during angiogenesis, and studies using histone deacetylase inhibitors suggest an involvement in endometrial
proliferation and differentiation. The processes of decidualization and implantation are also associated with epigenetic change and epigenetic
modulators show variable expression across the menstrual cycle. Our own studies found that endometrial global histone acetylation, as
determined by western blotting, changed throughout the menstrual cycle and correlated well with expected transcription activity during
the different phases. This suggests that epigenetics may be involved in the regulation of endometrial gene expression during the menstrual
cycle and that abnormal epigenetic modications may therefore be associated with implantation failure and early pregnancy loss as well as
with other endometrial pathologies.
Key words: endometrium / epigenetics / histone acetylation / menstrual cycle / uterus
Introduction
Implantation is a highly controlled process, involving a dialogue between
the endometrium and the implanting embryo which is crucial for the
establishment and maintenance of pregnancy (Norwitz et al., 2001;
Dey et al., 2004; Makker and Singh, 2006; Palis et al., 2007). Little is
known about the regulation of implantation (Makker and Singh,
2006), however inadequate uterine receptivity is thought to be respon-
sible for two-thirds of implantation failures (Simon et al., 1998). Control
of gene expression is crucial to the developmental processes which are
regularly implemented as the endometrium undergoes cyclic periods of
growth and differentiation. Inappropriate epigenetic modication result-
ing in aberrant expression of endometrial regulatory genes or proteins
may therefore provide a partial explanation for the failure of embryo
implantation (Kao et al., 2003).
Deregulation of implantation may have consequences beyond the
failure of implantation and subsequent infertility. In mice, even a transi-
ent postponement of blastocyst attachment is sufcient to cause a
detrimental ripple effect throughout the pregnancy, with aberrant
spacing of embryos, defective placentation, resorption and retarded
development of foetuses observed (Wilcox et al., 1999; Song et al.,
2002; Ye et al., 2005). Poor implantation and placentation have also
been associated with end results such as intrauterine growth restriction,
pre-eclampsia, and preterm birth in humans (Wang and Dey, 2006).
Epigenetic modication, resulting in altered chromatin structure and
transcriptional activity, controls gene expression and may therefore
control the functional changes in the endometrium which occur
throughout the menstrual cycle. Here, we examine the role epige-
netics plays in several key processes which occur during the menstrual
cycle and the establishment of pregnancy and discuss whether
normal human endometrium could be under epigenetic regulation.
Epigenetics
For mainly historical reasons, there has been some confusion about
what epigenetics actually refers to. A standard denition has been
mitotically and/or meiotically heritable changes in gene function
& The Author 2010. Published by Oxford University Press on behalf of the European Society of Human Reproduction and Embryology. All rights reserved.
For Permissions, please email: journals.permissions@oxfordjournals.org
Molecular Human Reproduction, Vol.16, No.5 pp. 297310, 2010
Advanced Access publication on February 5, 2010 doi:10.1093/molehr/gaq010

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

that cannot be explained by changes in gene sequence (Russo et al.,
1996). Of these, heritability is perhaps the most restrictive require-
ment: for example, although neurons undergo stable alterations,
since they rarely divide the changes are not heritable in the simplest
sense; whereas other alterations seem to be heritable, but are not
maintained in a stable manner. These issues led to a proposed de-
nition of epigenetics as the structural adaptation of chromosomal
regions so as to register, signal or perpetuate altered activity states
(Bird, 2007), and more recently; an epigenetic trait is a stably inher-
ited phenotype resulting from changes in a chromosome without
alterations in the DNA sequence (Berger et al., 2009).
Changes in chromatin architecture can be mediated by DNA
methylation and the post-translational modication of histone tails
(Dolinoy et al., 2007; Jirtle and Skinner, 2007). These histone modi-
cations are also involved in the recruitment of transcription factor
complexes (Berger, 2007) and therefore act to regulate gene
expression. It is thought that non-protein coding RNAs may also regu-
late chromatin structure, with RNA mediated interplay between the
environment, epigenome and transcriptome also occurring, however
this is still an emerging area of research (Mattick et al., 2009). Cellular
components such as DNA methyltransferases (DNMTs), histone dea-
cetylases (HDACs) and histone acetyltransferases (HATs) act to main-
tain or alter these modications as appropriate. Two of the most
well-known epigenetic-related phenomena are X chromosome inacti-
vation and genomic imprinting (Jirtle and Skinner, 2007). A range of
environmental factors, including nutrition and exposure to xenobiotic
chemicals, can inuence the establishment and maintenance of epige-
netic patterning and thus long-term gene transcription (Dolinoy et al.,
2007; Jirtle and Skinner, 2007).
Both DNA methylation and histone modications have been found
to be altered in human cancers within the promoter regions of tumour
suppressor genes and oncogenes (Jirtle and Skinner, 2007). It is
becoming increasingly clear that such changes are also associated
with other disease states (Guo, 2009; Trenkmann et al., 2009;
Turunen et al., 2009).
DNA methylation
DNA methylation is the covalent modication of post-replicative DNA
by the addition of a methyl group to the cytosine ring to form methyl
cytosine; a process which is catalyzed by DNMTs (Ohgane et al.,
2008). Three main DNMTs are present in mammals: DNMT1,
which maintains methylation by recognizing uni-strand methylation
after replication, and DNMTs 3a and 3b, which predominantly func-
tion as de novo methyltransferases acting at previously un-methylated
sites (Brero et al., 2006).
In mammals, DNA methylation is thought to be exclusively associ-
ated with cytosine-phosphate-guanine (CpG) dinucleotides, occurring
on both strands at the cytosine residue (Brero et al., 2006). In ver-
tebrates, large un-methylated GC rich regions can be found at the
5
0
end of many genes and are termed CpG islands (Bird, 1987).
DNA methylation at promoter associated CpG islands is strongly
linked to repression of transcriptional activity; however, the precise
role of CpG methylation is an area of ongoing research (Bird,
2002). Approximately 1020% of genes display DNA methylation
patterns in a tissue-specic manner (Song et al., 2005; Eckhardt
et al., 2006; Rakyan et al., 2008) and these so-called tissue-specic dif-
ferentially methylated regions have been associated with tissue-specic
patterns of gene expression (Rakyan et al., 2008). The repressive
nature of DNA methylation is thought to result from either the
methyl group inhibiting the binding of regulatory molecules to a
CpG site or by acting as a recognition signal for methyl-CpG-binding
proteins with repressive properties (Bird, 2002), both of which are
mechanisms likely to act in a site-specic manner. However, it
should be noted that DNA methylation is not always associated
with transcriptional repression. A recent study conrmed previous
ndings of a negative correlation between DNA methylation and
gene expression at promoters with a medium or high CpG density,
however, contrary to previous notions, even some low-CpG density
promoters showed this correlation. In contrast, gene-body methyl-
ation was found to positively correlate with gene expression suggesting
novel roles for DNA methylation (Rakyan et al., 2008). Further
research is required to elucidate the sequence and context-specic
nature of this epigenetic modication.
Histone modications
Histones are integral to the higher order structure of chromatin.
Approximately 146 bp of DNA is wrapped around a histone
octamer consisting of two sets of the core histonesH2A, H2B,
H3 and H4to form a nucleosome. These nucleosomes are linked
by loops of DNA and the linker histone H1. Although the globular
domains of histones associate closely with each other, the trailing
amino acid tails protrude past the surrounding DNA and are
subject to post-translational modication. New histone proteins are
produced during the S (synthesis) phase of the cell cycle when
DNA is being replicated (Lucchini and Sogo, 1995). The existing his-
tones from the parental strand are randomly segregated between par-
ental and daughter strands during replication and the rest of the
histones, synthesized de novo, are then deposited (Sogo et al.,
1986). Two H3H4 heterodimers are deposited either sequentially
or as a heterotetramer, followed by the two H2AH2B dimers,
thus forming the octamer (Worcel et al., 1978; English et al., 2006).
Many distinct modications of histones exist, with those identied
including lysine acetylation, lysine methylation, arginine methylation,
phosphorylation, ubiquitylation, sumoylation, ADP ribosylation, deimi-
nation and proline isomerization (Kouzarides, 2007; Lindner, 2008).
Modication of histones can be difcult to study as histone modi-
cations are very dynamic. All known histone post-translational modi-
cations have now been shown to be reversible, and as some histone
modications have been shown to change within minutes of a stimulus,
it is uncertain as to which modications can be considered as truly
stable and therefore epigenetic. It is also unknown how the persist-
ence of chromatin state is achieved, and which modications are
therefore truly heritable (Berger, 2007; Kouzarides, 2007).
Histone deacetylation, one of the best characterized histone modi-
cations, is associated with gene silencing, whereas histone acetylation
is associated with transcriptional activation (Fuks, 2005) (Fig. 1).
Histone acetylation patterns are maintained by two opposing groups
of enzymes, the HATs and HDACs (Hildmann et al., 2007). There
is also evidence for crosstalk between elements of the DNA methyl-
ation machinery and those involved in histone modication, with
DNMTs able to recruit complexes containing HDACs (Fuks, 2005).
A group of proteins with methyl binding domains are also able to
recognize methylated DNA sequences and recruit HDACs (Nan
et al., 1998; Irvine et al., 2002; Jorgensen and Bird, 2002). A reverse
298 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

regulatory pathway whereby histone acetylation mediates DNA
demethylation may also exist (Szyf et al., 1985; Selker, 1998;
Cervoni and Szyf, 2001) as some data show that HDAC inhibition
can lead to changes in DNA methylation (Cosgrove and Cox, 1990;
Chen and Pikaard, 1997; Hu et al., 1998; Selker, 1998; Hu et al.,
2000; Dobosy and Selker, 2001; Maass et al., 2002), although, this is
not always the case (Cameron et al., 1999; Singal et al., 2001;
El-Osta et al., 2002). HDACs and HATs, like many other enzymes
involved in histone modication, have actions which are not specic
to the histones and are able to modify other proteins. Over 60 tran-
scription factors including the tumour suppressor p53 are subjected to
acetylation and some HATs also physically associate with non-protein
coding RNA (Yang and Seto, 2007). Similar actions are seen with
other histone modifying enzymes, such as lysine-specic
demethylase-1, which in addition to histone demethylation, has
been found to be involved in the demethylation and stabilization of
DNMT1, which is required for the maintenance of global DNA
methylation (Hotz and Peters, 2009; Wang et al., 2009).
The menstrual cycle
Human endometrium undergoes cyclic morphological changes invol-
ving precise periods of growth, differentiation and regression, with
each new menstrual cycle starting again on average every 28 days
during a womans reproductive years (Curry and Osteen, 2003).
These changes are the visible result of the interactions of an array
of biological communication pathways and many local factors including
cytokines, which are under the overall control of the ovarian steroids
estrogen and progesterone (Curry and Osteen, 2003). Rising serum
estrogen levels from the ovary stimulate endometrial growth during
the proliferative phase. Epithelial and stromal cells from the functional
layer proliferate extensively, with both cell types increasingly acquiring
receptors to both estrogen and progesterone in parallel to the rise in
estrogen levels (Garcia et al., 1988). Following ovulation, usually occur-
ring around Day 14 of the menstrual cycle, there is a secretory trans-
formation of the endometrium in response to rising serum
progesterone levels produced by the corpus luteum (Brenner and
West, 1975; Lessey, 2000). Changes in the glandular epithelium are
apparent by cycle Days 1516 (Noyes et al., 1950; Lessey, 2000). Fol-
lowing fertilization, signals from the developing blastocyst act in
concert with steroid hormones to induce a receptive state supportive
of implantation (Paria et al., 2002; Makker and Singh, 2006). The endo-
metrial epithelium only transiently acquires this receptive state and a
variety of architectural, cellular, molecular and biochemical events
must be coordinated (Makker and Singh, 2006). The endometrial
stroma becomes more vascular and oedematous during this phase,
with the glands displaying enhanced secretory activity (Norwitz
et al., 2001). Additional morphological changes occur with the
process of decidualization by which stromal cells differentiate to
support pregnancy (Carson et al., 2000). The secretory phase will,
in the absence of an embryo, terminate with the shedding of the func-
tionalis layer approximately 2 weeks after ovulation, a process known
as menstruation (Carson et al., 2000). As menstruation results in
bleeding and tissue loss, it is followed by a period of repair and regen-
eration of the functionalis prior to the next cycle of proliferation. The
regenerative capacity of the endometrium is such that within a cycle it
grows to a thickness of 57 mm from the initial 0.51 mm post-
menstruation (McLennan and Rydell, 1965). Consequently, the
endometrium requires a continuing activation of processes such as
angiogenesis that are normally only associated with early development
(Curry and Osteen, 2003). As many of these processes have also been
Figure 1 Histone post-translational modications (PTMs) inuence chromatin structure. This leads to either a more condensed state, associated
with transcriptional repression or a more relaxed state associated with transcriptional activation. PTMs are the net result from activity of epigenetic
modulators which form large complexes with some adding, others removing different modications. Chromatin state is also linked to the degree of
DNA methylation with crosstalk between elements of the DNA methylation machinery and those involved in histone modication.
Epigenetic regulation of human endometrium 299

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

associated with epigenetic alterations, we propose that the endome-
trium may be subject to epigenetic regulation throughout the men-
strual cycle (Fig. 2).
Regeneration of the endometrium
The stratum basalis or basal layer of the endometrium remains intact
throughout the menstrual cycle and contains the cells from which the
functional layer will regenerate following menstruation (Noyes et al.,
1950; Curry and Osteen, 2003; Diedrich et al., 2007). The regenera-
tive capacity of human endometrium led to the proposal that there
might be a local population of adult stem cells contributing to the
regeneration of the functional layer (Padykula, 1991) (Fig. 2). Adult
stem cells are small populations of quiescent cells that undergo asym-
metrical cell divisions when activated, allowing the maintenance of
their own population in addition to the production of daughter
cellslike embryonic stem cellsare likely to be under the control
of DNA methylation and chromatin modications. These adult stem
cells/progenitor cells can proliferate extensively (Fuchs and Segre,
2000) differentiating to form the different cell types of that lineage
and represent a more differentiated version of the stem cells seen
in early development.
It is clear that there are major epigenetic alterations of the genome
during mammalian development and during differentiation of embryo-
nic stem cells (Reik, 2007). Embryonic stem cells differ in their global
gene expression prole from the more lineage restricted adult stem
cells, with a gradual loss of pluripotent gene expression and gain of
lineage-specic genes occurring during differentiation (Perry et al.,
2004; Yeo et al., 2007; Atkinson and Armstrong, 2008; Zardo et al.,
2008). Changes in chromatin architecture are thought to be respon-
sible for these altered gene expression proles and thus stem cell iden-
tity, with the epigenetic code determining a stem cells
responsiveness to developmental signals (Cirillo et al., 2002; Schrem
et al., 2002; Zaret et al., 2008; Collas, 2009). External stimuli such
as extracellular growth factors may also result in direct alteration of
the chromatin status, impeding or facilitating stem cell differentiation
(Song and Ghosh, 2004; Snykers et al., 2009). The unique properties
of stem/progenitor cells, pluripotent differentiation and self-renewal
seem to be under the control of DNA methylation and chromatin
modications (Shukla et al., 2008). In light of this, it is not surprising
that alteration of epigenetic state has been found to play an important
role in the reprogramming of cell fate, with studies showing that vir-
tually any cell can be regressed to a less mature state with the right
combination of factors (Cowan et al., 2005; Ivanova et al., 2006;
Takahashi et al., 2007; Mikkelsen et al., 2008; Zardo et al., 2008).
Of note is that most human cancers are associated with altered
epigenetic patterns in tissue stem/progenitor cells (Feinberg et al.,
2006) emphasizing the importance of epigenetics in control of cell
proliferation and differentiation.
The rst endometrial stem/progenitor cell candidates were ident-
ied in 2004, with a small population of both human endometrial epi-
thelial cells and stromal cells shown to possess clonogenic activity
in vitro (Chan et al., 2004). It has been suggested that these colony-
forming unit (CFU) cells might be responsible for the regeneration
of both cycling and atrophic endometrium (Gargett et al., 2008),
with CFU cells being found in normal cycling endometrium, inactive
peri-menopausal endometrium and in the endometrium of women
taking oral contraceptives (Schwab et al., 2005). The existence of epi-
thelial and stromal stem/progenitor cells in human and mouse endo-
metrium have been conrmed using stem cell activity assays (Gargett
et al., 2007; Gargett et al., 2009). Assays also recently identied differ-
ent CFU cells as being of an epithelial progenitor cell type (Gargett
et al., 2009), which is expected to reside somewhere in the basalis
layer (Gargett, 2007), and an endometrial mesenchymal stem/
progenitor cell type (Reynolds and Rietze, 2005; Dimitrov et al.,
2008; Gargett et al., 2009) recently localized to perivascular niches
in both the basal and functional layers of the endometrium (Schwab
and Gargett, 2007). Epithelial side population cells have also
been identied in human endometrial cell suspensions and cultures
Figure 2 Many individual events during the menstrual cycle are associated with epigenetic change. Epigenetics has been implicated in the processes
of stem cell differentiation, angiogenesis, implantation and decidualization, and it is likely that given the dynamic nature of the endometrium, that it is
under epigenetic regulation.
300 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

(Kato et al., 2007; Tsuji et al., 2008) and are reported to be capable of
reconstituting various endometrial components and structures when
xenotransplanted into mice (Maruyama, 2009).
It has been proposed that inappropriate shedding of endometrial
progenitor cells may contribute to the development of endometriotic
lesions, with evidence of monoclonality (Jimbo et al., 1997; Starzinski-
Powitz et al., 2001; Leyendecker et al., 2002; Tanaka et al., 2003; Wu
et al., 2003; Gargett, 2007; Gargett et al., 2008). With a change in
environment following retrograde menstruation potentially altering
both paracrine and endocrine signalling, it is likely that a number of
epigenetic changes arise in ectopic endometrium, consequently
giving rise to the pathology of endometriosis. Indeed, endometriosis
is increasingly being identied as an epigenetic disease and the evi-
dence for this has recently been reviewed (Guo, 2009). Evidence of
distinct epigenetic changes arising in endometriosis suggests a mechan-
ism by which key endometrial genes involved in proliferation and
differentiation may already be under epigenetic control, with acti-
vation/repression in response to environmental cues. It is likely that
epigenetic changes assist in the regulation of progenitor cell differen-
tiation into the variety of lineage-restricted endometrial cell types
required during regeneration. Interestingly, the initial stages of endo-
metrial repair do not appear to require estrogen. Within 48 h of shed-
ding, when circulating estrogen levels are low, epithelial cells migrate
over the denuded surface of the endometrium (Okulicz and Scarrell,
1998). At this time, endometrial epithelial cells also lack the
expression of estrogen receptor alpha (ERa). A mouse model of
endometrial restoration following menstruation has conrmed regen-
eration can occur in the absence of estrogen (Kaituu-Lino et al.,
2007). Rapid proliferation of epithelial label-retaining cells in response
to estrogen despite the lack of receptor indicates a transmission of
proliferative signals from stromal niche cells (which do express ERa)
to candidate stem/progenitor cells (Chan and Gargett, 2006;
Gargett, 2007) and exemplies the importance of local environment
in the control of endometrial repair and regeneration.
Proliferation
During the proliferative phase, rising serum estrogen levels from the
ovary stimulate endometrial growth with extensive proliferation of
the epithelial and stromal cells of the functional layer, which increas-
ingly acquire receptors to both estrogen and progesterone (Garcia
et al., 1988). Studies into endometrial cancer have allowed us some
insight into the potential for epigenetic regulation of normal endo-
metrial proliferation. For example promoter hypo-methylation of the
transcription factor paired-box gene 2 (PAX2) has been identied in
endometrial cancer, with 75% of endometrial carcinoma in one
study exhibiting PAX2 promoter hypo-methylation (Wu et al.,
2005a). This hypo-methylation was found to allow inappropriate
activation by estrogen and tamoxifen, and resulted in enhanced cell
proliferation (Wu et al., 2005a). This suggests that endometrial
responsiveness to estrogen signalling may be dependent not only on
receptor expression but also on the appropriate epigenetic modi-
cation of other effectors. Receptiveness to steroid hormones and
their antagonists is a particular problem with endometrial tissue, par-
ticularly in patients receiving treatment for breast cancer. Tamoxifen, a
selective ER modulator which acts as an ER antagonist in the breast
and is used for breast cancer therapy, acts as a partial agonist in the
endometrium (Gallo and Kaufman, 1997) leading to an increased
risk of developing uterine adenocarcinoma (Fisher et al., 1994). Treat-
ment with the HDAC; Valproic acid (VPA) has been shown to block
the proliferation of uterine endometrial cells exposed to both tamox-
ifen and estrogen in Ishikawa cells (Hodges-Gallagher et al., 2007). The
Ishikawa cell line is known to be heterogeneous (Nishida et al., 1996),
which may account for reports of VPA both stimulating (Graziani et al.,
2003) and inhibiting (Takai et al., 2004) proliferation in Ishikawa cells,
but this does indicate a potential link between endometrial prolifera-
tive capacity and histone acetylation status.
Angiogenesis
Angiogenesis is the process of new blood vessel growth and occurs reg-
ularly during every menstrual cycle in human endometrium, with the
growth and regression of blood vessels occurring under the inuence
of estrogen and progesterone. The regulation of angiogenesis in the
endometrium has been difcult to elucidate, with a large number of
angiogenic factors being secreted. The direct correlation of these
factors with endometrial angiogenesis is hard to conrm as endothelial
cell proliferation, a marker for angiogenic activity, occurs throughout
the menstrual cycle (Gargett and Rogers, 2001; Rogers et al., 2009). In
animal models, endothelial cell proliferation is closely linked to circulating
levels of estrogen and progesterone; however, regulation is not so simple
in humans, with evidence that estrogen can both promote and inhibit
angiogenesis under different circumstances (Girling and Rogers, 2005).
In mouse models, endothelial cell proliferation occurs following the
elevation of estrogen levels post-estrus, and in conjunction with rising
plasma progesterone levels (Walter et al., 2005). In ovariectomized
mice, this proliferation can be induced to occur within 24 hours by
injection with 100 ng of estrogen 7 days post-surgery (Heryanto and
Rogers, 2002). Treatment with either antibodies against vascular
endothelial growth factor (VEGF) or VEGF receptor 2 inhibitors, com-
pletely blocks this response (Heryanto et al., 2003), suggesting that
estrogen acts through the VEGF pathway. An inux of VEGF expres-
sing neutrophils into the endometrium also occurs following estrogen
injection, which when blocked, results in decreased proliferation of
endothelial cells in response to estrogen implicating circulating leuco-
cytes in the mediation of this response (Heryanto et al., 2004). Pro-
gesterone injection alone, without estrogen priming, also stimulated
endothelial cell proliferation, although this effect is only mediated in
part through the VEGF pathway (Walter et al., 2005).
In the human endometrium, VEGF is the principle angiogenic factor
and its action is not restricted to vascular smooth muscle cells
(Charnock-Jones et al., 1993). VEGF mRNA is expressed mainly in
the stroma during the proliferative phase (Li et al., 1994; Kawano
et al., 2000), with a subset of stromal cells exhibiting much higher
expression, and a low level of expression is also observed in the gland-
ular epithelium (Charnock-Jones et al., 1993). During the secretory
phase stromal VEGF expression is more uniform and glandular
expression is increased (Charnock-Jones et al., 1993), with overall
expression three to ve times higher in the secretory phase when
compared with the early proliferative phase (Charnock-Jones et al.,
1994). A high level of VEGF mRNA can also be detected in menstrual
tissue (Charnock-Jones et al., 1993). Focal VEGF is also located in neu-
trophils associated with micro-vessel walls (Gargett et al., 2001). In
stromal cells, VEGF production is accompanied by the production of
matrix metalloproteinases (MMPs) which are involved in extracellular
matrix degradation and regulation of angiogenesis (Li et al., 1994;
Epigenetic regulation of human endometrium 301

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

Moses and Langer, 1991; Kawano et al., 2000). Treatment with
thrombin leads to increased VEGF, MMP-1 and active MMP-2, and
this up-regulation appears to act through PAR-1 (a member of the
G-protein-coupled protease-activated receptors) and mitogen-
activated protein kinase (Furukawa et al., 2009).
Evidence fromother tissues suggests that epigenetic mechanisms play
a role in the process of angiogenesis. For example, studies have shown
that VEGF actions on endothelial cell proliferation are mediated through
regulation of the HDAC-7. Like the other mammalian Class II HDACs,
HDAC-7 is responsive to extracellular signals and contains an N-
terminal extension that interacts with other transcription cofactors
(Verdin et al., 2003; Wang et al., 2008). Phosphorylation of conserved
serine residues in the N-terminal domain result in the removal of this
class of HDACs from the nucleus to the cytoplasm, activating gene
expression (Grozinger and Schreiber, 2000; McKinsey et al., 2000;
Vega et al., 2004; Mottet et al., 2007; Ha et al., 2008; Wang et al.,
2008), and blockade of HDAC-7 phosphorylation represses VEGF-
mediated endothelial cell proliferation and migration (Wang et al.,
2008). HDAC-7 can maintain vascular integrity by repressing MMP-10
expression in endothelial cells and in vitro assays support a role for
the modulation of angiogenesis by HDAC-7 (Chang et al., 2006;
Mottet et al., 2007). Inhibition of HDAC-7 gene transcription results
in altered endothelial cell morphology, migration and capillary forming
capacity but does not affect proliferation, adhesion or apoptosis
(Mottet et al., 2007). It also resulted in the up-regulation of platelet-
derived growth factor-B (PDGF-B) and its receptor. Inhibition of
HDACs blocks the differentiation of endothelial progenitor cells and
this appears to act through the down-regulation of homeobox
protein HOXA9, as over expression of HOXA9 partially rescues the
differentiation blockade observed after HDAC inhibitor treatment
(Rossig et al., 2005). HOXA9 appears to act as a regulator of
endothelial-committed genes such as endothelial nitric oxide synthase,
VEGFR-2 and VE-cadherin, and also mediates the shear stress-induced
maturation of endothelial cells (Rossig et al., 2005). This implicates
histone acetylation in the control of angiogenesis, with VEGF signalling
mediated through HDAC-7 resulting in the up-regulation of a number
of genes critical to angiogenesis, and it is possible that similar mechan-
isms may be acting in the endometrium (Fig. 2). HOXA9 mRNA is
expressed in human endometrium with a decrease in expression
reported between the early and mid-secretory phase (Carson et al.,
2002). Interestingly, the HOXA9 gene in endometrium has been
shown to be methylated in women with ovarian cancer, which is
thought to result from abnormal methylation of HOXA9 in mullerian
tract stem cells (Widschwendter et al., 2009). PDGF-B transcripts
have been identied in human endometrium (Boehm et al., 1990),
and both PGDF-B and MMP10 are highly expressed by a population
of cells isolated frommenstrual blood termed endometrial regenerative
cells (ERC) which were capable of differentiation into a number of
lineages including epithelial and endothelial lineages (Meng et al.,
2007). In addition, MMP10 production by endometrial stromal cells
has been shown to increase in response to conditioned media from
human trophoblasts (Hess et al., 2007), which may assist in vascular
remodelling associated with placentation.
Differentiation
Rising serum progesterone levels post-ovulation result in a secretory
transformation of the newly regenerated endometrium as the different
cell types begin to differentiate (Brenner and West, 1975; Lessey,
2000). Studies involving Ishikawa cells suggest a role for histone acety-
lation in the control of endometrial differentiation. Although Ishikawa
cells are a well-differentiated adenocarcinoma cell line rather than a
normal endometrial cell line type (Nishida, 2002), treatment with
the related HDAC inhibitors Trichostatin A (TSA) and suberoylanilide
hydroxamic acid (SAHA), results in differentiation which closely
resemble normal secretory phase endometrial epithelium in a time-
and dose-dependent manner. This is accompanied by the expression
of secretory phase-specic proteins (Uchida et al., 2005b) and the
effects on proliferation and differentiation by TSA and SAHA were
comparable to treatment with estrogen and progesterone suggesting
a parallel between histone hyper-acetylation, differentiation and
secretory phase hormone exposure. Surprisingly, the HDAC inhibitor
effects could be blocked by gene silencing of glycodelin, an established
differentiation marker for endometrial glandular cells (Seppala et al.,
2002) which suggests some level of feedback.
Decidualization is a process of differentiation of the broblast-like
stromal cells of the endometrium and is necessary to support the
developing embryo (Sakai et al., 2003). These cells, primed by estro-
gen in the proliferative stage, respond to progesterone by becoming
larger and rounder and initiating a range of biochemical pathways.
Decidual cells arise rst in the vicinity of the spiral arteries, and
begin to spread throughout the endometrium during the secretory
phase of the menstrual cycle. Following implantation of an embryo,
decidualization extends throughout the stroma, resulting in the for-
mation of the decidua characteristic of pregnancy (Noyes et al.,
1950; Sakai et al., 2003).
The process of decidualization has been associated with epigenetic
changes in vitro (Fig. 2). HDAC inhibitor treatment of human endo-
metrial stromal cells with TSA resulted in morphological change and
expression of differentiation marker proteins such as insulin-like
growth factor (IGFBP-1) and prolactin, indicative of decidualization
(Sakai et al., 2003). It was found that histones H3 and H4 became
acetylated upon decidualization and acetylated H4 in turn was associ-
ated with the ovarian steroid induced promoter activation of IGFBP-1
(Sakai et al., 2003). In addition, human endometrial adenocarcinoma
cell lines treated with TSA or SAHA have been shown to undergo
morphological transformation to a attened and widespread epithelial
phenotype, with induction of differentiation markers such as glycode-
lin, leukaemia inhibitory factor (LIF), interleukin-1 receptor and glyco-
gen synthesissimilar to responses seen with ovarian steroid
hormones (Uchida et al., 2005a).
Implantation
With epigenetics implicated in the control of decidualization (Sakai
et al., 2003; Uchida et al., 2007b), it is likely that epigenetic regulation
plays a role in the establishment of endometrial receptivity and in sub-
sequent embryo implantation (Fig. 2). DNA methylation and histone
acetylation have been directly correlated with the expression of
implantation related genes in other contexts, including LIF (Uchida
et al., 2005b), HOXA10 (Wu et al., 2005b; Yoshida et al., 2006), gly-
codelin (Uchida et al., 2005b; Uchida et al., 2007a), MMPs, tissue
inhibitors of matrix metalloproteinase (Clark et al., 2007), E-cadherin
(Yoshiura et al., 1995; Saito et al., 2003; Rahnama et al., 2006;
Rahnama et al., 2009) and mucin (MUC)1 (Yamada et al., 2008).
The steroid hormone receptors themselves are susceptible to
302 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

epigenetic modulation with promoter associated CpG islands
(Campan et al., 2006) and are capable of inducing the modication
to the chromatin structure of other genes (Leu et al., 2004). In turn
these epigenetically modied genes are involved in the expression of
MUC1 (Horne et al., 2006), the b3 subunit (Achache and Revel,
2006), Osteopontin (Makker and Singh, 2006), heparin-binding epi-
dermal growth factor (Makker and Singh, 2006) and MMPs and has
been linked to the formation of pinopodes (Bagot et al., 2001;
Aghajanova et al., 2003), with many secondary interactions between
these factors. It is conceivable that many other genes involved are
similarly regulated.
Expression of epigenetic modulators during
the menstrual cycle
Signicant changes in the expressions of DNMTs have been observed
in endometrium during the menstrual cycle with a reported decrease
in DNMT mRNA during mid-late secretory phase, and following
estrogen and medroxy-progesterone acetate treatment in stromal
cell cultures (Yamagata et al., 2009). DNMT protein levels in human
endometrium remain less clear, with reports of no change in
DNMT expression during the menstrual cycle by one group (Yamagata
et al., 2009) and that DNMT1 expression is restricted to the prolifera-
tive phase by another (Liao et al., 2008). In proliferative endometrium,
immunostaining reveals a high level of methylated cytosine residues in
cells of the glandular epithelium and the stroma. In the secretory phase
levels of methylated cytosines in the cells of the glandular epithelium
decrease, while there is an increase in the stroma (Ghabreau et al.,
2004).
A recent study investigated the expression of class 1 HDACs and
two HATs in human endometrium (Krusche et al., 2007). The
HDAC13 were found to have constitutive mRNA expression
throughout the menstrual cycle, likewise the mRNA of one of the
HATs; p300/CBP-associated factor is also expressed in human endo-
metrium without cyclic changes. In contrast, mRNA expression of
another HAT; general control non-derepressible 5 was reduced
during secretory phase compared with the proliferative phase
(Krusche et al., 2007). The same study found that HDAC1, and
HDAC3 protein expression was constitutive throughout the menstrual
cycle, though HDAC1 protein expression was found to be highly vari-
able between individuals (Krusche et al., 2007). HDAC2 protein,
however, was found to be slightly but signicantly elevated during
the secretory phase (Krusche et al., 2007).
New developments
With epigenetic regulation seemingly involved in many of the pro-
cesses occurring in human endometrium during the menstrual cycle,
and some evidence for altered expression of epigenetic modulators
both during the menstrual cycle and in endometrial pathologies, it is
of interest to discover whether global epigenetic change occurs
during the menstrual cycle. Certainly, in rats, injection of estradiol
results in increased histone acetylation in the uterus (Libby, 1972;
Guo and Gorski, 1989) suggesting a role by which the ovarian
steroid hormones could act through chromatin alterations to effect
gene transcription. Likewise, it has been shown that HDAC inhibitor
treatment can enhance the proliferative and morphogenetic actions
of estrogen in mice (Gunin et al., 2005).
We recently began investigating global histone acetylation levels
during the menstrual cycle in the human endometrium, constructing a
preliminary prole of global histone acetylation levels during the men-
strual cycle. Little information has been available on histone acetylation
in the endometrium with ours being the rst study to report on levels of
histone acetylation and how they alter during the menstrual cycle. To
date, two studies have investigated the expression of epigenetic modu-
lators during the menstrual cycle, with a report on the expression of
class 1 HDACs (Krusche et al., 2007) and recently on the expression
of DNMTs (Yamagata et al., 2009), but how these results correlate
with the global histone acetylation or DNA methylation status of the
endometrium during the menstrual cycle has not yet been elucidated.
We found that global histone acetylation levels of H2AK5, H3K9
and H4K8 were increased in the early proliferative phase, sub-
sequently declining until ovulation, a trend shared by H3K14/18
(Figs 35). Histone acetylation is associated with transcription acti-
vation, so increased levels are consistent with the initiation of many
genes and pathways that would be required to regenerate the endo-
metrium following menstruation. Once regeneration has occurred
and proliferation has begun, it is expected that many of these pathways
are no longer required, so a subsequent decrease in acetylation levels
is not surprising. Likewise, the increase in acetylation post-ovulation
makes sense in light of the switch from proliferation to differentiation
by many of the cell types of the endometrium. This increase was stat-
istically signicant for H4K8 and a marked trend is observed for H3K9
and H4K14/18 which is supported by a previous study (Sakai et al.,
2003) that found that endometrial stromal cells cultured with estrogen
and progesterone exhibited a mild but signicant increase in the acety-
lation of H4K8 and H3K9/14. A decline in global acetylation levels in
the late secretory phase is also expected with the regression of the
corpus luteum and breakdown of the endometrium in the absence
of pregnancy. Global H2BK12 acetylation does not appear to share
the trend seen with the other acetylation sites during the proliferative
phase; however, the extent to which both H2A and H2B acetylation
contribute to transcriptional activation is less understood with most
studies focusing on histones H3 and H4.
H4K5 acetylation did not change during the menstrual cycle, although
this was expected as H4K5 (and H4K12; not examined in this study) is
associated with the deposition of newly transcribed histones onto
DNA, requiring the presence of histone chaperones (Verreault,
2000). The initial acetylation of histones following transcription is
thought to be important for their assembly into nucleosomes by
histone chaperones (Shahbazian and Grunstein, 2007). In many eukar-
yotes, newly synthesised histone H4 is acetylated at H4K5 and
H4K12 (Sobel et al., 1995). Hence, constant levels of H4K5 acetylation
could reect a maintained proportion of new histone when compared
with total histone throughout the menstrual cycle.
It is noted that these are preliminary data and not without limit-
ations. In particular, these data relate to whole endometrial tissue,
and therefore data relates to a mixed population of cells, the relative
mix of which will be changing somewhat during the menstrual cycle.
Experiments are continuing to address these concerns and to eluci-
date which changes result from alterations in steroid hormone
levels. In addition, as global bulk acetylation levels were examined,
data are not restricted to particular regions of the genome. Localized
Epigenetic regulation of human endometrium 303

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

increases in acetylation may therefore have been masked in these
studies, and promoter-specic analyses (e.g. chromatin immunopreci-
pitation) will be required to determine whether this is the case.
Implications for future research/clinical
translation
Many studies have now examined the changes in endometrial gene
expression during the menstrual cycle both from freshly frozen
samples and using cell culture experiments (Popovici et al., 2000;
Brar et al., 2001; Carson et al., 2002; Kao et al., 2002; Borthwick
et al., 2003; Okada et al., 2003; Riesewijk et al., 2003;
Tierney et al., 2003; Ponnampalam et al., 2004; Mirkin et al., 2005;
Punyadeera et al., 2005; Talbi et al., 2006), often with little correlation
between studies. If epigenetic changes are indeed occurring during the
menstrual cycle then this could explain the large numbers of genes
showing differential expression across the cycle and between
studies. In particular, the effect of epigenetic alterations in cancer
cell lines must be considered when elucidating endometrial function,
with many studies to date using Ishikawa and other such cell lines
when examining events such as implantation.
Figure 4 Global acetylation of histone H3 at lysine 9 (A) and at
lysines 14 and 18 (B), in human endometrium during the menstrual
cycle. Changes in histone acetylation status were determined by
western blotting and data were normalized to histone H3 protein
levels and given as mean optical densities [relative to a sodium buty-
rate (HDAC inhibitor) treated positive control] +SEM. Statistical sig-
nicance denoted by: *P ,0.05. Cycle stages are early proliferative
[EP, (A) n 4, (B) n 5], mid-proliferative (MP, n 9), late prolif-
erative (LP, n 3), early secretory (ES, n 5), mid-secretory (MS,
n 4) and late secretory (LS, n 7) as determined by
histopathology.
Figure 3 Global acetylation of histones H2A (lysine 5) (A) and
H2B (lysine 12) (B), in human endometrium during the menstrual
cycle. Changes in histone acetylation status were determined by
western blotting and data were normalized to histone H2A/B
protein levels and given as mean optical densities [relative to a
sodium butyrate (HDAC inhibitor) treated positive control] +SEM.
Statistical signicance denoted by: **P ,0.01, ***P ,0.001. Cycle
stages are early proliferative [EP, (A) n 3, (B) n 4], mid-
proliferative (MP, n 9), late proliferative (LP, n 3), early secretory
(ES, n 5), mid-secretory (MS, n 4) and late secretory (LS, n 7)
as determined by histopathology. Numbers above bars indicate
P-values for non-signicant trends.
304 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

In addition, many studies into endometrial abnormalities compare
epigenetic state of the affected tissue/cells with normal endome-
trium. However, little regard is given to endometrial cycle stage,
with data invariably being pooled if no change is seen between the
proliferative and secretory phases. Given the apparent changes
within these phases reported here, much potential insight into these
disorders could be going unidentied.
Epigenetics appears to play a key role in developmental processes
that occur during the menstrual cycle and the establishment of preg-
nancy. It is potentially involved in the initial regeneration of the endo-
metrium through the changes in the epigenetic prole of stem cells and
is likely to be involved in endometrial proliferation and angiogenesis.
Evidence for altered epigenetic modulators or marks can be seen in
cancer and endometriosis, and epigenetics is also implicated in the
process of decidualization. Modulation of epigenetic state may there-
fore represent an important means of elucidating specic functions and
lead to therapeutic intervention.
Ethics
Ethical approval for the study was obtained from Northern X Ethics
Committee (NTX/08/02/008).
Acknowledgements
We thank the staff of Auckland City Surgical Services, Auckland City
Hospital and Greenlane Clinical Centre, Auckland, New Zealand for
their assistance in collection of the endometrial tissues, the clinicians
and patients involved, and the staff of Diagnostic Medlab for assistance
with histopathology of samples. We also thank Dr Kevin Dudley for
critical review of this manuscript.
Funding
This work was supported by the National Research Centre for
Growth and Development; a James Cook Research Fellowship to
MDM; and the Foundation for Research, Science and Technology
[grant# UOAX 0814].
References
Achache H, Revel A. Endometrial receptivity markers, the journey to
successful embryo implantation. Hum Reprod Update 2006;12:731746.
Aghajanova L, Stavreus-Evers A, Nikas Y, Hovatta O, Landgren BM.
Coexpression of pinopodes and leukemia inhibitory factor, as well as
its receptor, in human endometrium. Fertil Steril 2003;79:808814.
Atkinson S, Armstrong L. Epigenetics in embryonic stem cells: regulation of
pluripotency and differentiation. Cell Tissue Res 2008;331:2329.
Bagot CN, Kliman HJ, Taylor HS. Maternal Hoxa10 is required for pinopod
formation in the development of mouse uterine receptivity to embryo
implantation. Dev Dyn 2001;222:538544.
Berger SL. The complex language of chromatin regulation during
transcription. Nature 2007;447:407412.
Berger SL, Kouzarides T, Shiekhattar R, Shilatifard A. An operational
denition of epigenetics. Genes Dev 2009;23:781783.
Bird AP. CpG islands as gene markers in the vertebrate nucleus. Trends
Genet 1987;3:342347.
Bird A. DNA methylation patterns and epigenetic memory. Genes Dev
2002;16:621.
Bird A. Perceptions of epigenetics. Nature 2007;447:396398.
Boehm KD, Daimon M, Gorodeski IG, Sheean LA, Utian WH, Ilan J.
Expression of the insulin-like and platelet-derived growth factor genes
in human uterine tissues. Mol Reprod Dev 1990;27:93101.
Borthwick JM, Charnock-Jones DS, Tom BD, Hull ML, Teirney R,
Phillips SC, Smith SK. Determination of the transcript prole of
human endometrium. Mol Hum Reprod 2003;9:1933.
Brar AK, Handwerger S, Kessler CA, Aronow BJ. Gene induction and
categorical reprogramming during in vitro human endometrial
broblast decidualization. Physiol Genomics 2001;7:135148.
Figure 5 Global acetylation of histone H4 at lysines 5 (A) and 8
(B), in human endometrium during the menstrual cycle. Changes in
histone acetylation status were determined by western blotting and
data were normalized to histone H4 protein levels and given as
mean optical densities [relative to a sodium butyrate (HDAC inhibi-
tor) treated positive control] +SEM. Statistical signicance denoted
by: *P , 0.05, ***P ,0.001. Cycle stages are early proliferative (EP,
n 4), mid-proliferative (MP, n 9), late proliferative (LP, n 3),
early secretory (ES, n 5), mid-secretory (MS, n 4) and late
secretory (LS, n 7) as determined by histopathology.
Epigenetic regulation of human endometrium 305

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

Brenner RM, West NB. Hormonal regulation of the reproductive tract in
female mammals. Annu Rev Physiol 1975;37:273302.
Brero A, Leonhardt H, Cardoso MC. Replication and translation of
epigenetic information. Curr Top Microbiol Immunol 2006;301:2144.
Cameron EE, Bachman KE, Myohanen S, Herman JG, Baylin SB. Synergy of
demethylation and histone deacetylase inhibition in the re-expression of
genes silenced in cancer. Nat Genet 1999;21:103107.
Campan M, Weisenberger DJ, Laird PW. DNA methylation proles of
female steroid hormone-driven human malignancies. Curr Top Microbiol
Immunol 2006;310:141178.
Carson DD, Bagchi I, Dey SK, Enders AC, Fazleabas AT, Lessey BA,
Yoshinaga K. Embryo implantation. Dev Biol 2000;223:217237.
Carson DD, Lagow E, Thathiah A, Al-Shami R, Farach-Carson MC,
Vernon M, Yuan L, Fritz MA, Lessey B. Changes in gene expression
during the early to mid-luteal (receptive phase) transition in human
endometrium detected by high-density microarray screening. Mol
Hum Reprod 2002;8:871879.
Cervoni N, Szyf M. Demethylase activity is directed by histone acetylation.
J Biol Chem 2001;276:4077840787.
Chan RW, Gargett CE. Identication of label-retaining cells in mouse
endometrium. Stem Cells 2006;24:15291538.
Chan RW, Schwab KE, Gargett CE. Clonogenicity of human endometrial
epithelial and stromal cells. Biol Reprod 2004;70:17381750.
Chang S, Young BD, Li S, Qi X, Richardson JA, Olson EN. Histone
deacetylase 7 maintains vascular integrity by repressing matrix
metalloproteinase 10. Cell 2006;126:321334.
Charnock-Jones DS, Sharkey AM, Rajput-Williams J, Burch D, Schoeld JP,
Fountain SA, Boocock CA, Smith SK. Identication and localization of
alternately spliced mRNAs for vascular endothelial growth factor in
human uterus and estrogen regulation in endometrial carcinoma cell
lines. Biol Reprod 1993;48:11201128.
Charnock-Jones DS, Sharkey AM, Boocock CA, Ahmed A, Plevin R,
Ferrara N, Smith SK. Vascular endothelial growth factor receptor
localization and activation in human trophoblast and choriocarcinoma
cells. Biol Reprod 1994;51:524530.
Chen ZJ, Pikaard CS. Epigenetic silencing of RNA polymerase I
transcription: a role for DNA methylation and histone modication in
nucleolar dominance. Genes Dev 1997;11:21242136.
Cirillo LA, Lin FR, Cuesta I, Friedman D, Jarnik M, Zaret KS. Opening of
compacted chromatin by early developmental transcription factors
HNF3 (FoxA) and GATA-4. Mol Cell 2002;9:279289.
Clark IM, Swingler TE, Young DA. Acetylation in the regulation of
metalloproteinase and tissue inhibitor of metalloproteinases gene
expression. Front Biosci 2007;12:528535.
Collas P. Epigenetic states in stem cells. Biochim Biophys Acta 2009;
1790:900905.
Cosgrove DE, Cox GS. Effects of sodium butyrate and 5-azacytidine on
DNA methylation in human tumor cell lines: variable response to
drug treatment and withdrawal. Biochim Biophys Acta 1990;1087:8086.
Cowan CA, Atienza J, Melton DA, Eggan K. Nuclear reprogramming of
somatic cells after fusion with human embryonic stem cells. Science
2005;309:13691373.
Curry TE Jr, Osteen KG. The matrix metalloproteinase system: changes,
regulation, and impact throughout the ovarian and uterine
reproductive cycle. Endocr Rev 2003;24:428465.
Dey SK, Lim H, Das SK, Reese J, Paria BC, Daikoku T, Wang H. Molecular
cues to implantation. Endocr Rev 2004;25:341373.
Diedrich K, Fauser BC, Devroey P, Griesinger G. The role of the
endometrium and embryo in human implantation. Hum Reprod Update
2007;13:365377.
Dimitrov R, Timeva T, Kyurkchiev D, Stamenova M, Shterev A, Kostova P,
Zlatkov V, Kehayov I, Kyurkchiev S. Characterization of clonogenic
stromal cells isolated from human endometrium. Reproduction 2008;
135:551558.
Dobosy JR, Selker EU. Emerging connections between DNA methylation
and histone acetylation. Cell Mol Life Sci 2001;58:721727.
Dolinoy DC, Weidman JR, Jirtle RL. Epigenetic gene regulation: linking
early developmental environment to adult disease. Reprod Toxicol
2007;23:297307.
Eckhardt F, Lewin J, Cortese R, Rakyan VK, Attwood J, Burger M, Burton J,
Cox TV, Davies R, Down TA et al. DNA methylation proling of human
chromosomes 6, 20 and 22. Nat Genet 2006;38:13781385.
El-Osta A, Kantharidis P, Zalcberg JR, Wolffe AP. Precipitous release of
methyl-CpG binding protein 2 and histone deacetylase 1 from the
methylated human multidrug resistance gene (MDR1) on activation.
Mol Cell Biol 2002;22:18441857.
English CM, Adkins MW, Carson JJ, Churchill ME, Tyler JK. Structural basis
for the histone chaperone activity of Asf1. Cell 2006;127:495508.
Feinberg AP, Ohlsson R, Henikoff S. The epigenetic progenitor origin of
human cancer. Nat Rev Genet 2006;7:2133.
Fisher B, Costantino JP, Redmond CK, Fisher ER, Wickerham DL,
Cronin WM. Endometrial cancer in tamoxifen-treated breast cancer
patients: ndings from the National Surgical Adjuvant Breast and
Bowel Project (NSABP) B-14. J Natl Cancer Inst 1994;86:527537.
Fuchs E, Segre JA. Stem cells: a new lease on life. Cell 2000;100:143155.
Fuks F. DNA methylation and histone modications: teaming up to silence
genes. Curr Opin Genet Dev 2005;15:490495.
Furukawa Y, Kawano Y, Fukuda J, Matsumoto H, Narahara H. The
production of vascular endothelial growth factor and metalloproteinase
via protease-activated receptor in human endometrial stromal cells.
Fertil Steril 2009;91:535541.
Gallo MA, Kaufman D. Antagonistic and agonistic effects of tamoxifen:
signicance in human cancer. Semin Oncol 1997;24:S1-71S1-80.
Garcia E, Bouchard P, De Brux J, Berdah J, Frydman R, Schaison G,
Milgrom E, Perrot-Applanat M. Use of immunocytochemistry of
progesterone and estrogen receptors for endometrial dating. J Clin
Endocrinol Metab 1988;67:8087.
Gargett CE. Uterine stem cells: what is the evidence? Hum Reprod Update
2007;13:87101.
Gargett CE, Rogers PA. Human endometrial angiogenesis. Reproduction
2001;121:181186.
Gargett CE, Lederman F, Heryanto B, Gambino LS, Rogers PA. Focal
vascular endothelial growth factor correlates with angiogenesis in
human endometrium. Role of intravascular neutrophils. Hum Reprod
2001;16:10651075.
Gargett CE, Chan RW, Schwab KE. Endometrial stem cells. Curr Opin
Obstet Gynecol 2007;19:377383.
Gargett CE, Chan RW, Schwab KE. Hormone and growth factor signaling
in endometrial renewal: role of stem/progenitor cells. Mol Cell
Endocrinol 2008;288:2229.
Gargett CE, Schwab KE, Zillwood RM, Nguyen HP, Wu D. Isolation and
culture of epithelial progenitors and mesenchymal stem cells from
human endometrium. Biol Reprod 2009;80:11361145.
Ghabreau L, Roux JP, Niveleau A, Fontaniere B, Mahe C, Mokni M,
Frappart L. Correlation between the DNA global methylation status
and progesterone receptor expression in normal endometrium,
endometrioid adenocarcinoma and precursors. Virchows Arch 2004;
445:129134.
Girling JE, Rogers PA. Recent advances in endometrial angiogenesis
research. Angiogenesis 2005;8:8999.
Graziani G, Tentori L, Portarena I, Vergati M, Navarra P. Valproic acid
increases the stimulatory effect of estrogens on proliferation of
human endometrial adenocarcinoma cells. Endocrinology 2003;144:
28222828.
306 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

Grozinger CM, Schreiber SL. Regulation of histone deacetylase 4 and 5 and
transcriptional activity by 14-3-3-dependent cellular localization. Proc
Natl Acad Sci USA 2000;97:78357840.
Gunin AG, Kapitova IN, Suslonova NV. Effects of histone deacetylase
inhibitors on estradiol-induced proliferation and hyperplasia formation
in the mouse uterus. J Endocrinol 2005;185:539549.
Guo SW. Epigenetics of endometriosis. Mol Hum Reprod 2009;
15:587607.
Guo JZ, Gorski J. Estrogen effects on modications of chromatin proteins
in the rat uterus. J Steroid Biochem 1989;32:1320.
Ha CH, Jhun BS, Kao HY, Jin ZG. VEGF stimulates HDAC7
phosphorylation and cytoplasmic accumulation modulating matrix
metalloproteinase expression and angiogenesis. Arterioscler Thromb
Vasc Biol 2008;28:17821788.
Heryanto B, Rogers PA. Regulation of endometrial endothelial cell
proliferation by oestrogen and progesterone in the ovariectomized
mouse. Reproduction 2002;123:107113.
Heryanto B, Lipson KE, Rogers PA. Effect of angiogenesis inhibitors on
oestrogen-mediated endometrial endothelial cell proliferation in the
ovariectomized mouse. Reproduction 2003;125:337346.
Heryanto B, Girling JE, Rogers PA. Intravascular neutrophils partially
mediate the endometrial endothelial cell proliferative response to
oestrogen in ovariectomised mice. Reproduction 2004;127:613620.
Hess AP, Hamilton AE, Talbi S, Dosiou C, Nyegaard M, Nayak N,
Genbecev-Krtolica O, Mavrogianis P, Ferrer K, Kruessel J et al.
Decidual stromal cell response to paracrine signals from the
trophoblast: amplication of immune and angiogenic modulators. Biol
Reprod 2007;76:102117.
Hildmann C, Riester D, Schwienhorst A. Histone deacetylasesan
important class of cellular regulators with a variety of functions. Appl
Microbiol Biotechnol 2007;75:487497.
Hodges-Gallagher L, Valentine CD, Bader SE, Kushner PJ. Inhibition of
histone deacetylase enhances the anti-proliferative action of
antiestrogens on breast cancer cells and blocks tamoxifen-induced
proliferation of uterine cells. Breast Cancer Res Treat 2007;105:297309.
Horne AW, Lalani EN, Margara RA, White JO. The effects of sex steroid
hormones and interleukin-1-beta on MUC1 expression in endometrial
epithelial cell lines. Reproduction 2006;131:733742.
Hotz HR, Peters AH. Protein demethylation required for DNA
methylation. Nat Genet 2009;41:1011.
Hu JF, Oruganti H, Vu TH, Hoffman AR. The role of histone acetylation in
the allelic expression of the imprinted human insulin-like growth factor II
gene. Biochem Biophys Res Commun 1998;251:403408.
Hu JF, Pham J, Dey I, Li T, Vu TH, Hoffman AR. Allele-specic histone
acetylation accompanies genomic imprinting of the insulin-like growth
factor II receptor gene. Endocrinology 2000;141:44284435.
Irvine RA, Lin IG, Hsieh CL. DNA methylation has a local effect on
transcription and histone acetylation. Mol Cell Biol 2002;22:66896696.
Ivanova N, Dobrin R, Lu R, Kotenko I, Levorse J, DeCoste C, Schafer X,
Lun Y, Lemischka IR. Dissecting self-renewal in stem cells with RNA
interference. Nature 2006;442:533538.
Jimbo H, Hitomi Y, Yoshikawa H, Yano T, Momoeda M, Sakamoto A,
Tsutsumi O, Taketani Y, Esumi H. Evidence for monoclonal expansion
of epithelial cells in ovarian endometrial cysts. Am J Pathol 1997;
150:11731178.
Jirtle RL, Skinner MK. Environmental epigenomics and disease
susceptibility. Nat Rev Genet 2007;8:253262.
Jorgensen HF, Bird A. MeCP2 and other methyl-CpG binding proteins.
Ment Retard Dev Disabil Res Rev 2002;8:8793.
Kaituu-Lino TJ, Morison NB, Salamonsen LA. Estrogen is not essential for
full endometrial restoration after breakdown: lessons from a mouse
model. Endocrinology 2007;148:51055111.
Kao LC, Tulac S, Lobo S, Imani B, Yang JP, Germeyer A, Osteen K,
Taylor RN, Lessey BA, Giudice LC. Global gene proling in human
endometrium during the window of implantation. Endocrinology 2002;
143:21192138.
Kao LC, Germeyer A, Tulac S, Lobo S, Yang JP, Taylor RN, Osteen K,
Lessey BA, Giudice LC. Expression proling of endometrium from
women with endometriosis reveals candidate genes for disease-based
implantation failure and infertility. Endocrinology 2003;144:28702881.
Kato K, Yoshimoto M, Kato K, Adachi S, Yamayoshi A, Arima T,
Asanoma K, Kyo S, Nakahata T, Wake N. Characterization of
side-population cells in human normal endometrium. Hum Reprod
2007;22:12141223.
Kawano Y, Matsui N, Kamihigashi S, Narahara H, Miyakawa I. Effects of
interferon-gamma on secretion of vascular endothelial growth
factor by endometrial stromal cells. Am J Reprod Immunol 2000;43:
4752.
Kouzarides T. Chromatin modications and their function. Cell 2007;
128:693705.
Krusche CA, Vloet AJ, Classen-Linke I, von Rango U, Beier HM, Alfer J.
Class I histone deacetylase expression in the human cyclic
endometrium and endometrial adenocarcinomas. Hum Reprod 2007;
22:29562966.
Lessey BA. Endometrial receptivity and the window of implantation. Best
Pract Res Clin Obstet Gynaecol 2000;14:775788.
Leu YW, Yan PS, Fan M, Jin VX, Liu JC, Curran EM, Welshons WV,
Wei SH, Davuluri RV, Plass C et al. Loss of estrogen receptor
signaling triggers epigenetic silencing of downstream targets in breast
cancer. Cancer Res 2004;64:81848192.
Leyendecker G, Herbertz M, Kunz G, Mall G. Endometriosis results from
the dislocation of basal endometrium. Hum Reprod 2002;
17:27252736.
Li XF, Gregory J, Ahmed A. Immunolocalisation of vascular endothelial
growth factor in human endometrium. Growth Factors 1994;
11:277282.
Liao X, Siu MK, Chan KY, Wong ES, Ngan HY, Chan QK, Li AS, Khoo US,
Cheung AN. Hypermethylation of RAS effector related genes and DNA
methyltransferase 1 expression in endometrial carcinogenesis. Int J
Cancer 2008;123:296302.
Libby PR. Histone acetylation and hormone action. Early effects of
oestradiol-17beta on histone acetylation in rat uterus. Biochem J 1972;
130:663669.
Lindner HH. Analysis of histones, histone variants, and their
post-translationally modied forms. Electrophoresis 2008;29:25162532.
Lucchini R, Sogo JM. Replication of transcriptionally active chromatin.
Nature 1995;374:276280.
Maass N, Biallek M, Rosel F, Schem C, Ohike N, Zhang M, Jonat W,
Nagasaki K. Hypermethylation and histone deacetylation lead to
silencing of the maspin gene in human breast cancer. Biochem Biophys
Res Commun 2002;297:125128.
Makker A, Singh MM. Endometrial receptivity: clinical assessment in
relation to fertility, infertility, and antifertility. Med Res Rev 2006;
26:699746.
Maruyama T. Stem/progenitor cells and the regeneration potentials in the
human uterus. Reprod Med Biol 2009;18. Epub ahead of print 26 Aug
2009, doi:10.1007/s12522-009-0032-y.
Mattick JS, Amaral PP, Dinger ME, Mercer TR, Mehler MF. RNA regulation
of epigenetic processes. Bioessays 2009;31:5159.
McKinsey TA, Zhang CL, Lu J, Olson EN. Signal-dependent nuclear export
of a histone deacetylase regulates muscle differentiation. Nature 2000;
408:106111.
McLennan CE, Rydell AH. Extent of endometrial shedding during normal
menstruation. Obstet Gynecol 1965;26:605621.
Epigenetic regulation of human endometrium 307

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

Meng X, Ichim TE, Zhong J, Rogers A, Yin Z, Jackson J, Wang H, Ge W,
Bogin V, Chan KW et al. Endometrial regenerative cells: a novel stem
cell population. J Transl Med 2007;5:57.
Mikkelsen TS, Hanna J, Zhang X, Ku M, Wernig M, Schorderet P,
Bernstein BE, Jaenisch R, Lander ES, Meissner A. Dissecting direct
reprogramming through integrative genomic analysis. Nature 2008;
454:4955.
Mirkin S, Arslan M, Churikov D, Corica A, Diaz JI, Williams S, Bocca S,
Oehninger S. In search of candidate genes critically expressed in the
human endometrium during the window of implantation. Hum Reprod
2005;20:21042117.
Moses MA, Langer R. Inhibitors of angiogenesis. Biotechnology (N Y) 1991;
9:630634.
Mottet D, Bellahcene A, Pirotte S, Waltregny D, Deroanne C, Lamour V,
Lidereau R, Castronovo V. Histone deacetylase 7 silencing alters
endothelial cell migration, a key step in angiogenesis. Circ Res 2007;
101:12371246.
Nan X, Ng HH, Johnson CA, Laherty CD, Turner BM, Eisenman RN,
Bird A. Transcriptional repression by the methyl-CpG-binding protein
MeCP2 involves a histone deacetylase complex. Nature 1998;
393:386389.
Nishida M. The Ishikawa cells from birth to the present. Hum Cell 2002;
15:104117.
Nishida M, Kasahara K, Oki A, Satoh T, Arai Y, Kubo T. Establishment of
eighteen clones of Ishikawa cells. Hum Cell 1996;9:109116.
Norwitz ER, Schust DJ, Fisher SJ. Implantation and the survival of early
pregnancy. N Engl J Med 2001;345:14001408.
Noyes RW, Hertig AI, Rock J. Dating the endometrial biopsy. Fertil Steril
1950;1:325.
Ohgane J, Yagi S, Shiota K. Epigenetics: the DNA methylation prole of
tissue-dependent and differentially methylated regions in cells. Placenta
2008;29:S29S35.
Okada H, Nakajima T, Yoshimura T, Yasuda K, Kanzaki H. Microarray
analysis of genes controlled by progesterone in human endometrial
stromal cells in vitro. Gynecol Endocrinol 2003;17:271280.
Okulicz WC, Scarrell R. Estrogen receptor alpha and progesterone
receptor in the rhesus endometrium during the late secretory phase
and menses. Proc Soc Exp Biol Med 1998;218:316321.
Padykula HA. Regeneration in the primate uterus: the role of stem cells.
Ann N Y Acad Sci 1991;622:4756.
Palis J, Batistatou A, Iliopoulou A, Tsanou E, Bakogiannis A,
Dassopoulos G, Charalabopoulos K. Expression of adhesion
molecules during normal pregnancy. Cell Tissue Res 2007;329:111.
Paria BC, Reese J, Das SK, Dey SK. Deciphering the Cross-Talk of
Implantation: Advances and Challenges. Science 2002;296:21852188.
Perry P, Sauer S, Billon N, Richardson WD, Spivakov M, Warnes G,
Livesey FJ, Merkenschlager M, Fisher AG, Azuara V. A dynamic switch
in the replication timing of key regulator genes in embryonic stem
cells upon neural induction. Cell Cycle 2004;3:16451650.
Ponnampalam AP, Weston GC, Trajstman AC, Susil B, Rogers PA.
Molecular classication of human endometrial cycle stages by
transcriptional proling. Mol Hum Reprod 2004;10:879893.
Popovici RM, Kao L-C, Giudice LC. Discovery of New Inducible Genes in
in vitroDecidualized Human Endometrial Stromal Cells Using Microarray
Technology. Endocrinology 2000;141:35103515.
Punyadeera C, Dassen H, Klomp J, Dunselman G, Kamps R, Dijcks F,
Ederveen A, de Goeij A, Groothuis P. Oestrogen-modulated gene
expression in the human endometrium. Cell Mol Life Sci 2005;
62:239250.
Rahnama F, Shaei F, Gluckman PD, Mitchell MD, Lobie PE. Epigenetic
regulation of human trophoblastic cell migration and invasion.
Endocrinology 2006;147:52755283.
Rahnama F, Thompson B, Steiner M, Shaei F, Lobie PE, Mitchell MD.
Epigenetic regulation of E-cadherin controls endometrial receptivity.
Endocrinology 2009;150:14661472.
Rakyan VK, Down TA, Thorne NP, Flicek P, Kulesha E, Graf S,
Tomazou EM, Backdahl L, Johnson N, Herberth M et al. An
integrated resource for genome-wide identication and analysis of
human tissue-specic differentially methylated regions (tDMRs).
Genome Res 2008;18:15181529.
Reik W. Stability and exibility of epigenetic gene regulation in mammalian
development. Nature 2007;447:425432.
Reynolds BA, Rietze RL. Neural stem cells and neurospheres
re-evaluating the relationship. Nat Methods 2005;2:333336.
Riesewijk A, Martin J, van Os R, Horcajadas JA, Polman J, Pellicer A,
Mosselman S, Simon C. Gene expression proling of human
endometrial receptivity on days LH 2 versus LH 7 by microarray
technology. Mol Hum Reprod 2003;9:253264.
Rogers PA, Donoghue JF, Walter LM, Girling JE. Endometrial angiogenesis,
vascular maturation, and lymphangiogenesis. Reprod Sci 2009;
16:147151.
Rossig L, Urbich C, Bruhl T, Dernbach E, Heeschen C, Chavakis E,
Sasaki K, Aicher D, Diehl F, Seeger F et al. Histone deacetylase
activity is essential for the expression of HoxA9 and for endothelial
commitment of progenitor cells. J Exp Med 2005;201:18251835.
Russo VEA, Martienssen RA, Riggs AD. Epigenetic Mechanisms of Gene
Regulation. Cold Spring Harbour Laboratory Press, 1996, New York.
Saito T, Nishimura M, Yamasaki H, Kudo R. Hypermethylation in
promoter region of E-cadherin gene is associated with tumor
dedifferention and myometrial invasion in endometrial carcinoma.
Cancer 2003;97:10021009.
Sakai N, Maruyama T, Sakurai R, Masuda H, Yamamoto Y, Shimizu A,
Kishi I, Asada H, Yamagoe S, Yoshimura Y. Involvement of histone
acetylation in ovarian steroid-induced decidualization of human
endometrial stromal cells. J Biol Chem 2003;278:1667516682.
Schwab KE, Gargett CE. Co-expression of two perivascular cell markers
isolates mesenchymal stem-like cells from human endometrium. Hum
Reprod 2007;22:29032911.
Schrem H, Klempnauer J, Borlak J. Liver-enriched transcription factors in
liver function and development. Part I: the hepatocyte nuclear factor
network and liver-specic gene expression. Pharmacol Rev 2002;
54:129158.
Schwab KE, Chan RW, Gargett CE. Putative stem cell activity of human
endometrial epithelial and stromal cells during the menstrual cycle.
Fertil Steril 2005;84:11241130.
Selker EU. Trichostatin A causes selective loss of DNA methylation in
Neurospora. Proc Natl Acad Sci USA 1998;95:94309435.
Seppala M, Taylor RN, Koistinen H, Koistinen R, Milgrom E. Glycodelin: a
major lipocalin protein of the reproductive axis with diverse actions in
cell recognition and differentiation. Endocr Rev 2002;23:401430.
Shahbazian MD, Grunstein M. Functions of site-specic histone acetylation
and deacetylation. Annu Rev Biochem 2007;76:75100.
Shukla V, Vaissiere T, Herceg Z. Histone acetylation and chromatin
signature in stem cell identity and cancer. Mutat Res 2008;637:
115.
Simon C, Moreno C, Remohi J, Pellicer A. Cytokines and embryo
implantation. J Reprod Immunol 1998;39:117131.
Singal R, van Wert J, Bashambu M. Cytosine methylation represses
glutathione S-transferase P1 (GSTP1) gene expression in human
prostate cancer cells. Cancer Res 2001;61:48204826.
Snykers S, Henkens T, De Rop E, Vinken M, Fraczek J, De Kock J, De
Prins E, Geerts A, Rogiers V, Vanhaecke T. Role of epigenetics in
liver-specic gene transcription, hepatocyte differentiation and stem
cell reprogrammation. J Hepatol 2009;51:187211.
308 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

Sobel RE, Cook RG, Perry CA, Annunziato AT, Allis CD. Conservation of
deposition-related acetylation sites in newly synthesized histones H3
and H4. Proc Natl Acad Sci USA 1995;92:12371241.
Sogo JM, Stahl H, Koller T, Knippers R. Structure of replicating simian virus
40 minichromosomes. The replication fork, core histone segregation
and terminal structures. J Mol Biol 1986;189:189204.
Song MR, Ghosh A. FGF2-induced chromatin remodeling regulates
CNTF-mediated gene expression and astrocyte differentiation. Nat
Neurosci 2004;7:229235.
Song H, Lim H, Paria BC, Matsumoto H, Swift LL, Morrow J, Bonventre JV,
Dey SK. Cytosolic phospholipase A2alpha is crucial [correction of
A2alpha deciency is crucial] for on-time embryo implantation
that directs subsequent development. Development 2002;129:
28792889.
Song F, Smith JF, Kimura MT, Morrow AD, Matsuyama T, Nagase H,
Held WA. Association of tissue-specic differentially methylated
regions (TDMs) with differential gene expression. Proc Natl Acad Sci
USA 2005;102:33363341.
Starzinski-Powitz A, Zeitvogel A, Schreiner A, Baumann R. In search of
pathogenic mechanisms in endometriosis: the challenge for molecular
cell biology. Curr Mol Med 2001;1:655664.
Szyf M, Eliasson L, Mann V, Klein G, Razin A. Cellular and viral DNA
hypomethylation associated with induction of Epstein-Barr virus lytic
cycle. Proc Natl Acad Sci USA 1985;82:80908094.
Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K,
Yamanaka S. Induction of pluripotent stem cells from adult human
broblasts by dened factors. Cell 2007;131:861872.
Takai N, Desmond JC, Kumagai T, Gui D, Said JW, Whittaker S,
Miyakawa I, Koefer HP. Histone deacetylase inhibitors have a
profound antigrowth activity in endometrial cancer cells. Clin Cancer
Res 2004;10:11411149.
Talbi S, Hamilton AE, Vo KC, Tulac S, Overgaard MT, Dosiou C, Le
Shay N, Nezhat CN, Kempson R, Lessey BA et al. Molecular
phenotyping of human endometrium distinguishes menstrual cycle
phases and underlying biological processes in normo-ovulatory
women. Endocrinology 2006;147:10971121.
Tanaka T, Nakajima S, Umesaki N. Cellular heterogeneity in long-term
surviving cells isolated from eutopic endometrial, ovarian
endometrioma and adenomyosis tissues. Oncol Rep 2003;10:11551160.
Tierney EP, Tulac S, Huang S-TJ, Giudice LC. Activation of the protein
kinase A pathway in human endometrial stromal cells reveals
sequential categorical gene regulation. Physiol Genomics 2003;16:4766.
Trenkmann M, Brock M, Ospelt C, Gay S. Epigenetics in rheumatoid
arthritis. Clin Rev Allergy Immunol 2009. Epub ahead of print 27 Aug
2009, doi:10.1007/s12016-009-8166-6.
Tsuji S, Yoshimoto M, Takahashi K, Noda Y, Nakahata T, Heike T. Side
population cells contribute to the genesis of human endometrium.
Fertil Steril 2008;90:15281537.
Turunen MP, Aavik E, Yla-Herttuala S. Epigenetics and atherosclerosis.
Biochim Biophys Acta 2009;1790:886891.
Uchida H, Maruyama T, Arase T, Ono M, Nagashima T, Masuda H,
Asada H, Yoshimura Y. Histone acetylation in reproductive organs:
Signicance of histone deacetylase inhibitors in gene transcription.
Reprod Med Biol 2005a;4:115122.
Uchida H, Maruyama T, Nagashima T, Asada H, Yoshimura Y. Histone
deacetylase inhibitors induce differentiation of human endometrial
adenocarcinoma cells through up-regulation of glycodelin.
Endocrinology 2005b;146:53655373.
Uchida H, Maruyama T, Ohta K, Ono M, Arase T, Kagami M, Oda H,
Kajitani T, Asada H, Yoshimura Y. Histone deacetylase
inhibitor-induced glycodelin enhances the initial step of implantation.
Hum Reprod 2007a;22:26152622.
Uchida H, Maruyama T, Ohta K, Ono M, Arase T, Kagami M, Oda H,
Kajitani T, Asada H, Yoshimura Y. Histone deacetylase
inhibitor-induced glycodelin enhances the initial step of implantation.
Hum Reprod 2007b;22:26152622.
Vega RB, Harrison BC, Meadows E, Roberts CR, Papst PJ, Olson EN,
McKinsey TA. Protein kinases C and D mediate agonist-dependent
cardiac hypertrophy through nuclear export of histone deacetylase 5.
Mol Cell Biol 2004;24:83748385.
Verdin E, Dequiedt F, Kasler HG. Class II histone deacetylases: versatile
regulators. Trends Genet 2003;19:286293.
Verreault A. De novo nucleosome assembly: new pieces in an old puzzle.
Genes Dev 2000;14:14301438.
Walter LM, Rogers PA, Girling JE. The role of progesterone in endometrial
angiogenesis in pregnant and ovariectomised mice. Reproduction 2005;
129:765777.
Wang H, Dey SK. Roadmap to embryo implantation: clues from mouse
models. Nat Rev Genet 2006;7:185199.
Wang S, Li X, Parra M, Verdin E, Bassel-Duby R, Olson EN. Control
of endothelial cell proliferation and migration by VEGF signaling
to histone deacetylase 7. Proc Natl Acad Sci USA 2008;105:
77387743.
Wang J, Hevi S, Kurash JK, Lei H, Gay F, Bajko J, Su H, Sun W, Chang H,
Xu G et al. The lysine demethylase LSD1 (KDM1) is required
for maintenance of global DNA methylation. Nat Genet 2009;41:
125129.
Widschwendter M, Apostolidou S, Jones AA, Fourkala EO, Arora R,
Pearce CL, Frasco MA, Ayhan A, Zikan M, Cibula D et al. HOXA
methylation in normal endometrium from premenopausal women is
associated with the presence of ovarian cancer: a proof of principle
study. Int J Cancer 2009;125:22142218.
Wilcox AJ, Baird DD, Weinberg CR. Time of implantation of the
conceptus and loss of pregnancy. N Engl J Med 1999;340:17961799.
Worcel A, Han S, Wong ML. Assembly of newly replicated chromatin. Cell
1978;15:969977.
Wu Y, Basir Z, Kajdacsy-Balla A, Strawn E, Macias V, Montgomery K,
Guo SW. Resolution of clonal origins for endometriotic lesions using
laser capture microdissection and the human androgen receptor
(HUMARA) assay. Fertil Steril 2003;79:710717.
Wu H, Chen Y, Liang J, Shi B, Wu G, Zhang Y, Wang D, Li R, Yi X,
Zhang H et al. Hypomethylation-linked activation of PAX2 mediates
tamoxifen-stimulated endometrial carcinogenesis. Nature 2005a;
438:981987.
Wu Y, Halverson G, Basir Z, Strawn E, Yan P, Guo S-W. Aberrant
methylation at HOXA10 may be responsible for its aberrant
expression in the endometrium of patients with endometriosis. Am J
Obstet Gynecol 2005b;193:371380.
Yamada N, Nishida Y, Tsutsumida H, Hamada T, Goto M, Higashi M,
Nomoto M, Yonezawa S. MUC1 expression is regulated by DNA
methylation and histone H3 lysine 9 modication in cancer cells.
Cancer Res 2008;68:27082716.
Yamagata Y, Asada H, Tamura I, Lee L, Maekawa R, Taniguchi K,
Taketani T, Matsuoka A, Tamura H, Sugino N. DNA
methyltransferase expression in the human endometrium:
down-regulation by progesterone and estrogen. Hum Reprod 2009;
24:11261132.
Yang XJ, Seto E. HATs and HDACs: from structure, function and
regulation to novel strategies for therapy and prevention. Oncogene
2007;26:53105318.
Ye X, Hama K, Contos JJ, Anliker B, Inoue A, Skinner MK, Suzuki H,
Amano T, Kennedy G, Arai H et al. LPA3-mediated lysophosphatidic
acid signalling in embryo implantation and spacing. Nature 2005;
435:104108.
Epigenetic regulation of human endometrium 309

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

Yeo S, Jeong S, Kim J, Han JS, Han YM, Kang YK. Characterization of DNA
methylation change in stem cell marker genes during differentiation of
human embryonic stem cells. Biochem Biophys Res Commun 2007;
359:536542.
Yoshiura K, Kanai Y, Ochiai A, Shimoyama Y, Sugimura T, Hirohashi S.
Silencing of the E-cadherin invasion-suppressor gene by CpG methylation
in human carcinomas. Proc Natl Acad Sci USA 1995;92:74167419.
Yoshida H, Broaddus R, Cheng W, Xie S, Naora H. Deregulation
of the HOXA10 homeobox gene in endometrial carcinoma:
role in epithelial-mesenchymal transition. Cancer Res 2006;66:
889897.
Zardo G, Cimino G, Nervi C. Epigenetic plasticity of chromatin in
embryonic and hematopoietic stem/progenitor cells: therapeutic
potential of cell reprogramming. Leukemia 2008;22:15031518.
Zaret KS, Watts J, Xu J, Wandzioch E, Smale ST, Sekiya T. Pioneer factors,
genetic competence, and inductive signaling: programming liver and
pancreas progenitors from the endoderm. Cold Spring Harb Symp
Quant Biol 2008;73:119126.
310 Munro et al.

b
y

g
u
e
s
t

o
n

A
u
g
u
s
t

3
1
,

2
0
1
4
h
t
t
p
:
/
/
m
o
l
e
h
r
.
o
x
f
o
r
d
j
o
u
r
n
a
l
s
.
o
r
g
/
D
o
w
n
l
o
a
d
e
d

f
r
o
m

You might also like