ArticlePDF Available

Epigenetic Hierarchy Governing Nestin Expression

Authors:

Abstract

Nestin is an intermediate filament protein expressed specifically in neural stem cells and progenitor cells of the central nervous system. DNA demethylation and histone modifications are two types of epigenetic modifications working in a coordinate or synergistic manner to regulate the expression of various genes. This study investigated and elucidated the epigenetic regulation of Nestin gene expression during embryonic differentiation along the neural cell lineage. Nestin exhibits differential DNA methylation and histone acetylation patterns in Nestin-expressing and nonexpressing cells. In P19 embryonic carcinoma cells, activation of Nestin expression is mediated by both trichostatin A and 5-aza-2′-deoxycytidine treatment, concomitant with histone acetylation, but not with DNA demethylation. Nestin transcription is also mediated by treatment with retinoic acid, again in the absence of DNA demethylation. Thus, histone acetylation is sufficient to mediate the activation of Nestin transcription. This study proposed that the regulation of Nestin gene expression can be used as a model to study the epigenetic regulation of gene expression mediated by histone acetylation, but not by DNA demethylation. Disclosure of potential conflicts of interest is found at the end of this article.
Author contributions: D.W.H.: Conception and design, data collection and/or assembly, data analysis and interpretation, manuscript writing; J.T.D.:
Data analysis and interpretation, data collection and/or assembly; M.J.A.-B.: Data analysis and interpretation; S.H.L.: Data analysis and
interpretation; A.M.: Data analysis and interpretation, provision of study materials; H.T.L.: Data analysis and interpretation; R.J.: Data analysis and
interpretation, provision of study materials; H.R.S.: Conception and design, data analysis and interpretation, financial support, manuscript writing.
4Corresponding author: Hans R. Schöler, Ph. D., Email: h.schoeler@mpi-muenster.mpg.de, Tel: 49-251-70365-300, Fax: 49-251-70365-399.
Received August 14, 2008; accepted for publication February 10, 2009; first published online in Stem Cells Express February 19, 2009. ©AlphaMed
Press 1066-5099/2009/$30.00/0 doi: 10.1002/stem.43
STEM CELLS®
E
MBRYONIC
S
TEM
C
ELLS
/I
NDUCED
P
LURIPOTENT
S
TEM
C
ELLS
Epigenetic Hierarchy Governing Nestin Expression
Dong Wook Han1,2, Jeong Tae Do1, Marcos J. Araúzo-Bravo1, Sung Ho Lee2, Alexander
Meissner3, Hoon Taek Lee2, Rudolf Jaenisch3, and Hans R. Schöler1,4
1Department of Cell and Developmental Biology, Max Planck Institute for Molecular Biomedicine, Röntgenstrasse 20,
48149 Münster, Germany; 2Department of Bioscience and Biotechnology, Bio-Organ Research Center, Konkuk
University, 1 Hwayang-dong, Gwangjin-Gu, Seoul 143-701, South Korea; 3Whitehead Institute and Department of
Biology, Massachusetts Institute of Technology, 9 Cambridge Center, Cambridge, MA 02142, USA
Key words. DNA methylation/epigenetic regulation/histone acetylation/Nestin
ABSTRACT
Nestin is an intermediate filament protein expressed
specifically in neural stem cells and progenitor cells
of the central nervous system. DNA demethylation
and histone modifications are two types of epigenetic
modifications working in a coordinate or synergistic
manner to regulate the expression of various genes.
This study investigates and elucidates the epigenetic
regulation of Nestin gene expression during
embryonic differentiation along the neural cell
lineage. Nestin exhibits differential DNA methylation
and histone acetylation patterns in Nestin-expressing
and non-expressing cells. In P19 embryonic
carcinoma (EC) cells, activation of Nestin expression
is mediated by both trichostatin A (TSA) and 5-aza-
2’-deoxycytidine (5-aza-dC) treatment, concomitant
with histone acetylation, but not with DNA
demethylation. Nestin transcription is also mediated
by treatment with retinoic acid (RA), again in the
absence of DNA demethylation. Thus, histone
acetylation is sufficient to mediate activation of
Nestin transcription. This study proposes that
regulation of Nestin gene expression can be used as a
model to study the epigenetic regulation of gene
expression mediated by histone acetylation, but not
by DNA demethylation.
INTRODUCTION
Nestin is an intermediate filament protein
expressed specifically in neural stem cells
(NSCs) and neural progenitor cells of the
central nervous system (CNS) ranging from
embryonic to adult stages [1] and thus has been
widely used as a marker gene to identify and
isolate NSCs [2-4]. As NSCs differentiate into
neuronal and glial cells during CNS
development, Nestin expression diminishes and
the protein is gradually replaced by
neurofilaments in neurons and glial fibrillary
acidic proteins (GFAP) in glial cells [5].
However, Nestin is also expressed in myogenic
precursor cells [6] and in developing heart [7],
differentiating testis [8], and developing tooth
bud [9] tissues during embryogenesis.
The Nestin gene is well-conserved among
species containing four exons and three introns
[2, 4, 10]. Two different tissue-specific
enhancers located in the first and second
introns of the Nestin gene drive Nestin
expression in somatic muscle progenitor cells
and NSCs, respectively. The CNS-specific
second intron enhancer is considered to be an
important element in the regulation of Nestin
expression [11]. In addition, an enhancer
element located in the 3’ half of the second
intron enhancer called the 3’enhancer is also
known to be involved in Nestin regulation in
the CNS [12-14]. A recent study has shown
that activation of Nestin expression is
dependent upon the binding of transcription
factors, such as SOX and POU factors, to the
Nestin 3’ enhancer [15]. However, although
factors involved in the transcriptional
regulation of Nestin expression have been
identified, evidence for their participation in an
epigenetic mechanism has remained elusive.
Epigenetic modifications such as DNA
methylation and chromatin remodeling are
important components in the transcriptional
regulation of many genes. DNA demethylation
and histone modifications have been shown to
play a synergic role in the activation of gene
transcription [16-19]. Despite the intense
characterization and the widespread use of
Nestin as a marker for NSCs, relatively little is
known about the epigenetic mechanism
governing Nestin transcription. The current
study aims to determine whether DNA
methylation and histone modifications are
involved in the regulation of Nestin regulation.
To this end, the study investigated 1) the effect
of inhibiting DNA methylation and histone
deacetylation (i.e. of using histone deacetylase
[HDAC]) on Nestin transcription; 2) the
methylation status of the Nestin regulatory
elements in both Nestin-expressing and non-
expressing cells; 3) the effect of inhibiting
DNA methylation and histone deacetylation on
DNA methylation patterns of Nestin regulatory
elements; 4) the transcriptional status of Nestin
in DNA methyltransferases knock-out (Dnmts
KO) embryonic stem (ES) cells; 5) the histone
acetylation status of Nestin regulatory elements
in both Nestin-expressing and non-expressing
cells; and 6) the epigenetic changes in Nestin
mediated by inhibition of DNA methylation
and histone deacetylation and by
differentiation induced by retinoic acid (RA).
MATERIALS AND METHODS
Mice
All mouse strains were either bred and housed
at the mouse facility of the Max Planck
Institute (MPI) or were bought from Halan or
Jackson laboratories. Animal handling was in
accordance with the MPI animal protection
guidelines and the German animal protection
laws.
Cell culture and treatment with reagents
P19 EC cells were grown on gelatin-coated
(0.1% in PBS) dishes in standard EC cell
media: high-glucose DMEM (Gibco BRL)
containing 15% fetal calf serum (FCS; Gibco
BRL), 1X penicillin/streptomycin/glutamine,
and 1X nonessential amino acids (Gibco BRL).
For the derivation of NSCs, brain tissue was
collected from 12.5- to 16.5-dpc
OG2/ROSA26 heterozygous female mice
according to a previous protocol [31]. Isolated
NSCs were grown in NS-A medium
(Euroclone) supplemented with N2
supplement, 10 ng/ml EGF, 10 ng/ml bFGF
(both from Invitrogen), 50 µg/ml BSA
(Fraction V; Gibco BRL), and 1X
penicillin/streptomycin/glutamine and 1X
nonessential amino acids (Gibco BRL).
P19 EC cells were first pre-cultured for 48 hrs
in standard EC cell media and then cultured
with either 5-aza-dC (5 ȝM and 10 ȝM) for 8,
16, 24, 48, and 72 hrs or with TSA (200 nM)
for 4, 6, 8, and 16 hrs. Pre-cultured P19 EC
cells were also treated with retinoic acid (RA;
5ȝM) for 12, 24, and 48 hrs.
RNA extraction, cDNA synthesis, and RT-
PCR
Total RNA was isolated from each sample
using the ToTALLY RNA kit (Ambion).
Reverse transcription was performed with 1 ȝg
of DNase-treated total RNA using the
RETROscript kit (Ambion) following the
manufacturer’s protocol. cDNA was amplified
by PCR (35 cycles) using SuperTaq
polymerase (Ambion) with the primers
described in a previous publication by this
laboratory [32].
Bisulfite sequencing analysis
To assess the DNA methylation status of
Nestin regulatory elements, genomic DNA was
treated with sodium bisulfite to convert all
unmethylated cytosine residues into uracil ones
using the One Day MSP kit (In2Gen, Korea)
according to a previous protocol [33]. In brief,
purified genomic DNA (0.5-1 ȝg) was
2
Epigenetic hierarchy governing
Nestin
expression
denatured with 3N NaOH (Sigma-Aldrich,
USA) at 37ºC for 10 min. Sodium bisulfite (5
M) was then added and the mixture was
incubated at 50ºC for 16 hrs in the dark.
Following desulfonation, neutralization, and
desalting, the modified DNA was dissolved in
20 ȝl of distilled water. Bisulfite PCR
amplification was then carried out using
aliquots of 1 ȝl of modified DNA per PCR
reaction.
PCR amplification and sequencing of the
bisulfite-treated genomic DNA
In this study, the methylation status of the
Nestin promoter and the 5’ and 3’enhancers
was compared in Nestin-expressing and non-
expressing cells. Following bisulfite treatment,
all genomic regions selected were amplified
using a nested primer approach. PCR
amplifications were performed using SuperTaq
polymerase (Ambion) in a total volume of 25
ȝl. All PCR amplifications consisted of a total
of 50 cycles of denaturation at 94ºC for 30s,
annealing at the appropriate temperature for
each target region for 30s, extension at 72ºC
for 30s with a 1st denaturation at 94ºC for 5
min, and a final extension at 72ºC for 10 min.
The primer sequences and annealing
temperatures used were as follows: Nestin
promoter 1st sense 5’-
GTTTTGGGTTGTTTGGTTGTATT-3’,
Nestin promoter 1st antisense 5’-
ATAACCCTTAAAACTTTTAAAAAAAA-3’
(298 bp, 45ºC), Nestin promoter 2nd sense 5’-
TTTGTTGAGTTGGGATGATGTAG-3’,
Nestin promoter 2nd antisense 5’-
TAACCCTTAAAACTTTTAAAAAAAA-3’
(202 bp, 50ºC), 5’enhancer 1st sense 5’-
TAAAGAGGTTGTTTGGTTTGGTAGT-3’,
5’enhancer 1st antisense 5’-
CTATTCCACTCAACCTTCCTAAAAC-3’
(394 bp, 45ºC), 5’enhancer 2nd sense 5’-
TAGTTTTTAGGGAGGAGATTAGAGG-3’,
5’enhancer 2nd antisense 5’-
CTCTTACCCCAAACACAACTAAAAC-3’
(188 bp, 58ºC), 3’enhancer 1st sense 5’-
TTTGTTATTAGTTTTGGGGGTTTAG-3’,
3’enhancer 1st antisense 5’-
CATATCCTACCACTACAAAATCACTCTT
-3’ (387 bp, 45ºC), 3’enhancer 2nd sense 5’-
TTTTTAGATGGTAGTGTGGATAAAAG-
3’, and 3’enhancer 2nd antisense 5’-
CCCATATTTAAAACTCAAAAAAACC-3’
(203 bp, 58ºC). For each primer set, 3 ȝl of
product from the first round of PCR was used
in the second round of PCR. The amplified
products were verified by electrophoresis on
1% agarose gel. The PCR products were
subcloned using the PCR 2.1-TOPO vector
(Invitrogen) according to the manufacturer’s
protocol. Reconstructed plasmids were purified
using the QIAprep Spin Miniprep kit (Qiagen)
and individual clones were sequenced (GATC-
biotech, Germany). Clones were only accepted
if there was at least 90% cytosine conversion
and all possible clonalities were excluded
based on the criteria from BiQ Analyzer
software (Max Planck Society, Germany). For
each fusion hybrid, results were confirmed by
performing at least 10 replicates per selected
genomic region and at least three separate
bisulfite treatments.
ChIP analysis
A chromatin immunoprecipitation assay was
performed using anti-acetylated histone H3
(06-599) and H4 (06-598) antibodies (Upstate
Biotechnology). In brief, cells were rinsed with
phosphate-buffered saline (PBS) and treated
with 1% formaldehyde at 37ºC for 10 min to
cross-link histones and proteins. Each sample
was first sonicated to obtain DNA fragments
that were 200- to 1,000-bp long and then
incubated with or without the antibodies
overnight at 4ºC to detect specific or
nonspecific binding, respectively. Heating at
65ºC for 4 hrs reversed covalent histone-DNA
bonds. DNA was recovered by
phenol/chloroform extraction and amplified by
PCR (35 cycles) with the following primer
sets: Hprt sense 5’-
TCAGGCCCACCTAGTCAGAT-3’, Hprt
antisense 5’-
CGGAAAGCAGTGAGGTAAGC-3’ (198 bp,
62ºC), Nestin promoter sense 5’-
TCCCTTTCTCCCCCTTAAAA-3’, Nestin
promoter antisense 5’-
TGAGCTCCCACATCTGAAAA-3’ (133 bp,
52ºC), 5’enhancer-1 sense 5’-
AGGGAGGAAACCAGAGGGTA-3’,
5’enhancer-1 antisense 5’-
AAGGAAGGCAGAAGGCTAGG-3’ (207 bp,
62ºC), 5’enhancer-2 sense 5’-
AACCACTGAGCCATCTGTCC-3’,
3
5’enhancer-2 antisense 5’-
TGCCTGCCTTTCAAAGAACT-3’ (270 bp,
62ºC), 3’enhancer-1 sense 5’-
GCCCCAGTCAGTCTTCTGAG-3’,
3’enhancer-1 antisense 5’-
TGGGAATTCTCAGGCTGTTC-3’ (181 bp,
62ºC), 3’enhancer-2 sense 5’-
CATGGGAACAGCCTGAGAAT-3’,
3’enhancer-2 antisense 5’-
AGGGCCAAGTGAATTGCTAA-3’ (205 bp,
58ºC), Oct4 sense 5’-
CCTCCGTCTGGAAGACACAGGCAGATA
GCG-3’, and Oct4 antisense 5’-
CGAAGTCTGAAGCAGGTGTCCAGCCAT
GG-3’ (221 bp, 55ºC). All PCR reactions were
performed in triplicate with at least two
independently prepared samples. Each PCR
product was quantitated using an ethidium
bromide–stained gel image and fold
enrichment was calculated using the
Alphaimager HP software (Hewlett Packard).
RESULTS
Activation of Nestin expression mediated by
epigenetic modifications
Nestin is expressed by cells of neural origin—
such as neural stem cells and neural progenitor
cells—but not by non-neural embryonic
carcinoma (EC) cells as determined by RT-
PCR (Fig. 1), consistent with previous results
[20]. Since P19 EC cells are frequently used as
an in-vitro model system to study
developmental milestone events, such as the
differentiation of neurons [21], P19 EC cells
were treated with reagents that induce
epigenetic modifications to determine whether
Nestin undergoes epigenetic regulation. Cells
were treated with either 5-aza-2’-deoxycytidine
(5-aza-dC), an inhibitor of DNA methylation,
for 8, 16, 24, 48, and 72 hrs or trichostatin A
(TSA), an inhibitor of histone deacetylation,
for 4, 6, 8, and 16 hrs. Nestin mRNA was first
detected after 4 hrs of TSA treatment in Nestin
non-expressing P19 EC cells; i.e. time-
dependent expression (Fig. 1A, Fig. S2A). F9
EC cells, another type of EC cell, also
exhibited the same expression pattern (Fig.
1A). Activation of Nestin expression was also
observed in P19 EC cells treated with different
concentrations of 5-aza-dC for 4, 6, 8, 16, 24,
48, and 72 hrs. Nestin mRNA was first
detected after 24 hrs of 5-aza-dC treatment, in
a time- and dose-dependent fashion (Fig. 1B,
Fig. S2B). These data were also confirmed
with F9 EC cells (Fig. 1B). The time-
dependent expression of Nestin mRNA
following inhibition of DNA methylation (Fig.
1B) and histone deacetylation (Fig. 1A) attests
to the multitude of regulatory mechanisms
governing Nestin expression. However, since
both TSA and 5-aza-dC could mediate
activation of Nestin expression in Nestin non-
expressing P19 EC cells, it was postulated that
Nestin expression may be epigenetically
regulated at both the DNA and histone levels.
DNA methylation status of promoter and
second intron enhancer regions of Nestin
DNA methylation is thought to play an
important regulatory role in Nestin expression,
suggesting that Nestin regulatory elements are
differentially methylated in Nestin-expressing
and non-expressing cells. To confirm this
hypothesis, the DNA methylation status of
Nestin-expressing NSCs was compared with
that of non-expressing P19 EC cells by
bisulfite sequencing analysis. The results
showed that the Nestin promoter region (-161
to + 183 bp), which exhibits the highest
activity among promoter regions [22], was
completely unmethylated in both NSCs and
P19 EC cells (Fig. 2A). Previous studies had
determined the second intron enhancer (i.e.
CNS enhancer) located in the second intron of
Nestin to be a critical regulatory element of
Nestin expression, as the 5’ and 3’enhancer
elements of the second intron enhancer play
different roles in the regulation of Nestin
transcription [11]. In particular, the 3’enhancer
element, which harbors SOX and POU binding
sites, has been shown to be essential in Nestin
regulation [15]. Thus, the DNA methylation
status of both 5’ and 3’enhancers in Nestin-
expressing NSCs was compared with that of
non-expressing P19 EC cells. The 3’enhancer,
along with the promoter region, was found to
be completely unmethylated in both NSCs and
P19 EC cells, while the 5’enhancer was
differentially methylated in NSCs and P19 EC
cells (Fig. 2A). Nestin-expressing NSCs were
found to be nearly completely unmethylated
(1.5% methylation), but P19 EC cells, which
do not express Nestin, were considerably
4
Epigenetic hierarchy governing
Nestin
expression
methylated (68.6% methylation). This result
was also confirmed with other embryonic
cells—F9 EC cells (methylation rate: 58.6%)
and ES cells (methylation rate: 73.3%). Taken
together, these data suggest that DNA
methylationídependent gene silencing
regulates Nestin expression.
DNA demethylation is not sufficient to
mediate activation of Nestin transcription in
EC cells
To delve deeper into the series of events
involved in Nestin regulation, first the DNA
methylation status of Nestin regulatory
elements was assessed in 5-aza-dCítreated P19
EC cells. Onset of demethylation of the Nestin
5’enhancer element in P19 EC cells was
observed only after 16 hrs of 5-aza-dC
treatment, suggesting that 5-aza-dC treatment
caused cell cycle–dependent passive
demethylation of the Nestin 5’enhancer (Fig.
2B). This demethylation pattern was similarly
observed for the Nestin 5’enhancer in 5-aza-
dCítreated F9 EC cells (Fig. 2B). Next, the
DNA methylation status of the Nestin
5’enhancer was assessed in TSA-treated P19
EC cells. TSA can induce not only gene-
specific, but also global DNA demethylation
[23, 24]; TSA-induced demethylation has been
attributed to Dnmt1 and Dnmt3b down-
regulation [25, 26]. Onset of Nestin 5’enhancer
demethylation was observed in P19 EC cells
after 16 hrs of TSA treatment, with a
methylation rate of only 30%, which is about
2-fold lower than that in untreated P19 EC
cells and even lower than that in P19 EC cells
after 24 hrs of 5-aza-dC treatment (Fig. 2B, C).
The same pattern was observed in F9 EC cells
(Fig. 2B, C). But since the onset of Nestin
expression in P19 EC cells occurs after 4 hrs of
TSA treatment, demethylation of Nestin
regulatory element may not be essential for
activation of Nestin transcription in P19 EC
cells. Moreover, since Nestin expression was
first detected after 24 hrs of 5-aza-dC treatment
and 8 hours after the onset of 5’enhancer
demethylation, DNA demethylation does not
appear to be sufficient to mediate activation of
Nestin transcription. However, we cannot
exclude that at a later stage activator
complexes interact with the demethylated
enhancer and result in delayed activation of
Nestin expression in 5-aza-dC treated cells. In
such a case the recruitment of an activator
complex would take several hours, and the
onset of Nestin expression would occur at
about the same time in TSA-treated cells as in
5-aza-dC–treated cells. However the onset of
Nestin expression was observed after the cells
had been treated with TSA for only 4 hours.
Therefore, DNA demethylation is not sufficient
to cause Nestin expression.
Transcription status of Nestin in Dnmts-
deficient ES cells
To further discern any potential involvement of
DNA methylation in the regulation of Nestin
transcription—as suggested by the differential
methylation pattern of the 5’enhancer in
Nestin-expressing and non-expressing cells—
the DNA methylation pattern was assessed in
DNA methyltransferase knock-out ES cell
lines (Dnmt KO ES cells). Genome-wide
demethylation has been observed in the Dnmts
KO ES cells [27]. If DNA methylation were to
play a role in Nestin expression, the Nestin
gene would be expected to be transcriptionally
active in Dnmts KO ES cells, which exhibit
Nestin enhancer demethylation. The Nestin
5’enhancer was assessed in wild-type ES cells
and in three different Dnmt KO ES cells lines:
Dnmt1 KO ES (Dnmt1-deficient), Dnmt 3a, 3b
KO ES (Dnmt3a- and 3b-deficient), and Dnmt
total knock-out (TKO) ES cells [28].
Hypermethylation was found in the wild-type
ES cells and hypomethylation in all Dnmt KO
ES cells—3.1% methylation rate in Dnmt1 KO
ES, 0.0% in Dnmt3 KO ES, and 0.0% in Dnmt
TKO ES cells (Fig. 3A). Although the degree
of Nestin enhancer hypomethylation in all
Dnmts KO ES cells was similar to that of
Nestin-expressing NSCs, Nestin expression
could not be detected in untreated Dnmts KO
ES cells, but only in TSA-treated cells (Fig.
3B). This demonstrates that DNA
demethylation of the Nestin enhancer element
is not sufficient to mediate activation of Nestin
transcription. It is possible that differentiation
of Dnmt KO ES lines into NSCs is
accompanied by enhanced Nestin expression
due to their unmethylated promoter. To test
this we have attempted to derive NSCs from
each Dnmt KO ES cell line based on the
published protocol [31]. However, although
5
differentiation into early precursors of NSCs
was possible, we were unable to generate NSC
lines from any of the Dnmt KO ES cell lines.
These early precursors of NSCs all stained
positive for Nestin and Sox2 as determined by
immunocytochemistry, similar to derivatives of
control ES cells (data not shown). However,
we were unable to obtain evidence for the
earlier activation of Nestin. Thus, enhanced
induction of Nestin activation could not be
observed in differentiated Dnmt KO ES cells,
indicating that DNA demethylation alone does
not support the activation of Nestin expression.
Chromatin structure of Nestin promoter
and enhancer elements
Since both TSA and 5-aza-dC treatment could
mediate the activation of Nestin transcription in
EC cells, the chromatin status of Nestin was
assessed in NSCs and P19 EC cells. Histone
acetylation of the Nestin regulatory elements in
P19 EC cells and NSCs was examined by
chromatin immunoprecipitation (ChIP) assay
using antibodies that recognize acetylated
versions of histones H3 and H4. The
acetylation level of histone H3 on the Nestin
promoter element was not found to differ
between NSCs and P19 EC cells. The same
region was completely unmethylated in both
NSCs and P19 EC cells (Fig. 2A), while
histone H4 on the promoter element was
slightly more acetylated in NSCs compared
with P19 EC cells (Fig. 4, Fig. S4B). However,
the acetylation level of histones H3 and H4 on
both 5’ and 3’enhancers was significantly
higher in NSCs compared with P19 EC cells
(Fig. 4, Fig. S4A, B) and F9 EC cells (Fig. S3).
Moreover, there was a nearly 3-fold difference
in histone H4 acetylation on the Nestin second
intron enhancer elements between NSCs and
P19 EC cells. On the other hand, there was a
2.4- to 3.3-fold higher acetylation of Oct4, a
marker for pluripotent cells, in P19 EC cells
compared with NSCs (Fig. 4, Fig. S4A, B).
Taken together, these data demonstrate the
epigenetic regulation of histones on the Nestin
promoter and enhancer elements.
Histone acetylation is sufficient to mediate
activation of Nestin transcription
Since activation of Nestin transcription can
occur without DNA demethylation after 4 hrs
of TSA treatment (Fig. 1A, Fig. 2C), histone
acetylation—which can be induced by TSA—
may be sufficient to mediate activation of
Nestin transcription. To test this possibility, the
acetylation of histones H3 and H4 on the
Nestin enhancer after TSA treatment was
assessed in a time-dependent fashion (4, 6, 8,
and 16 hrs). Strikingly, a dramatic increase in
the acetylation levels of both histones H3 and
H4 was found after 4 hrs of TSA treatment,
comparable with levels in NSCs, thereby
suggesting that TSA treatment induces rapid
histone acetylation on the Nestin enhancer,
followed by activation of Nestin transcription
(Fig. 5A, Fig. S4C). In a previous study,
histone hyperacetylation was observed after 5-
aza-dC treatment [29]. Thus, histone
acetylation on the Nestin enhancer after 5-aza-
dC treatment was assessed to discern any
association between histone acetylation on the
Nestin enhancer and Nestin transcription.
Although DNA demethylation of the Nestin
enhancer was observed after 16 hrs of 5-aza-
dC treatment (Fig. 2B), acetylation of histone
H3 on the 5’enhancer was seen to increase
after 24 hrs (Fig. 5B, Fig. S4D), consistent
with the onset of Nestin transcription, thus
suggesting that histone modification is a
prerequisite for Nestin activation. A significant
increase in the acetylation of histone H4—but
not histone H3—on the 3’enhancer was found.
While an increase in histone H4 acetylation on
one of the 3’enhancer elements was not
observed, there was a dramatic increase in
histone H4 acetylation on the other enhancer
elements, including 5’ and 3’enhancers, after
24 hrs of 5-aza-dC treatment (Fig. 5B, Fig.
S4D). Taken together, these data suggest that
histone acetylation is sufficient to mediate
activation of Nestin transcription. Moreover,
histone acetylation on the Nestin gene in P19
EC cells after 8 to 16 hrs of TSA treatment was
comparable to that in NSCs, but the histone
acetylation level in 5-aza-dCítreated P19 EC
cells was still lower than that in NSCs even
after 48 hrs of 5-aza-dC treatment (Fig. 5, Fig.
S4C, D). Similarly, since there was lower
Nestin expression after 48 hrs of 5-aza-dC
treatment than after 4 hrs of TSA treatment
(Fig. S2), DNA demethylation is likely to
occur subsequent to Nestin transcription.
6
Epigenetic hierarchy governing
Nestin
expression
Epigenetic hierarchy in the regulation of
Nestin expression
The data presented herein show that Nestin
transcription could be activated following
acetylation of histones in the absence of DNA
demethylation; however, DNA demethylation
after 5-aza-dC treatment is not sufficient to
mediate Nestin activation. Moreover, TSA and
5-aza-dC influence the epigenetic status of the
Nestin gene by completely different
mechanisms—TSA inhibits HDAC and 5-aza-
dC inhibits DNA methylation. Therefore, to
exclude the possibility of any other epigenetic
modifications induced by TSA or 5-aza-dC
treatment, the role of DNA methylation
following random activation of Nestin
transcription without specific inhibition of
either histone deacetylation or DNA
methylation was next investigated. First, Nestin
non-expressing P19 EC cells were treated with
retinoic acid (RA) for 12, 24, and 48 hrs to
induce cellular differentiation, and the
expression pattern of Nestin was then
examined by RT-PCR. Onset of Nestin
expression in P19 EC cells was observed after
24 hrs of RA treatment, with expression
increasing gradually over time (Fig. 6A).
Down-regulation of Oct4 expression was
detected after RA treatment, which is
consistent with RA-induced differentiation of
pluripotent cells (Fig. S1B). Nestin
transcription was also examined by real-time
RT-PCR (Fig. S1A). Since activation of Nestin
expression was observed after 24 hrs of RA
treatment, but not after 12 hrs, it was
postulated that passive demethylation might
occur after 24 hrs of RA treatment, accounting
for Nestin activation. However, DNA
demethylation of the Nestin enhancer could not
be observed after up to 48 hrs of RA treatment,
suggesting that Nestin activation was mediated
by RA in the absence of DNA demethylation
of the Nestin enhancer elements (Fig. 6B).
Again, the acetylation status of histones H3
and H4 was assessed following RA treatment.
All selected regions exhibited a gradual
increase in H3 acetylation and a more dramatic
increase (1.4-6.8-fold) in H4 acetylation after
24 hrs of RA treatment (Fig. 6C, Fig. S4E).
Taken together, these findings suggest that
DNA demethylation is not sufficient to mediate
Nestin activation, as Nestin expression could
not be detected in P19 EC cells after 5-aza-dC
treatment, i.e. in cells with unmethylated
enhancer elements. However, histone
modification (acetylation) is sufficient to
mediate Nestin activation, as Nestin expression
could be observed after TSA and RA
treatment, i.e. in cells without DNA
demethylation of the Nestin enhancer elements
(Fig. 7C). Therefore, histone modification (i.e.
acetylation) is postulated to be at the top of the
epigenetic regulatory hierarchy governing
Nestin expression, while DNA demethylation
appears to be a dispensable event.
DISCUSSION
Although Nestin has been widely used as a
marker gene to identify and isolate NSCs [2-4],
its regulation remains relatively poorly
understood. The current study was undertaken
to determine the impact of the key epigenetic
features histone deacetylation and DNA
methylation on Nestin gene transcription. An
earlier study examining the regulation of
Nestin expression in transgenic mice [11]
showed that the Nestin upstream promoter
element does not harbor any regulatory
elements specific to NSCs, but it identified two
cell type-specific enhancers in the intron
regions of Nestin: a myogenic precursor
cellspecific enhancer within the first intron
and a CNS-specific enhancer within the second
intron of the rat Nestin gene. The CNS-specific
second intron enhancer is a crucial element in
the regulation of Nestin expression in NSCs
[13]. Moreover, the 3’ part of the second intron
enhancer (3’enhancer) harbors SOX and POU
binding sites, which when mutated lead to
silencing of Nestin expression, indicating that
the Nestin 3’enhancer is the main regulatory
element involved in the positive regulation of
Nestin transcription. The current study showed
that the Nestin gene is differentially methylated
in Nestin-expressing and non-expressing cells
and, moreover, that only the 5’ part of second
intron enhancer (5’enhancer) is differentially
methylated, indicating that 5’enhancer may
also be a crucial element in Nestin
transcription. Nestin transcription was
therefore assessed in Nestin non-expressing EC
cells that had been treated with TSA and 5-aza-
dC—TSA inhibits histone deacetylation and 5-
7
aza-dC inhibits DNA methylation—to
determine the role of DNA methylation in the
regulation of Nestin expression. DNA
demethylation and histone modifications
typically act in a synergistic manner to regulate
the transcription of many genes [16-19].
However, it was found that activation of Nestin
expression could be mediated by the inhibition
of histone acetylation (i.e. of histone
deacetylase [HDAC]) through TSA treatment
in the absence of DNA demethylation on the
5’enhancer. Other neural marker genes were
also found to display similar transcriptional
activation patterns (Fig. S5), indicating that not
all genes are regulated by either equal or
synergic contributions of DNA methylation
and histone modifications. Moreover, onset of
Nestin expression was observed only after 24
hours of 5-aza-dC treatment, even though the
5’enhancer had already been demethylated.
These data suggest two possible roles of DNA
methylation in the mechanisms underlying
Nestin transcription. First, DNA methylation
simply reflects the transcriptional status of
Nestin. Second, and more feasibly, DNA
methylation may be a process secondary to the
activation of Nestin transcription. To verify
this hypothesis, an attempt was made to discern
the presence of an association between Nestin
transcription and DNA methylation of Nestin
in Dnmts KO ES cells, wherein genome-wide
DNA demethylation has normally been
observed. Although undermethylation of the
Nestin 5’enhancer (methylation rate: 0.0-3.1%)
was observed in all Dnmt KO ES cell lines,
there was no activation of Nestin expression in
the absence of histone deacetylation inhibition
induced by TSA treatment (Fig. 3). Therefore,
DNA methylation is not sufficient to mediate
Nestin transcription; it appears to be
dispensable for Nestin activation. Activation of
Nestin transcription was observed
concomitantly with the absence of DNA
demethylation in P19 EC cells that had been
treated with TSA for 4 hrs. However, DNA
demethylation of the 5’enhancer element—
observed after 16 hrs of TSA treatment—was
found to lead to enhanced Nestin transcription
(Fig. S2A), indicating that although DNA
methylation is dispensable for Nestin
activation, it may still represent an important
process secondary to Nestin regulation.
Furthermore, DNA demethylation of the Nestin
5’enhancer was indirectly mediated by TSA
treatment, suggesting that DNA methylation
may play a role in ensuring maximal Nestin
expression. Moreover, upon induction of
further differentiation of Olig2-GFP–positive
NSCs, which had been generated from Olig2-
GFP ES cells that had differentiated into
Olig2-GFP–negative cells, including neurons,
astrocytes, and oligodendrocyte (Fig. 7A),
hypermethylation of the Nestin 5’enhancer
element was observed. This result suggests that
although DNA methylation is not important for
Nestin activation, it may be a secondary
mechanism for suppressing Nestin expression
(Fig. 7B). The histone acetylation status of the
Nestin regulatory elements was then assessed
in both Nestin-expressing NSCs and non-
expressing EC cells. As expected, P19 EC cells
and NSCs were found to display completely
different acetylation patterns of both histones
H3 and H4 on the second intron enhancer,
including the 5’ and 3’enhancers (Fig. 4, Fig.
S4A, B). Although the Nestin upstream
promoter element does not harbor any specific
regulatory machinery in NSCs, significantly
more acetylation (1.8-fold) was observed on
histone H4 on the Nestin promoter in Nestin-
expressing NSCs, indicating that the Nestin
promoter may also be involved in the
epigenetic regulation of Nestin transcription.
Interestingly, histone acetylation of Nestin was
always observed concomitantly with activation
of Nestin transcription in both TSA- and 5-aza-
dC–treated cells. Moreover, Nestin silencing
could be disturbed by RA treatment in the
absence of DNA demethylation, and activation
of Nestin transcription was also observed
concomitantly with histone acetylation of the
Nestin second intron enhancer (Fig. 6, Fig.
7C). A recent study showed that histone
acetylation precedes DNA demethylation, with
a time lag between histone acetylation and
DNA demethylation in the transcription of an
in-vitro methylated reporter plasmid, while the
DNA demethylation of a reporter plasmid
always occurred in a replication-dependent
manner [30]. However, since histone
acetylation appears to be sufficient for
activation of Nestin transcription in the absence
of DNA demethylation, investigations into the
regulation of Nestin expression provide a
8
Epigenetic hierarchy governing
Nestin
expression
model to study the role of the epigenetic
modification histone acetylation in the absence
of DNA demethylation. To elucidate the roles
played by HDACs and histone
acetyltransferase (HATs) in Nestin regulation,
the levels of all HDACs or HATs were
determined in NSCs, ES cells, and MEF cells.
The expression patterns of all HDACs and
HATs were unexpectedly found to be very
similar between cell lines, suggesting that
Nestin activation may be mediated by the
combined effects of many HDACs and HATs
(Fig. S6). The results reviewed in the current
study suggest that histone modifications are
sufficient to mediate activation of Nestin
transcription and thus histone modification is a
higher-order epigenetic feature necessary for
Nestin transcription. Although we were unable
to show that Nestin activation is merely
dependent on the acetylation induced by TSA,
the finding that Nestin activation is always
accompanied by acetylation of the Nestin
enhancer indicates that this acetylation is
indeed both critical and a prerequisite for
Nestin transcription.
ACKNOWLEDGMENTS
We are indebted to all members of the Schöler
laboratory for helpful discussions of the
results. We are especially grateful to Dr.
Natalia Tapia for valuable comments on the
manuscript. This work was supported by the
Federal Ministry of Education and Research
(BMBF) initiative “Cell-Based Regenerative
Medicine” (Grant 01GN0539).
REFERENCES
1. Temple S. The development of neural stem cells.
Nature 2001;414:112-117.
2. Dahlstrand J, Zimmerman LB, McKay RD et al.
Characterization of the human nestin gene reveals a
close evolutionary relationship to neurofilaments. J
Cell Sci 1992;103 (Pt 2):589-597.
3. Kawaguchi A, Miyata T, Sawamoto K et al. Nestin-
EGFP transgenic mice: visualization of the self-
renewal and multipotency of CNS stem cells. Mol
Cell Neurosci 2001;17:259-273.
4. Lendahl U, Zimmerman LB, McKay RD. CNS stem
cells express a new class of intermediate filament
protein. Cell 1990;60:585-595.
5. Frederiksen K, McKay RD. Proliferation and
differentiation of rat neuroepithelial precursor cells
in vivo. J Neurosci 1988;8:1144-1151.
6. Kachinsky AM, Dominov JA, Miller JB. Myogenesis
and the intermediate filament protein, nestin. Dev
Biol 1994;165:216-228.
7. Kachinsky AM, Dominov JA, Miller JB. Intermediate
filaments in cardiac myogenesis: nestin in the
developing mouse heart. J Histochem Cytochem
1995;43:843-847.
8. Frojdman K, Pelliniemi LJ, Lendahl U et al. The
intermediate filament protein nestin occurs
transiently in differentiating testis of rat and mouse.
Differentiation 1997;61:243-249.
9. Terling C, Rass A, Mitsiadis TA et al. Expression of
the intermediate filament nestin during rodent tooth
development. Int J Dev Biol 1995;39:947-956.
10. Jin ZG, Liu L, Zhong H et al. Second intron of
mouse nestin gene directs its expression in
pluripotent embryonic carcinoma cells through POU
factor binding site. Acta Biochim Biophys Sin
(Shanghai) 2006;38:207-212.
11. Zimmerman L, Parr B, Lendahl U et al. Independent
regulatory elements in the nestin gene direct
transgene expression to neural stem cells or muscle
precursors. Neuron 1994;12:11-24.
12. Josephson R, Muller T, Pickel J et al. POU
transcription factors control expression of CNS stem
cell-specific genes. Development 1998;125:3087-
3100.
13. Lothian C, Lendahl U. An evolutionarily conserved
region in the second intron of the human nestin gene
directs gene expression to CNS progenitor cells and
to early neural crest cells. Eur J Neurosci
1997;9:452-462.
14. Yaworsky PJ, Kappen C. Heterogeneity of neural
progenitor cells revealed by enhancers in the nestin
gene. Dev Biol 1999;205:309-321.
15. Tanaka S, Kamachi Y, Tanouchi A et al. Interplay of
SOX and POU factors in regulation of the Nestin
gene in neural primordial cells. Mol Cell Biol
2004;24:8834-8846.
16. Ghoshal K, Datta J, Majumder S et al. Inhibitors of
histone deacetylase and DNA methyltransferase
synergistically activate the methylated
metallothionein I promoter by activating the
transcription factor MTF-1 and forming an open
chromatin structure. Mol Cell Biol 2002;22:8302-
8319.
17. Hwang CK, Song KY, Kim CS et al. Evidence of
endogenous mu opioid receptor regulation by
epigenetic control of the promoters. Mol Cell Biol
2007;27:4720-4736.
18. Yamada N, Hamada T, Goto M et al. MUC2
expression is regulated by histone H3 modification
and DNA methylation in pancreatic cancer. Int J
Cancer 2006;119:1850-1857.
19. Zhang Y, Fatima N, Dufau ML. Coordinated
changes in DNA methylation and histone
9
modifications regulate silencing/derepression of
luteinizing hormone receptor gene transcription. Mol
Cell Biol 2005;25:7929-7939.
20. Lin P, Kusano K, Zhang Q et al. GABAA receptors
modulate early spontaneous excitatory activity in
differentiating P19 neurons. J Neurochem
1996;66:233-242.
21. McBurney MW, Reuhl KR, Ally AI et al.
Differentiation and maturation of embryonal
carcinoma-derived neurons in cell culture. J Neurosci
1988;8:1063-1073.
22. Cheng L, Jin Z, Liu L et al. Characterization and
promoter analysis of the mouse nestin gene. FEBS
Lett 2004;565:195-202.
23. Ou JN, Torrisani J, Unterberger A et al. Histone
deacetylase inhibitor Trichostatin A induces global
and gene-specific DNA demethylation in human
cancer cell lines. Biochem Pharmacol 2007;73:1297-
1307.
24. Selker EU. Trichostatin A causes selective loss of
DNA methylation in Neurospora. Proc Natl Acad Sci
U S A 1998;95:9430-9435.
25. Januchowski R, Dabrowski M, Ofori H et al.
Trichostatin A down-regulate DNA
methyltransferase 1 in Jurkat T cells. Cancer Lett
2007;246:313-317.
26. Xiong Y, Dowdy SC, Podratz KC et al. Histone
deacetylase inhibitors decrease DNA
methyltransferase-3B messenger RNA stability and
down-regulate de novo DNA methyltransferase
activity in human endometrial cells. Cancer Res
2005;65:2684-2689.
27. Li E, Bestor TH, Jaenisch R. Targeted mutation of
the DNA methyltransferase gene results in
embryonic lethality. Cell 1992;69:915-926.
28. Meissner A, Gnirke A, Bell GW et al. Reduced
representation bisulfite sequencing for comparative
high-resolution DNA methylation analysis. Nucleic
Acids Res 2005;33:5868-5877.
29. Takebayashi S, Nakao M, Fujita N et al. 5-Aza-2'-
deoxycytidine induces histone hyperacetylation of
mouse centromeric heterochromatin by a mechanism
independent of DNA demethylation. Biochem
Biophys Res Commun 2001;288:921-926.
30. D'Alessio AC, Weaver IC, Szyf M. Acetylation-
induced transcription is required for active DNA
demethylation in methylation-silenced genes. Mol
Cell Biol 2007;27:7462-7474.
31. Conti L, Pollard SM, Gorba T et al. Niche-
independent symmetrical self-renewal of a
mammalian tissue stem cell. PLoS Biol 2005;3:e283.
32. Do JT, Scholer HR. Comparison of neurosphere cells
with cumulus cells after fusion with embryonic stem
cells: reprogramming potential. Reprod Fertil Dev
2005;17:143-149.
33. Han DW, Do JT, Gentile L et al. Pluripotential
reprogramming of the somatic genome in hybrid
cells occurs with the first cell cycle. Stem Cells
2008;26:445-454.
See www.StemCells.com for supporting information available
online.
10
Epigenetic hierarchy governing
Nestin
expression
Figure 1. Activation of Nestin in P19 and F9 EC cells after TSA and 5-aza-dC treatments. A,B.
The expression profiles of Nestin in TSA- (A) and 5-aza-dC(B) treated P19 and F9 EC cells were
determined by conventional RT-PCR. NSCs were used as a positive control for Nestin expression.
ȕ-actin was used as an RT control.
11
Figure 2. DNA methylation status of the Nestin regulatory elements in P19 and F9 EC cells. A.
DNA methylation status of Nestin regulatory elements, including the promoter and the 5’ and
3’enhancers, was assessed using bisulfite sequencing, with selected regions indicated with gray
boxes. B,C. Demethylation patterns of 5’enhancer after 5-aza-dC (B) and TSA (C) treatments.
Open and filled circles indicate unmethylated and methylated CpGs, respectively.
12
Epigenetic hierarchy governing
Nestin
expression
Figure 3. Transcription and DNA methylation status of Nestin in Dnmts KO ES cell lines. A. DNA
methylation status of the 5’enhancer was assessed using bisulfite sequencing in Dnmts KO ES cell
lines, with the selected region examined indicated with a gray box. Open and filled circles indicate
unmethylated and methylated CpGs, respectively. B. Expression profiles of Nestin in Dnmts KO
ES cell lines were determined by conventional RT-PCR. Wild-type ES cells were used as a
negative control for Nestin expression and ȕ-actin was used as an RT control. To activate Nestin
expression in Dnmts KO ES cell lines, each cell line was treated with TSA for 16 hrs.
13
Figure 4. ChIP analysis of the Nestin regulatory elements in NSCs and P19 EC cells. The upper
panel is a schematic diagram of the Nestin regulatory elements. The gray boxes indicate the primer
positions used for PCR of the ChIP assay. NSCs and P19 EC cells were subjected to ChIP analysis
using antibodies of acetylated versions of histone H3 (AcH3) and H4 (AcH4). Immunoprecipitation
in the absence of antibodies served as negative control. Aliquots of the sonicated cell supernatant
were also analyzed as an input control. PCR of Hprt served as positive control for the antibodies,
and the starting materials (templates) from each sample were empirically determined by measuring
the intensities of Hprt bands. The fold-enrichment values are shown below the respective lanes. In
addition, the acetylation status of Oct4 gene in P19 EC cells was examined, and was consistent the
ChIP analysis.
14
Epigenetic hierarchy governing
Nestin
expression
Figure 5. ChIP analysis of the Nestin enhancer after TSA and 5-aza-dC treatments. A. Histone
acetylation status of P19 EC cells was examined after TSA treatment. Pre-cultured P19 EC cells
were treated with TSA for 4, 6, 8, and 16 hrs. B. Histone acetylation status of P19 EC cells was
examined after 5-aza-dC treatment for 8, 16, 24, 48, and 72 hrs (10 ȝM). The fold-enrichment
values are shown below the respective lanes.
15
Figure 6. Activation of Nestin expression after RA treatment and sequential epigenetic
modifications in P19 EC cells. A. The effect of RA on Nestin expression was determined by
conventional RT-PCR. B. Demethylation pattern of the 5’enhancer after RA treatment as
evidenced by bisulfite sequencing analysis. Open and filled circles indicate unmethylated and
methylated CpGs, respectively. C. Histone acetylation status of Nestin in P19 EC cells was
examined after RA treatment for 12, 24, and 48 hrs (5 ȝM). The fold-enrichment values are shown
below the respective lanes.
16
Epigenetic hierarchy governing
Nestin
expression
Figure 7. Schematic representation of the regulatory mechanism governing Nestin expression. A.
Olig2-GFP-positive NSCs were derived from Olig2-GFP-negative ES cells and further
differentiated into Olig2-GFP-negative cells including neurons (Tuj1), astrocytes (GFAP), and
oligodendrocytes (O4). B. DNA methylation status of 5’enhancer was assessed in Olig2-GFP ES
cells, Olig2-GFP NSCs and differentiated cells. C. The relationship between sequential epigenetic
modifications, such as gene expression, DNA demethylation, and histone acetylation, and treatment
with TSA, 5-aza-dC, and RA is depicted.
17
18
... The second intron of Nestin, which contains a central nervous system (CNS)-specific enhancer [18], can be divided into a 5′ and a 3′ part. Although 3′ part of the second intron of Nestin harbors SOX and POU binding sites, only the 5′ part is differentially methylated in Nestinexpressing neural stem cells (NSC) and -nonexpressing carcinoma cells [19]. We, therefore, measured the methylation in the 5′ part enhancer by PCR with a forward (5′-TAG TTT TTA GGG AGG AGA TTA GAG G-3′) and a reversal primer (5′-CTC TTA CCC CAA ACA CAA CTA AAA C-3′). ...
... In contrast, the first and second introns of Nestin harbor a myogenic precursor cell-specific and an NSC-specific enhancer, respectively [18]. The enhancer from the second intron controls Nestin expression in NSC by methylation [19,25]. We found that the methylation levels at 5′ part of the second intron were significantly decreased at multiple CpG sites (+2333 bp, +2429 bp, +2438 bp) in neurotoxicity group as compared to those in control (Fig. 5a, b). ...
... The increased expressions of Nanog and Nestin in the retina challenged with the neurotoxicity were accompanied by hypomethylation of the genomic regions characterized as a promoter and an NSC-specific enhancer, respectively. Specifically, a significantly reduced methylation was observed in the 5′ part of the enhancer (Fig. 5), which directs transgene expression to NSC and is differentially methylated in Nestin-expressing NSC and -nonexpressing carcinoma cells [18,19]. Methylation or demethylation is a major strategy employed by the cells to regulate gene transcription. ...
Article
Full-text available
Retinal Müller glial cells in mammals acquire stem and progenitor cell properties after neurotoxic treatment. However, the molecular mechanisms underlying proliferation and dedifferentiation of adult Müller cells in the mammalian retina were unclear. In this study, treatments with N-methyl-d-aspartate (NMDA) plus epidermal growth factor (EGF) led to the proliferation of Müller cells and expression of stem cell markers including Nanog and Nestin in the retina. The increased mRNA for Nanog and Nestin were coincident with reduced methylation of a Nanog promoter and a Nestin enhancer specific in the neural stem cells, respectively. We found that Apolipoprotein B mRNA editing catalytic subunit 1 (Apobec1) was upregulated early in the retina treated with NMDA and EGF. Moreover, overexpression of Apobec1 in primary Müller cells increased expression of Nestin and reduced methylation of the Nestin enhancer. The data suggest that neurotoxicity-induced Apobec1 may promote expression of Nestin and help cell cycle reentry of retinal Müller cells via DNA demethylation. This study provides novel insights into the molecular mechanisms underlying dedifferentiation and proliferation of Müller cells in the mammalian retina.
... Moreover, the Nestin gene is controlled by epigenetic modifications such as DNA methylation/demethylation and histone acetylation [22]. In P19 cells, histone acetylation has been reported to be sufficient to mediate Nestin transcription, but DNA methylation was not [22]. ...
... Moreover, the Nestin gene is controlled by epigenetic modifications such as DNA methylation/demethylation and histone acetylation [22]. In P19 cells, histone acetylation has been reported to be sufficient to mediate Nestin transcription, but DNA methylation was not [22]. Thus, YKS may interact with histone modification protein(s) such as histone deacetylase(s), and thereby control Nestin transcription. ...
Article
Full-text available
Yokukansan (YKS), a traditional Japanese Kampo medicine, affects neurological and psychiatric disorders. It ameliorates hippocampal neurogenesis in animals. However, its effect on neuronal cell differentiation remains unclear. Therefore, we investigated the effects of YKS on pluripotent P19 embryonic carcinoma cells as neuronal differentiation model cells. Western blotting and immunocytochemistry revealed that 10 μg/mL YKS treatment during embryoid body formation or neuronal differentiation increased the expression of the neuronal stem cell marker, Nestin, by 1.9-fold and 1.7-fold, respectively, and of the mature neuron marker, NeuN, by 1.5-fold and 1.4-fold, respectively. We examined the effect of YKS on intracellular signaling pathways in P19 cells and found significant elevation in phospho-PDK1 and phospho-mTOR expression (1.1-fold and 1.2-fold, respectively). Therefore, we investigated the effect of PDK1 and mTOR inhibitors on the level of neuronal lineage markers. We found that the mTOR inhibitor significantly abolished the YKS effect on the level of neuronal lineage markers. Moreover, to identify the target(s) of YKS, antibody array analysis that simultaneously detects 16 phosphorylated proteins was performed. YKS significantly upregulated 10 phosphorylated proteins including PDK1, Akt, AMPK, PRAS40, mTOR, p70 S6 kinase, GSK-3α, Bad and ERK1/2 under cell proliferation conditions. These results suggest that YKS simultaneously activates multiple signaling pathways. Thus, we concluded that YKS enhances the level of neuronal lineage markers in differentiated P19 cells, however it does not induce neuronal differentiation. Furthermore, mTOR is the predominant mediator of the YKS effect on these cells.
... Recent studies have revealed several transcription factors responsible for the regulation of Nestin expression in tissue repair. 9,17 During liver fibrosis, the TGFb-Smad2/3 pathway was persistently activated and the transcription factor binding site prediction showed multiple SBEs on the promoter of both human and murine Nestin (Table S2). Thus, we tested whether the TGFb pathway could regulate Nestin expression. ...
Article
Full-text available
Background & aims Liver fibrosis is a wound-healing response that arises from various aetiologies. The intermediate filament protein, Nestin, has been reported to participate in maintaining tissue homeostasis during wound healing responses. However, little is known about the role Nestin plays in liver fibrosis. This study investigated the function and precise regulatory network of Nestin during liver fibrosis. Methods Nestin expression was assessed via immunostaining and quantitative real-time polymerase chain reaction (qPCR) in fibrotic/cirrhotic samples. The induction of Nestin expression by transforming growth factor beta (TGFβ)-Smad2/3 signalling was investigated through luciferase reporter assays. The functional role of Nestin in hepatic stellate cells (HSCs) was investigated by examining the pathway activity of pro-fibrogenic TGFβ-Smad2/3 signalling and degradation of TGFβ receptor I (TβRI) after interfering with Nestin. The in vivo effects of knocking down Nestin were examined with an adeno-associated virus vector (serotype 6, AAV6) carrying short hairpin RNA (shRNA) targeting Nestin in fibrotic mouse models. Results Nestin was mainly expressed in activated HSCs and increased with the progression of liver fibrosis. The pro-fibrogenic pathway TGFβ-Smad2/3 induced Nestin expression directly. Knocking down Nestin promoted Caveolin1 (Cav-1)−mediated TβRI degradation, resulting in TGFβ-Smad2/3 pathway impairment and reduced fibrosis marker expression in HSCs. In AAV6-treated murine fibrotic models, knocking down Nestin resulted in decreased levels of inflammatory infiltration, hepatocellular damage, and a reduced degree of fibrosis. Conclusion The expression of Nestin in HSCs was induced by TGFβ and positively correlated with the degree of liver fibrosis. Knockdown of Nestin decreased activation of TGFβ pathway and alleviated liver fibrosis both in vitro and in vivo. Our data demonstrate a novel role of Nestin in controlling HSC activation in liver fibrosis.
... However, analysis of NSCs derived from ectopic expression of Brn4, Sox2, Klf4, Tcf3, and with or without c-Myc, suggests that the fibroblast cell-fate network is still active in late passage NSC cultures (Han et al., 2012). Interestingly, the neural stem cell marker Nestin, displays differential DNA methylation in NSCs and fibroblasts, with the second intron of the loci being unmethylated in NSCs (Dong et al., 2009). This was evident in directly reprogrammed NSCs, suggesting direct reprogramming to the NSC state retains key DNA methylation signatures (Han et al., 2012). ...
Article
Full-text available
Understanding of cell-type specific transcription factors has promoted progress in methods for cellular reprogramming, such as directly reprogramming somatic cells to induced neurons (iN). Methods for direct reprogramming require neuronal-fate determining gene activation via neuron-specific microRNAs, chemical modulation of key neuronal signaling pathways or overexpression via viral vectors, with some reprogramming strategies requiring a combination of these methods to induce the neuronal-cell fate. These methods have been employed in a multitude of cell types, including fibroblasts, hepatocytes, peripheral blood mononuclear, and T cells. The ability to create iN from skin biopsies and blood samples coupled with recent advancements in artificially inducing age- and disease-associated phenotypes are accelerating the development of disease models for late-onset neurodegenerative disorders. Here, we review how activation of the neuronal transcriptome alters the epigenetic landscape of the donor cell to facilitate reprogramming to neurons. We also discuss the advantages of using DNA binding domains such as CRISPR/dCas9 to overcome epigenetic barriers to induce neuronal-cell fate by activating endogenous neuronal cell-fate determining genes.
... Briefly, PCR amplifications were performed using SuperTaq polymerase (Ambion) in a total volume of 25 mL and a protocol of a total of 40 cycles of denaturation at 94 C for 30 s, annealing at the appropriate temperature for each target region for 30 s, extension at 72 C for 30 s with a 1st denaturation at 94 C for 5 min, and a final extension at 72 C for 10 min. Primer sequences and annealing temperatures were described in our previous study and listed in Table S2 (Han et al., 2009;Hong et al., 2014). For each primer set, 3 mL of product from the first round of PCR was used in the second round of PCR as template. ...
Article
Full-text available
Simultaneous expression of Oct4, Klf4, Sox2, and cMyc induces pluripotency in somatic cells (iPSCs). Replacing Oct4 with the neuro-specific factor Brn4 leads to transdifferentiation of fibroblasts into induced neural stem cells (iNSCs). However, Brn4 was recently found to induce transient acquisition of pluripotency before establishing the neural fate. We employed genetic lineage tracing and found that induction of iNSCs with individual vectors leads to direct lineage conversion. In contrast, polycistronic expression produces a Brn4-Klf4 fusion protein that enables induction of pluripotency. Our study demonstrates that a combination of pluripotency and tissue-specific factors allows direct somatic cell transdifferentiation, bypassing the acquisition of a pluripotent state. This result has major implications for lineage conversion technologies, which hold potential for providing a safer alternative to iPSCs for clinical application both in vitro and in vivo.
... This silencing and remodeling displaces Brg1 for the Brahma-(Brm-) containing SWI/SNF complex on both the Oct4 and Nanog promoters, thus silencing transcription [275]. Although comparable conditions exist in ESCs [276] this is not a universal phenomenon as histone acetylation near the Nestin locus accompanies RAinduced differentiation in P19 cells [277]. It is interesting to note that RAR −/− F9 cells exhibit increased expression of Slc38a and Stmn2, which is normally associated with differentiated F9 cells due to hyperacetylation, suggesting that RA signaling might have a role in regulating histone 8 Stem Cells International modification [278]. ...
Article
Full-text available
Just over ten years have passed since the seminal Takahashi-Yamanaka paper, and while most attention nowadays is on induced, embryonic, and cancer stem cells, much of the pioneering work arose from studies with embryonal carcinoma cells (ECCs) derived from teratocarcinomas. This original work was broad in scope, but eventually led the way for us to focus on the components involved in the gene regulation of stemness and differentiation. As the name implies, ECCs are malignant in nature, yet maintain the ability to differentiate into the 3 germ layers and extraembryonic tissues, as well as behave normally when reintroduced into a healthy blastocyst. Retinoic acid signaling has been thoroughly interrogated in ECCs, especially in the F9 and P19 murine cell models, and while we have touched on this aspect, this review purposely highlights how some key transcription factors regulate pluripotency and cell stemness prior to this signaling. Another major focus is on the epigenetic regulation of ECCs and stem cells, and, towards that end, this review closes on what we see as a new frontier in combating aging and human disease, namely, how cellular metabolism shapes the epigenetic landscape and hence the pluripotency of all stem cells.
... Briefly, unmethylated cytosine nucleotides in genomic DNA were converted into uracil nucleotides upon sodium bisulfite treatment. To obtain enough PCR product, two rounds of PCR were performed using SuperTaq (Ambion, Waltham, MA, USA) according to our previous reports [28,41]. Three microliters of product from the first PCR reaction were used as template for the second round of PCR. ...
Article
Full-text available
Parkinson’s disease (PD) is a chronic, neurodegenerative disorder that results from the loss of cells in the substantia nigra (SN) which is located in the midbrain. However, no cure is available for PD. Recently, fibroblasts have been directly converted into induced neural stem cells (iNSCs) via the forced expression of specific transcription factors. Therapeutic potential of iNSC in PD has not been investigated yet. Here, we show that iNSCs directly converted from mouse fibroblasts enhanced functional recovery in an animal model of PD. The rotational behavior test was performed to assess recovery. Our results indicate that iNSC transplantation into the striatum of 6-hydroxydopamine (6-OHDA)-injected mice can significantly reduce apomorphine-induced rotational asymmetry. The engrafted iNSCs were able to survive in the striatum and migrated around the medial forebrain bundle and the SN pars compacta. Moreover, iNSCs differentiated into all neuronal lineages. In particular, the transplanted iNSCs that committed to the glial lineage were significantly increased in the striatum of 6-OHDA-injected mice. Engrafted iNSCs differentiated to dopaminergic (DA) neurons and migrated into the SN in the 6-OHDA lesion mice. Therefore, iNSC transplantation serves as a valuable tool to enhance the functional recovery in PD.
Article
Background: Nestin is an intermediate filament first reported in neuroepithelial stem cells. Nestin expression could be found in a variety of tissues throughout all systems of the body, especially during tissue development and tissue regeneration processes. Aim of review: This review aimed to summarize and discuss current studies on the distribution, contribution and regulation of nestin+ cells in different systems of the body, to discuss the feasibility ofusing nestin as a marker of multilineage stem/progenitor cells, and better understand the potential roles of nestin+ cells in tissue development, regeneration and pathological processes. Key scientific concepts of review: This review highlights the potential of nestin as a marker of multilineage stem/progenitor cells, and as a key factor in tissue development and tissue regeneration. The article discussed the current findings, limitations, and potential clinical implications or applications of nestin+ cells. Additionally, it included the relationship of nestin+ cells to other cell populations. We propose potential future research directions to encourage further investigation in the field.
Article
Full-text available
The transcription factor Ptf1a is a crucial helix–loop–helix (bHLH) protein selectively expressed in the pancreas, retina, spinal cord, brain, and enteric nervous system. Ptf1a is preferably assembled into a transcription trimeric complex PTF1 with an E protein and Rbpj (or Rbpjl). In pancreatic development, Ptf1a is indispensable in controlling the expansion of multipotent progenitor cells as well as the specification and maintenance of the acinar cells. In neural tissues, Ptf1a is transiently expressed in the post-mitotic cells and specifies the inhibitory neuronal cell fates, mostly mediated by downstream genes such as Tfap2a/b and Prdm13. Mutations in the coding and non-coding regulatory sequences resulting in Ptf1a gain- or loss-of-function are associated with genetic diseases such as pancreatic and cerebellar agenesis in the rodent and human. Surprisingly, Ptf1a alone is sufficient to reprogram mouse or human fibroblasts into tripotential neural stem cells. Its pleiotropic functions in many biological processes remain to be deciphered in the future.
Article
Full-text available
Induced neural stem cells (iNSCs) reprogrammed from somatic cells have great potentials in cell replacement therapies and in vitro modeling of neural diseases. Direct conversion of fibroblasts into iNSCs has been shown to depend on a couple of key neural progenitor transcription factors (TFs), raising the question of whether such direct reprogramming can be achieved by non-neural progenitor TFs. Here we report that the non-neural progenitor TF Ptf1a alone is sufficient to directly reprogram mouse and human fibroblasts into self-renewable iNSCs capable of differentiating into functional neurons, astrocytes and oligodendrocytes, and improving cognitive dysfunction of Alzheimer's disease mouse models when transplanted. The reprogramming activity of Ptf1a depends on its Notch-independent interaction with Rbpj which leads to subsequent activation of expression of TF genes and Notch signaling required for NSC specification, self-renewal, and homeostasis. Together, our data identify a non-canonical and safer approach to establish iNSCs for research and therapeutic purposes.
Article
Full-text available
Important features of adult neuronal number, location, and type are a consequence of early embryonic events that occur before neurons have differentiated. We have measured cell number during embryogenesis of the rat CNS. Markers that are expressed in the proliferating neuronal precursor are required to study the mechanisms controlling their proliferation and differentiation. By applying immunohistochemistry, fluorescence-activated cell sorting, and 3H-thymidine auto-radiography to dissociated rat CNS cells, we show that the monoclonal antibody Rat 401 recognizes a cell population with proliferative, temporal, and quantitative features expected of neuronal precursors.
Article
Full-text available
Multipotential stem cells in the neural tube give rise to the different neuronal cell types found in the brain. Abrupt changes in intermediate filament gene expression accompany this transition out of the precursor state: transcription of the intermediate filament nestin is replaced by that of the neurofilaments. In order to identify human neural precursor cells, and to learn more about the evolution of the intermediate filaments expressed in the central nervous system, we have isolated the human nestin gene. Despite considerable divergence between the human and rat nestin genes, in particular in the repetitive parts of the carboxy-terminal region, the positions of the introns are perfectly conserved. Two of the three intron positions are also shared by the neurofilaments, but not by other classes of intermediate filaments. This implies that nestin and the neurofilaments had a common ancestor after branching off from the other classes of intermediate filaments, and that nestin separated from the neurofilament branch before the different neurofilament genes diverged. The characterization of human nestin also facilitates the identification of human multipotential neural precursor cells. This will be of importance for central nervous system (CNS) tumor diagnosis and transplant-based clinical approaches to human neurodegenerative diseases.
Article
Full-text available
We have previously shown that retinoic acid-treated cultures of the P19 line of embryonal carcinoma cells differentiate into neurons, glia, and fibroblast-like cells (Jones-Villeneuve et al., 1982). We report here that the monoclonal antibody HNK-1 reacts with the neurons at a very early stage of their differentiation and is, therefore, an early marker of the neuronal lineage. Cells in differentiated P 19 cultures synthesized acetylcholine but not catecholamines, suggesting that at least some of the neurons are cholinergic. The neurons also carry high-affinity uptake sites for GABA but not for serotonin. In long-term cultures, neuronal processes differentiated into axons and dendrites, which formed synapses. This biological system should prove valuable for examining the development and maturation of cholinergic neurons, since their differentiation occurs in cell culture.
Article
Full-text available
Important features of adult neuronal number, location, and type are a consequence of early embryonic events that occur before neurons have differentiated. We have measured cell number during embryogenesis of the rat CNS. Markers that are expressed in the proliferating neuronal precursor are required to study the mechanisms controlling their proliferation and differentiation. By applying immunohistochemistry, fluorescence-activated cell sorting, and 3H-thymidine auto-radiography to dissociated rat CNS cells, we show that the monoclonal antibody Rat 401 recognizes a cell population with proliferative, temporal, and quantitative features expected of neuronal precursors.
Article
The intermediate filament protein nestin is expressed in the neural stem cells of the developing central nervous system (CNS). Promoter analysis revealed that the minimal promoter of the mouse nestin gene resides in the region −11 to +183 of the 5′-non-coding and upstream flanking region, and that two adjacent Sp1-binding sites are necessary for promoter activity. Electrophoretic mobility-shift assays (EMSA) and supershift assays showed that Sp1 and Sp3 proteins selectively bind to the upstream Sp1 site. These results demonstrate an important functionality of Sp1 and Sp3 in regulating the expression of the mouse nestin gene.
Article
Central nervous system (CNS) progenitor cells transiently proliferate in the embryonic neural tube and give rise to neurons and glial cells. A characteristic feature of the CNS progenitor cells is expression of the intermediate filament nestin and it was previously shown that the rat nestin second intron functions as an enhancer, directing gene expression to CNS progenitor cells. In this report we characterize the nestin enhancer in further detail. Cloning and sequence analysis of the rat and human nestin second introns revealed local domains of high sequence similarity in the 3′ portion of the introns. Transgenic mice were generated with the most conserved 714 bp in the 3′ portion of the intron, or with the complete, 1852 bp, human second intron, coupled to the reporter gene lacZ. The two constructs gave a very similar nestin-like expression pattern, indicating that the important control elements reside in the 714 bp element. Expression was observed starting in embryonic day (E)7.5 neural plate, and at E10.5 CNS progenitor cells throughout the neural tube expressed lacZ. At E12.5, lacZ expression was more restricted and confined to proliferating regions in the neural tube. An interesting difference, compared to the rat nestin second intron, was that the human intron at E10.5 mediated lacZ expression also in early migrating neural crest cells, which is a site of endogenous nestin expression. In conclusion, these data show that a relatively short, evolutionarily conserved region is sufficient to control gene expression in CNS progenitor cells, but that the same region differs between rodents and primates in its capacity to control expression in neural crest cells.
Article
P19 embryonic carcinoma (EC) stem cells are pluripotent and are efficiently induced to differentiate into neurons and glia with retinoic acid (RA) treatment. Within 5 days, a substantial number of differentiating P19 cells express gene products that are characteristic of a neuronal phenotype. P19 neurons were used as a model to explore the relationship between neuronal “differentiation” in vitro and the acquisition of γ-aminobutyric acid (GABAA) receptors and functional GABA responses. Pulse-labeling experiments using bromodeoxyuridine indicated that all neurons had become postmitotic within 3–4 days after treatment with RA. This was confirmed by a reduction in the immunocytochemical detection of the undifferentiated stem cell antigen SSEA-1. Subsequently, a transient expression of nestin was observed during the first 5 days in vitro (DIV) after exposure to RA. By 5–10 DIV after RA, a significant number of neurons (∼80–90%) expressed immunocytochemically detectable glutamate decarboxylase and GABA coincident with the acquisition of membrane binding sites for tetanus toxin. These phenotypic markers were maintained for >30 DIV after RA. Under current-clamp conditions, random, low-amplitude, spontaneous electrical activity appeared in neurons within the first few days after RA treatment and this was blocked by the specific GABAA receptor antagonist bicuculline. Thereafter, the appearance and progressive increases in the frequency of spontaneous action potentials in P19 neurons were observed that were similarly attenuated by bicuculline. In neurons > 5 DIV after RA, exogenous application of GABA elicited similar action potentials. The onset of excitatory responses to GABA or muscimol in voltage-clamped neurons appeared immediately after the cessation of neuronal mitosis and before the previously reported acquisition of responses to glutamate. In fura-2 imaging studies, the exogenous application of GABA resulted in neuron-specific increases in intracellular Ca2+. Thus, P19 neurons provide an in vitro model for the study of the early acquisition and properties of electrical excitability to GABA and the expression of functional GABAA receptors.
Article
We generated transgenic mice carrying enhanced green fluorescent protein (EGFP) under the control of the nestin second-intronic enhancer (E/nestin:EGFP). Flow cytometry followed by in vitro assays revealed that in situ EGFP expression in the embryonic brain correlated with the mitotic index, the cogeneration of both neurons and glia, and the frequency of neurosphere formation in vitro. High-level EGFP expressors derived from embryos included a distinct subpopulation of cells that were self-renewable and multipotent, criteria that define neural stem cells (NSCs). Such cells were largely absent among lower-level or non-EGFP expressors, thereby permitting us to enrich for NSCs using EGFP expression level. In adults, although E/nestin:EGFP-positive cells included the NSC population, the frequency of neurosphere formation did not correlate directly with the level of EGFP expression. However, moderately EGFP-expressing cells in adults gained EGFP intensity when they formed neurospheres, suggesting embryonic and adult NSCs exist in different microenvironments in vivo.
Article
Gene targeting in embryonic stem (ES) cells has been used to mutate the murine DNA methyltransferase gene. ES cell lines homozygous for the mutation were generated by consecutive targeting of both wild-type alleles; the mutant cells were viable and showed no obvious abnormalities with respect to growth rate or morphology, and had only trace levels of DNA methyltransferase activity. A quantitative end-labeling assay showed that the level of m5C in the DNA of homozygous mutant cells was about one-third that of wild-type cells, and Southern blot analysis after cleavage of the DNA with a methylation-sensitive restriction endonuclease revealed substantial demethylation of endogenous retroviral DNA. The mutation was introduced into the germline of mice and found to cause a recessive lethal phenotype. Homozygous embryos were stunted, delayed in development, and did not survive past mid-gestation. The DNA of homozygous embryos showed a reduction of the level of m5C similar to that of homozygous ES cells. These results indicate that while a 3-fold reduction in levels of genomic m5C has no detectable effect on the viability or proliferation of ES cells in culture, a similar reduction of DNA methylation in embryos causes abnormal development and embryonic lethality.
Article
Multipotential CNS stem cells receive and implement instructions governing differentiation to diverse neuronal and glial fates. Exploration of the mechanisms generating the many cell types of the brain depends crucially on markers identifying the stem cell state. We describe a gene whose expression distinguishes the stem cells from the more differentiated cells in the neural tube. This gene was named nestin because it is specifically expressed in neuroepithelial stem cells. The predicted amino acid sequence of the nestin gene product shows that nestin defines a distinct sixth class of intermediate filament protein. These observations extend a model in which transitions in intermediate filament gene expression reflect major steps in the pathway of neural differentiation.