ArticlePDF AvailableLiterature Review

Epigenetics in Stroke Recovery

Authors:

Abstract

While the death rate from stroke has continually decreased due to interventions in the hyperacute stage of the disease, long-term disability and institutionalization have become common sequelae in the aftermath of stroke. Therefore, identification of new molecular pathways that could be targeted to improve neurological recovery among survivors of stroke is crucial. Epigenetic mechanisms such as post-translational modifications of histone proteins and microRNAs have recently emerged as key regulators of the enhanced plasticity observed during repair processes after stroke. In this review, we highlight the recent advancements in the evolving field of epigenetics in stroke recovery.
genes
G C A T
T A C G
G C A T
Review
Epigenetics in Stroke Recovery
Haifa Kassis 1, Amjad Shehadah 1, Michael Chopp 1,2 and Zheng Gang Zhang 1, *
1Department of Neurology, Henry Ford Health System, Detroit, MI 48202, USA;
haifa.kassis@gmail.com (H.K.); amjad.shehadah@gmail.com (A.S.); michael.chopp@gmail.com (M.C.)
2Department of Physics, Oakland University, Rochester, MI 48309, USA
*Correspondence: zzhang1@hfhs.org; Tel.: +1-313-916-5456
Academic Editor: Dennis R. Grayson
Received: 18 November 2016; Accepted: 20 February 2017; Published: 27 February 2017
Abstract: Abstract:
While the death rate from stroke has continually decreased due to interventions
in the hyperacute stage of the disease, long-term disability and institutionalization have become
common sequelae in the aftermath of stroke. Therefore, identification of new molecular pathways that
could be targeted to improve neurological recovery among survivors of stroke is crucial. Epigenetic
mechanisms such as post-translational modifications of histone proteins and microRNAs have
recently emerged as key regulators of the enhanced plasticity observed during repair processes after
stroke. In this review, we highlight the recent advancements in the evolving field of epigenetics in
stroke recovery.
Keywords: epigenetics; stroke; recovery
1. Introduction
Stroke remains a major health care challenge despite the increasing availability of acute
thrombolytic interventions such as tissue plasminogen activator (tPA) and endovascular treatment
strategies [
1
]. While the death rate from stroke has continually decreased due to interventions in the
hyperacute stage of the disease, long-term disability and institutionalization have become common
sequelae in the aftermath of stroke. Stroke costs the US economy $34 billion annually, including the
costs of health care services, medications, and lost productivity [
2
]. Therefore, identification of new
molecular pathways that could be targeted to improve neurological recovery among survivors of
stroke is crucial.
After stroke, restorative processes, such as angiogenesis, neurogenesis, oligodendrogenesis,
synaptogenesis, and axonal outgrowth are induced in the areas adjacent to the infarct border
and contribute to neurological recovery after ischemia [
3
,
4
]. The underlying mechanisms of these
processes involve altered orchestrated expression of genes. Epigenetic mechanisms such as DNA
methylation, post-translational modifications of histone proteins and microRNAs (miRNAs) regulate
gene expression without directly changing the sequence of DNA. Considering the far-reaching effects
of epigenetic mechanisms on gene regulation, they can potentially regulate key aspects of the enhanced
plasticity observed during repair processes after stroke. In this review, we will highlight recent
studies that have expanded our understanding of the roles of epigenetics in regulating stroke recovery.
These exciting discoveries take us a step forward towards the development of epigenetic-based novel
therapies aiming to improve neurological function among stroke patients.
2. DNA Methylation
Methylation of cytosine in CpG (cytosine-guanine) dinucleotides is a repressive epigenetic
mechanism catalyzed by a family of DNA methyltransferase enzymes (DNMTs). DNMT1 is the most
Genes 2017,8, 89; doi:10.3390/genes8030089 www.mdpi.com/journal/genes
Genes 2017,8, 89 2 of 10
abundant DNMT in mammalian cells and binds to hemi-methylated DNA to maintain methylation,
while DNMT3a and 3b bind unmethylated DNA and are responsible for de novo methylation [5].
DNMTs are expressed in postmitotic neurons [
6
] and the global level of DNA methylation
increases
in vivo
acutely after mild ischemic injury/reperfusion. Administration of the broad-spectrum
DNMT inhibitors 5-aza-2
0
-deoxycytidine and Zebularine to wild type rodents provides resistance and
reduces ischemic injury which may indicate that DNA methylation contributes to cell death through
silencing of neuroprotective genes [
7
9
]. Clinical studies have also shown that aberrant blood DNA
methylation levels are associated with increased risk for atherosclerosis and stroke [10,11].
Studies have shown that DNA methylation is implicated in physiologic repair mechanisms such
as adult neurogenesis and synaptic plasticity. For example, members of the methyl-CpG-binding
proteins (MBDs) family of proteins which bind to methylated DNA play a role in the regulation of adult
hippocampal neurogenesis [
12
]. MBD1 binds to methylation sites in the promoter of basic fibroblast
growth factor 2 (FGF2) gene and loss of MBD1 impairs neural progenitor cells differentiation [
13
].
Methyl-CpG-binding protein 2 (MeCP2) is another MBD protein important in the formation of dendritic
spines and synaptogenesis of hippocampal neurons [
14
]. Both MBD1 and MeCP2 are highly expressed
in the brain and upregulated in the hippocampus at 24 h after stroke [
15
] which suggests a role for
DNA methylation in the regulation of stroke-induced neurogenesis and synaptogenesis.
3. Histone Acetylation in Stroke Recovery
The nucleosome is the basic unit of chromatin comprised of approximately 200 bp of DNA
wrapped around an octamer of histone proteins [
16
]. Each histone octamer consists of two copies
of each of the four core histone proteins (H2A, H2B, H3, and H4). Acetylation of lysine residues in
histone proteins removes their positive charge resulting in detachment of the negatively charged DNA
from histone proteins. Subsequently, acetylation leads to relaxation of chromatin (i.e., euochromatin),
which is usually associated with increased gene transcription due to increased access to DNA promoter
regions by transcription factors and RNA transcription machinery. Conversely, deacetylation restores
the positive charge to histone proteins leading to compaction of chromatin (i.e., heterochromatin) and
repression of gene expression [17].
Typically, histone acetylation and deacetylation states are controlled by maintenance of
an appropriate balance between the enzymatic activities of histone acetyltransferases (HATs) and
histone deacetylases (HDACs) [
18
]. Stroke induces a global reduction in acetylation levels of histones
H3 [
19
21
] and H4 [
22
] in the ischemic brain. Histone hypoacetylation starts within hours and lasts
at least up to two weeks [
23
] after the ischemic event. This disruption in the global acetylation
homeostasis appears to occur due to a combination of increased activity of HDACs [
24
] along with
decreased activity of HATs [25].
The observed histone hypoacetylation in the ischemic brain led to the hypothesis that HDAC
inhibitors that increase histone acetylation levels may be a viable approach to treat stroke. Indeed,
many non-selective HDAC inhibitors such as valproic acid (VPA) [
20
], suberoylanilide hydroxamic
acid (SAHA) [
26
], 4-phenylbutyrate [
27
], sodium butyrate, and trichostatin A [
19
] have been tested
in animal models of stroke and were shown to induce neuroprotection and reduce inflammation
when administered within a few hours after stroke onset. Most recently, HDAC inhibitors have
also been increasingly investigated as neurorestorative agents for stroke. Similar to their positive
effects on hyperacute neuronal protection from cell death, delayed administration (24 h after the
ischemic event) of non-selective HDAC inhibitors such as VPA and SAHA were found to improve
functional outcome through increased axonal and dendritic outgrowth and white matter repair [
24
,
28
].
VPA treatment also enhances stroke-induced angiogenesis in animal models of ischemic stroke likely
through upregulation of hypoxia-inducible factor 1-
α
(HIF-1
α
) and vascular endothelial growth factor
(VEGF) [
22
]. Furthermore, non-selective HDAC inhibition with sodium butyrate induces brain-derived
neurotrophic factor (BDNF)-receptor tropomyosin receptor kinase B (TrkB)-dependent stimulation of
neurogenesis in the ischemic brain [23].
Genes 2017,8, 89 3 of 10
HDACs comprise a super-family of enzymes grouped into four major classes (Classes I–IV) based
on their structure and homology to yeast enzymes [
29
]. The expression patterns of the different HDAC
isoforms are differentially regulated by ischemia and show cell- and region-specific patterns seven
days after stroke [
22
,
30
]. For example, oligodendrocyte progenitor cells in the peri-infarct white matter
exhibit increased expression of HDAC1 and HDAC2, concurrent with increased proliferation, while
mature oligodendrocytes, on the other hand, show decreased HDAC1 and increased HDAC2 [
22
].
These data suggest that individual HDAC isoforms within the same class may have differential effects
on neurorestoration and oligodendrogenesis during brain repair after stroke.
The actions of HDACs are even more complex as some of the isoforms possess the ability to
translocate between the cytoplasm and nucleus. For example, Class IIa HDACs (4 and 5) are mainly
expressed in the cytoplasm under physiological conditions, but are also able to shuttle into the nucleus.
In the nucleus, HDACs can access histone proteins and act as epigenetic regulators [
31
]. The role of
nuclear HDAC4 in neuronal cell death and survival is controversial and highly debated in the literature.
Previous studies have yielded contradictory results with some concluding that HDAC4 promotes cell
death [
32
,
33
], while others have provided evidence that nuclear HDAC4 shuttling protects neurons
from injury [
34
]. The controversy may stem from the diverse actions that HDAC4 may play depending
on environmental cues under different
in vitro
conditions [
35
]. During recovery
in vivo
, stroke induces
nuclear shuttling of HDAC4 (but not HDAC5) in neurons localized to the peri-infarct, and increased
nuclear HDAC4 is strongly associated with endogenous neuronal repair, suggesting a positive role for
HDAC4 in promoting neuronal recovery after ischemic injury [
36
]. HDAC remains highly localized to
the nucleus up to at least four weeks after the ischemic event. Interestingly, unlike the non-selective
HDAC inhibitors mentioned above, when the selective Class IIa HDAC inhibitor MC1568 was tested
in vivo
, it was found to increase mortality and lesion volume and impair neuronal remodeling after
stroke; possibly through downregulation of phosphorylated cAMP response element-binding protein
(CREB) and c-fos in neurons [
24
]. These studies further confirm the existence of isoform-specific roles
for HDACs in endogenous brain recovery after ischemic stroke. Some isoforms such as HDAC4 exhibit
restorative effects, while others such as HDAC6 [
37
] may promote cell death. Development of more
selective inhibitor/activators of HDAC isoforms may provide a novel therapeutic approach that can
potentially elucidate HDAC isoform specific functions and thereby facilitate enhanced brain recovery
after stroke.
Since very few studies in humans explored the roles of histone acetylation in stroke, most of our
knowledge comes from experimental animal studies. However, in an effort to identify genetic factors
that contribute to increased risk of stroke, a large genome-wide association study has identified that
a variant in HDAC9 on chromosome 7p21.1 is associated with a 1.4-fold increase in risk for large vessel
disease ischemic stroke [
38
]. HDAC9 expression was also found to be upregulated in human carotid
plaques compared with controls [
39
]. Whether HDAC9 plays a role during recovery after stroke is
currently unknown.
Utilizing imaging technology to measure HDAC expression acutely and during recovery in stroke
patients is clearly of great interest. A recent study shows that the expression of histone deacetylases in
the human brain can be non-invasively imaged using Positron Emission Tomography (PET) imaging
with [11C] Martinostat [
40
]. The authors found that HDACs are highly expressed throughout the
healthy human brain and display region-specific distribution. This technology can potentially provide
an additional tool for more clinical research in the area of histone acetylation.
Finally, many HDAC inhibitors are currently under intense investigation in clinical trials for
their potential use as anticancer drugs [
41
]. The U.S. Food and Drug Administration (FDA) has
approved a number of HDAC inhibitors including SAHA [
42
], romidepsin [
43
], belinostat [
44
], and
panobinostat [
45
] for the treatment of cutaneous/peripheral T-cell lymphoma and multiple myeloma.
Whether HDAC-based therapy will yield similar positive outcomes in stroke patients remains an open
question that needs to be addressed in future clinical trials.
Genes 2017,8, 89 4 of 10
4. MicroRNAs in Stroke Recovery
MicroRNAs are abundant small (20–25 nucleotides) non-coding RNAs that regulate gene
transcription via blockage of translation of messenger RNA (mRNAs) into proteins [
46
]. miRNAs are
produced by a multi-step canonical mechanism that include transcription of a long hairpin-containing
primary miRNA (pri-mRNA) by RNA polymerase II. The pri-miRNA is then cleaved by Drosha into
pre-miRNA and exported into the cytoplasm by Exportin 5. In the cytoplasm, pre-miRNA is cleaved
by Dicer and then binds with Argonaute (Ago) proteins in the RNA-induced silencing complexes
(RISCs), which silence specific mRNA transcripts based on complementary to unique 3
0
UTR sequence
motifs [47].
miRNAs are highly expressed in the nervous system where they play key roles in development,
physiology, and disease [
48
]. Considering the diversity of miRNA functions and their influence on a
large number of neuronal and non-neuronal genes in experience and activity-dependent manners [
49
],
it is not surprising that miRNAs have started to emerge as important players in stroke-induced
endogenous brain recovery events such as angiogenesis, neurogenesis, oligodendrogenesis, and axonal
outgrowth (Table 1).
Table 1. Key miRNAs altered by stroke and their potential roles in ischemic brain repair processes.
Brain Repair miRNA Effects in Ischemia References
Angiogenesis
miR-139; miR-335
Downregulated after stroke; promote
angiogenesis in vitro when upregulated [50]
miR-15a
Upregulated after stroke; promotes
angiogenesis when blocked via targeting of
BDNF and VEGF
[51,52]
miR-155 In vivo inhibition of miR-155 leads to
revascularization and BBB preservation [53]
miR-107 Upregulated in ischemic boundary zone
in vivo
; contributes to post-stroke angiogenesis
[54]
Neurogenesis miR-124 Downregulated after stroke;shows
pro-proliferative effect on SVZ stem cells [55]
miR-17-92 cluster
Upregulated after stroke; suppresses PTEN and
promotes neural stem cell proliferation [56]
Oligodendrogenesis miR-9; miR-200b Downregulated after stroke;regulate
oligodendrogenesis via targeting of SRF [57]
miR-146a Upregulated after stroke;enhances OPC
differentiation when overexpressed [58]
* miR denotes microRNA, BDNF brain-derived neurotrophic factor, VEGF vascular endothelial growth factor, BBB
blood–brain barrier, SVZ sub-ventricular zone, PTEN phosphatase and tensin homolog, SRF serum response factor,
OPC oligodendrocyte progenitor cell.
The adult brain vasculature is activated in response to various pathological conditions including
stroke [
59
]. Neurorestorative treatments, either cell-based or pharmacological therapies, in animal
models of stroke induce angiogenesis, the formation of new blood vessels, which is associated with
and may underlie improvements in neurological outcome [
3
]. Recent studies have indicated that
specific miRNAs play key roles in endothelial function and angiogenesis [
60
]. Altered miRNA
expression profiles were found in rat cerebral endothelial cells at seven days after stroke; 7 miRNAs
(miR-225, miR-335, miR-139-5p, miR-203, miR-708, miR-193*, and miR-494) are downregulated, and
another 9 miRNAs (miR-224, miR-210, miR-204, miR-322*, miR-100, miR-450a, miR-322, miR-331, and
miR-101a) are upregulated. Of those, miR-139 and miR-335 are most strikingly downregulated
at 60% and 90%, respectively; when the expressions of miR-139 and miR-335 are exogenously
restored in cultured cerebral endothelial cells, they promote angiogenesis as measured by the capillary
tube formation assay
in vitro
[
50
]. Similarly, another study has found that miR-15 is upregulated
in endothelial cells up to 16 h after oxygen deprivation exposure
in vitro
and provided evidence
Genes 2017,8, 89 5 of 10
suggesting that pharmacological inhibition of miR-15a may be a potential approach to increasing
angiogenesis after stroke through targeting of basic fibroblast growth factor (bFGF) and VEGF in
endothelial cells [
51
,
52
]. miR-155 is another key regulator of endothelial morphogenesis that was
recently implicated as a potential therapeutic target for stroke. When a specific miR-155 inhibitor was
administered two days after ischemia in an experimental mouse stroke model, it improved functional
outcome, decreased vascular leakage, and promoted revascularization through stabilization of tight
junctions and preservation of blood–brain barrier [
53
]. Another miRNA, miR-107, was also shown to be
upregulated in the ischemic boundary zone on day 3 and day 7 after permanent middle cerebral artery
occlusion in the rat and to contribute to post-stroke angiogenesis by targeting Dicer-1, an enzyme
which regulates processing of miRNA precursors and regulates of VEGF in endothelial cells [
54
].
Taken together, these studies highlight the important roles of endothelial miRNAs and their potential
as therapeutic targets for enhancing vascular remodeling and recovery after ischemia.
In addition to endothelial cell activation in response to injury, the adult brain neural stem cells
are also stimulated after ischemia. Stroke increases the proliferation rate of neural stem cells in
the sub-ventricular zone (SVZ) of the lateral ventricle and induces migration of newly generated
neuroblasts to the ischemic boundary area where they play a major role in promoting recovery after
stroke [
61
,
62
]. miRNA profiling studies have revealed that miRNAs involved in the regulation of key
signaling pathways in stem cell proliferation and differentiation such as Notch, Wnt, Sonic Hedgehog
(SHH), and TGF-
β
are significantly altered after ischemia [
63
,
64
]. For example, miR-124a, the most
abundant miRNA in neurons, is downregulated seven days after stroke and contributes to the dramatic
increase in neural progenitor cell proliferation after ischemia [
55
]. The pro-proliferative effects of
knocking down miR-124 are mediated through dampening of its inhibition of Jagged-1; Notch receptor
ligand 1, a major promoter of neural stem cell proliferation after stroke [
65
]. Indeed, exogenously
increasing the expression of miR-124 negatively impacts the proliferation of rat neural stem cells,
suggesting that stroke-induced downregulation of miR-124 is an important mediator of neurogenesis
after stroke.
miR-17, miR-18a, miR-19a, miR-20a, miR-19b-1, and miR-92a-1 are members of the miR-17-92 cluster,
also known as oncomiR-1 due to its involvement in tumorigenesis [
66
]. In addition, this cluster also
plays important roles in regulating neural progenitor cell proliferation and oligodendrogenesis during
development [
67
,
68
]. Stroke-induced activation of the SHH pathway upregulates the expression of the
miR-17-92 cluster in SVZ neural progenitor cells via c-Myc, one of the most potent oncogenic genes [
56
].
Overexpression of the miR 17-92 cluster members promotes stroke-induced neural progenitor cell
proliferation possibly through suppression of phosphatase and tensin homolog (PTEN) deleted on
chromosome 10, a tumor suppressor gene that negatively regulates cell proliferation [
56
]. Collectively,
these studies highlight the roles of SVZ miRNAs in regulating stroke-induced neurogenesis.
In addition to generation of new neuroblasts in the adult brain, SVZ progenitor cells also generate
oligodendrocyte progenitor cells (OPCs) able to differentiate into mature oligodendrocytes after
stroke [
69
]. miRs have been shown to play a pivotal role in regulating OPC proliferation and
differentiation under physiologic conditions [
70
]. Recent studies have implicated miRNAs in regulating
stroke-induced oligodendrogenesis. For example, miR-9 and miR-200b are downregulated in ischemic
white matter at two weeks after ischemic injury and regulate stroke-induced oligodendrogenesis
by targeting the transcription factor serum response factor (SRF) [
57
]. Another miRNA, miR-146a,
is upregulated by stroke in the corpus callosum and SVZ of the ischemic hemisphere. Overexpression
of miR-146a in neural progenitor cells
in vitro
significantly increased their differentiation into O4
+
OPCs via inversely regulating its target gene inteleukin-1 receptor-associated kinase 1 (IRAK1).
Furthermore, overexpression of miR-146a in primary OPCs in culture increases their expression
of myelin proteins, whereas downregulation of endogenous miR-146a suppresses the generation of
new myelin proteins [
58
]. Together, these data suggest that a number of specific miRNAs may mediate
stroke-induced oligodendrogenesis.
Genes 2017,8, 89 6 of 10
Recently, there is a growing interest in clinical studies in looking at whether circulating miRNAs
can be used as diagnostic biomarkers in stroke patients. Several studies have demonstrated alterations
in whole blood and plasma miRNA levels in stroke patients (extensively reviewed by Vijayan and
Reddy [
71
]). Large numbers of miRNAs were found to be up or downregulated in acute stroke
cases [
72
74
]. Of particular interest are levels of brain-specific miRNAs such as miR-124 and miR-9 [
75
]
that could potentially aid clinicians in diagnosing stroke especially when magnetic resonance imaging
is not available. However, the physiologic significance of miRNA blood levels relevant to injury
and recovery mechanisms in the brain tissue needs to be investigated. Furthermore, miRNA-target
interactions are cell type dependent [
76
]. The majority of studies have mainly analyzed miRNAs and
their mRNA targets in bulk tissue samples of the brain [
77
,
78
]. The
in vivo
identification and functional
analysis of miRNAs and their target miRNAs in neurons and glia are challenging after stroke.
5. Histone Deacetylases and microRNAs
miRNAs and HDACs may regulate each other in a manner that could critically coordinate their
functions [
79
]. This relationship is not yet fully understood but could represent a pivotal interaction
that influences gene expression and regulates restorative mechanisms during brain repair after stroke.
One study has found that HDAC inhibition with VPA can alter the expression of a large number
miRNAs in the ischemic cortex [
80
]. Interestingly, the predicted targets of these miRNAs include large
networks of genes involved in the development of the vascular and nervous systems.
Other studies have found that miRNAs can regulate the expression of specific HDACs, and
likewise, HDACs can also regulate the expression of specific miRNAs. For example, HDAC4 is
negatively regulated by miR-206 in a model of amyotrophic lateral sclerosis [
81
], and in a Huntington’s
disease model, miR-22 regulates HDAC4 [
82
]. On the other hand, SIRT1, a Class III HDAC, can
regulate miR-134, and a loss of SIRT1 leads to increased miR-134 resulting in impairment in synaptic
plasticity by a decrease in BDNF and CREB [83].
The interaction between HDACs and miRNAs in the ischemic brain during recovery has not been
investigated, and studies to look into its role in brain repair processes are clearly warranted.
6. Conclusions
In this review, we highlight the recent knowledge in the evolving field of epigenetics in stroke
recovery. Epigenetic players such as DNA methylation, HDACs, and miRNAs are potent modulators
of gene regulation, and an accumulating body of evidence suggests that they play a pivotal role
in regulating brain remodeling after stroke. While further studies to elucidate the functions of
specific HDACs, miRNAs, and the interplay between them are warranted, epigenetic-based therapies
carry a tremendous potential to serve as a novel multifaceted approach to enhance recovery among
stroke survivors.
Acknowledgments:
This publication was supported by National Institute of Neurological Disorders and Stroke
(NINDS) of the National Institutes of Health under award number RO1 NS075156 (ZGZ) and RO1 NS088656
(MC). The content is solely the responsibility of the authors and does not the necessarily represent the official
views of the National Institutes of Health.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Lackland, D.T.; Roccella, E.J.; Deutsch, A.F.; Fornage, M.; George, M.G.; Howard, G.; Kissela, B.M.; Kittner, S.J.;
Lichtman, J.H.; Lisabeth, L.D.; et al. Factors influencing the decline in stroke mortality: A statement from the
american heart association/american stroke association. Stroke J. Cereb. Circ.
2014
,45, 315–353. [CrossRef]
[PubMed]
2.
Writing Group Memebers; Mozaffarian, D.; Benjamin, E.J.; Go, A.S.; Arnett, D.K.; Blaha, M.J.; Cushman, M.;
Das, S.R.; de Ferranti, S.; Despres, J.P.; et al. Heart disease and stroke statistics-2016 update: A report from
the american heart association. Circulation 2016,133, e38-360. [CrossRef] [PubMed]
Genes 2017,8, 89 7 of 10
3.
Zhang, Z.G.; Chopp, M. Neurorestorative therapies for stroke: Underlying mechanisms and translation to
the clinic. Lancet Neurol. 2009,8, 491–500. [CrossRef]
4.
Chen, J.; Venkat, P.; Zacharek, A.; Chopp, M. Neurorestorative therapy for stroke. Front. Hum. Neurosci.
2014
,
8, 382. [CrossRef] [PubMed]
5.
Moore, L.D.; Le, T.; Fan, G. DNA methylation and its basic function. Neuropsychopharmacol. Off. Publ. Am.
Coll. Neuropsychopharmacol. 2013,38, 23–38. [CrossRef] [PubMed]
6.
Goto, K.; Numata, M.; Komura, J.I.; Ono, T.; Bestor, T.H.; Kondo, H. Expression of DNA methyltransferase
gene in mature and immature neurons as well as proliferating cells in mice. Differ. Res. Biol. Divers.
1994
,56,
39–44. [CrossRef]
7.
Endres, M.; Meisel, A.; Biniszkiewicz, D.; Namura, S.; Prass, K.; Ruscher, K.; Lipski, A.; Jaenisch, R.;
Moskowitz, M.A.; Dirnagl, U. DNA methyltransferase contributes to delayed ischemic brain injury.
J. Neurosci. Off. J. Soc. Neurosci. 2000,20, 3175–3181.
8.
Endres, M.; Fan, G.; Meisel, A.; Dirnagl, U.; Jaenisch, R. Effects of cerebral ischemia in mice lacking DNA
methyltransferase 1 in post-mitotic neurons. Neuroreport 2001,12, 3763–3766. [CrossRef] [PubMed]
9.
Dock, H.; Theodorsson, A.; Theodorsson, E. DNA methylation inhibitor zebularine confers stroke protection
in ischemic rats. Transl. Stroke Res. 2015,6, 296–300. [CrossRef] [PubMed]
10.
Turunen, M.P.; Aavik, E.; Yla-Herttuala, S. Epigenetics and atherosclerosis. Biochim. Biophys. Acta
2009
,1790,
886–891. [CrossRef] [PubMed]
11.
Baccarelli, A.; Wright, R.; Bollati, V.; Litonjua, A.; Zanobetti, A.; Tarantini, L.; Sparrow, D.; Vokonas, P.;
Schwartz, J. Ischemic heart disease and stroke in relation to blood DNA methylation. Epidemiology
2010
,21,
819–828. [CrossRef] [PubMed]
12.
Jobe, E.M.; McQuate, A.L.; Zhao, X. Crosstalk among epigenetic pathways regulates neurogenesis.
Front. Neurosci. 2012,6, 59. [CrossRef] [PubMed]
13.
Li, X.; Barkho, B.Z.; Luo, Y.; Smrt, R.D.; Santistevan, N.J.; Liu, C.; Kuwabara, T.; Gage, F.H.; Zhao, X.
Epigenetic regulation of the stem cell mitogen fgf-2 by mbd1 in adult neural stem/progenitor cells.
J. Biol. Chem. 2008,283, 27644–27652. [CrossRef] [PubMed]
14.
Smrt, R.D.; Eaves-Egenes, J.; Barkho, B.Z.; Santistevan, N.J.; Zhao, C.; Aimone, J.B.; Gage, F.H.; Zhao, X.
Mecp2 deficiency leads to delayed maturation and altered gene expression in hippocampal neurons.
Neurobiol. Dis. 2007,27, 77–89. [CrossRef] [PubMed]
15.
Jung, B.P.; Zhang, G.; Ho, W.; Francis, J.; Eubanks, J.H. Transient forebrain ischemia alters the mrna expression
of methyl DNA-binding factors in the adult rat hippocampus. Neuroscience 2002,115, 515–524. [CrossRef]
16.
Kornberg, R.D.; Thomas, J.O. Chromatin structure; oligomers of the histones. Science
1974
,184, 865–868.
[CrossRef] [PubMed]
17.
Li, B.; Carey, M.; Workman, J.L. The role of chromatin during transcription. Cell
2007
,128, 707–719. [CrossRef]
[PubMed]
18.
Yang, X.J.; Seto, E. Hats and hdacs: From structure, function and regulation to novel strategies for therapy
and prevention. Oncogene 2007,26, 5310–5318. [CrossRef] [PubMed]
19.
Kim, H.J.; Rowe, M.; Ren, M.; Hong, J.S.; Chen, P.S.; Chuang, D.M. Histone deacetylase inhibitors exhibit
anti-inflammatory and neuroprotective effects in a rat permanent ischemic model of stroke: Multiple
mechanisms of action. J. Pharmacol. Exp. Ther. 2007,321, 892–901. [CrossRef] [PubMed]
20.
Ren, M.; Leng, Y.; Jeong, M.; Leeds, P.R.; Chuang, D.M. Valproic acid reduces brain damage induced by
transient focal cerebral ischemia in rats: Potential roles of histone deacetylase inhibition and heat shock
protein induction. J. Neurochem. 2004,89, 1358–1367. [CrossRef] [PubMed]
21.
Wang, Z.; Leng, Y.; Tsai, L.K.; Leeds, P.; Chuang, D.M. Valproic acid attenuates blood-brain barrier disruption
in a rat model of transient focal cerebral ischemia: The roles of hdac and mmp-9 inhibition. J. Cereb. Blood
Flow Metab. 2011,31, 52–57. [CrossRef] [PubMed]
22.
Kassis, H.; Chopp, M.; Liu, X.S.; Shehadah, A.; Roberts, C.; Zhang, Z.G. Histone deacetylase expression in
white matter oligodendrocytes after stroke. Neurochem. Int. 2014,77, 17–23. [CrossRef] [PubMed]
23.
Kim, H.J.; Leeds, P.; Chuang, D.M. The hdac inhibitor, sodium butyrate, stimulates neurogenesis in the
ischemic brain. J. Neurochem. 2009,110, 1226–1240. [CrossRef] [PubMed]
24.
Kassis, H.; Shehadah, A.; Li, C.; Zhang, Y.; Cui, Y.; Roberts, C.; Sadry, N.; Liu, X.; Chopp, M.; Zhang, Z.G.
Class iia histone deacetylases affect neuronal remodeling and functional outcome after stroke. Neurochem. Int.
2016,96, 24–31. [CrossRef] [PubMed]
Genes 2017,8, 89 8 of 10
25.
Yildirim, F.; Ji, S.; Kronenberg, G.; Barco, A.; Olivares, R.; Benito, E.; Dirnagl, U.; Gertz, K.; Endres, M.;
Harms, C.; et al. Histone acetylation and creb binding protein are required for neuronal resistance against
ischemic injury. PLoS ONE 2014,9, e95465. [CrossRef] [PubMed]
26.
Faraco, G.; Pancani, T.; Formentini, L.; Mascagni, P.; Fossati, G.; Leoni, F.; Moroni, F.; Chiarugi, A.
Pharmacological inhibition of histone deacetylases by suberoylanilide hydroxamic acid specifically alters
gene expression and reduces ischemic injury in the mouse brain. Mol. Pharmacol.
2006
,70, 1876–1884.
[CrossRef] [PubMed]
27.
Qi, X.; Hosoi, T.; Okuma, Y.; Kaneko, M.; Nomura, Y. Sodium 4-phenylbutyrate protects against cerebral
ischemic injury. Mol. Pharmacol. 2004,66, 899–908. [CrossRef] [PubMed]
28.
Liu, X.S.; Chopp, M.; Kassis, H.; Jia, L.F.; Hozeska-Solgot, A.; Zhang, R.L.; Chen, C.; Cui, Y.S.; Zhang, Z.G.
Valproic acid increases white matter repair and neurogenesis after stroke. Neuroscience
2012
,220, 313–321.
[CrossRef] [PubMed]
29.
De Ruijter, A.J.; van Gennip, A.H.; Caron, H.N.; Kemp, S.; van Kuilenburg, A.B. Histone deacetylases (hdacs):
Characterization of the classical hdac family. Biochem. J. 2003,370, 737–749. [CrossRef] [PubMed]
30.
Baltan, S.; Bachleda, A.; Morrison, R.S.; Murphy, S.P. Expression of histone deacetylases in cellular
compartments of the mouse brain and the effects of ischemia. Transl. Stroke Res.
2011
,2, 411–423. [CrossRef]
[PubMed]
31.
Yang, X.J.; Gregoire, S. Class ii histone deacetylases: From sequence to function, regulation, and clinical
implication. Mol. Cell. Biol. 2005,25, 2873–2884. [CrossRef] [PubMed]
32.
Bolger, T.A.; Yao, T.P. Intracellular trafficking of histone deacetylase 4 regulates neuronal cell death. J. Neurosci.
2005,25, 9544–9553. [CrossRef] [PubMed]
33.
Yuan, H.; Denton, K.; Liu, L.; Li, X.J.; Benashski, S.; McCullough, L.; Li, J. Nuclear translocation of histone
deacetylase 4 induces neuronal death in stroke. Neurobiol. Dis. 2016,91, 182–193. [CrossRef] [PubMed]
34.
Majdzadeh, N.; Wang, L.; Morrison, B.E.; Bassel-Duby, R.; Olson, E.N.; D’Mello, S.R. Hdac4 inhibits cell-cycle
progression and protects neurons from cell death. Dev. Neurobiol.
2008
,68, 1076–1092. [CrossRef] [PubMed]
35.
Majdzadeh, N.; Morrison, B.E.; D’Mello, S.R. Class iia hdacs in the regulation of neurodegeneration.
Front. Biosci. 2008,13, 1072–1082. [CrossRef] [PubMed]
36.
Kassis, H.; Shehadah, A.; Chopp, M.; Roberts, C.; Zhang, Z.G. Stroke induces nuclear shuttling of histone
deacetylase 4. Stroke J. Cereb. Circ. 2015,46, 1909–1915. [CrossRef] [PubMed]
37.
Wang, Z.; Leng, Y.; Wang, J.; Liao, H.M.; Bergman, J.; Leeds, P.; Kozikowski, A.; Chuang, D.M. Tubastatin
a, an hdac6 inhibitor, alleviates stroke-induced brain infarction and functional deficits: Potential roles of
alpha-tubulin acetylation and fgf-21 up-regulation. Sci. Rep. 2016,6, 19626. [CrossRef] [PubMed]
38.
International Stroke Genetics Consortium (ISGC); Wellcome Trust Case Control Consortium 2 (WTCCC2);
Bellenguez, C.; Bevan, S.; Gschwendtner, A.; Spencer, C.C.; Burgess, A.I.; Pirinen, M.; Jackson, C.A.;
Traylor, M.; et al. Genome-wide association study identifies a variant in hdac9 associated with large
vessel ischemic stroke. Nat. Genet. 2012,44, 328–333.
39.
Markus, H.S.; Makela, K.M.; Bevan, S.; Raitoharju, E.; Oksala, N.; Bis, J.C.; O’Donnell, C.; Hainsworth, A.;
Lehtimaki, T. Evidence hdac9 genetic variant associated with ischemic stroke increases risk via promoting
carotid atherosclerosis. Stroke J. Cereb. Circ. 2013,44, 1220–1225. [CrossRef] [PubMed]
40.
Wey, H.Y.; Gilbert, T.M.; Zurcher, N.R.; She, A.; Bhanot, A.; Taillon, B.D.; Schroeder, F.A.; Wang, C.;
Haggarty, S.J.; Hooker, J.M. Insights into neuroepigenetics through human histone deacetylase pet imaging.
Sci. Transl. Med. 2016,8, 351ra106. [CrossRef] [PubMed]
41.
ClinicalTrials.gov Registry. Available online: http://www.Clinicaltrials.Gov/ (accessed on 30 January 2017).
42.
Duvic, M.; Talpur, R.; Ni, X.; Zhang, C.; Hazarika, P.; Kelly, C.; Chiao, J.H.; Reilly, J.F.; Ricker, J.L.;
Richon, V.M.; et al. Phase 2 trial of oral vorinostat (suberoylanilide hydroxamic acid, saha) for refractory
cutaneous t-cell lymphoma (ctcl). Blood 2007,109, 31–39. [CrossRef] [PubMed]
43.
VanderMolen, K.M.; McCulloch, W.; Pearce, C.J.; Oberlies, N.H. Romidepsin (istodax, nsc 630176, fr901228,
fk228, depsipeptide): A natural product recently approved for cutaneous t-cell lymphoma. J. Antibiot.
2011
,
64, 525–531. [CrossRef] [PubMed]
44.
West, A.C.; Johnstone, R.W. New and emerging hdac inhibitors for cancer treatment. J. Clin. Investig.
2014
,
124, 30–39. [CrossRef] [PubMed]
45. Garnock-Jones, K.P. Panobinostat: First global approval. Drugs 2015,75, 695–704. [CrossRef] [PubMed]
Genes 2017,8, 89 9 of 10
46.
He, L.; Hannon, G.J. Micrornas: Small rnas with a big role in gene regulation. Nat. Rev. Genet.
2004
,5,
522–531. [CrossRef] [PubMed]
47.
Ha, M.; Kim, V.N. Regulation of microrna biogenesis. Nat. Rev. Mol. Cell Biol.
2014
,15, 509–524. [CrossRef]
[PubMed]
48. Kosik, K.S. The neuronal microrna system. Nat. Rev. Neurosci. 2006,7, 911–920. [CrossRef] [PubMed]
49. Schratt, G. Micrornas at the synapse. Nat. Rev. Neurosci. 2009,10, 842–849. [CrossRef] [PubMed]
50.
Teng, H.; Chopp, M.; Liu, X.; Wang, X.; Chen, C.; Karp, L.; Zhang, Z. Stroke alters mirna expression profiles
in cerebral endothelial cells. In Proceedings of the Annual Society for Neuroscience Meeting, New Orleans,
LA, USA, 13–17 October 2012.
51.
Yin, K.J.; Deng, Z.; Hamblin, M.; Xiang, Y.; Huang, H.; Zhang, J.; Jiang, X.; Wang, Y.; Chen, Y.E. Peroxisome
proliferator-activated receptor delta regulation of mir-15a in ischemia-induced cerebral vascular endothelial
injury. J. Neurosci. Off. J. Soc. Neurosci. 2010,30, 6398–6408. [CrossRef] [PubMed]
52.
Yin, K.J.; Olsen, K.; Hamblin, M.; Zhang, J.; Schwendeman, S.P.; Chen, Y.E. Vascular endothelial cell-specific
microrna-15a inhibits angiogenesis in hindlimb ischemia. J. Biol. Chem.
2012
,287, 27055–27064. [CrossRef]
[PubMed]
53.
Caballero-Garrido, E.; Pena-Philippides, J.C.; Lordkipanidze, T.; Bragin, D.; Yang, Y.; Erhardt, E.B.; Roitbak, T.
In vivo
inhibition of mir-155 promotes recovery after experimental mouse stroke. J. Neurosci. Off. J.
Soc. Neurosci. 2015,35, 12446–12464. [CrossRef] [PubMed]
54.
Li, Y.; Mao, L.; Gao, Y.; Baral, S.; Zhou, Y.; Hu, B. Microrna-107 contributes to post-stroke angiogenesis by
targeting dicer-1. Sci. Rep. 2015,5, 13316. [CrossRef] [PubMed]
55.
Liu, X.S.; Chopp, M.; Zhang, R.L.; Tao, T.; Wang, X.L.; Kassis, H.; Hozeska-Solgot, A.; Zhang, L.; Chen, C.;
Zhang, Z.G. Microrna profiling in subventricular zone after stroke: Mir-124a regulates proliferation of neural
progenitor cells through notch signaling pathway. PLoS ONE 2011,6, e23461. [CrossRef] [PubMed]
56. Liu, X.S.; Chopp, M.; Wang, X.L.; Zhang, L.; Hozeska-Solgot, A.; Tang, T.; Kassis, H.; Zhang, R.L.; Chen, C.;
Xu, J.; et al. Microrna-17-92 cluster mediates the proliferation and survival of neural progenitor cells
after stroke. J. Biol. Chem. 2013,288, 12478–12488. [CrossRef] [PubMed]
57.
Buller, B.; Chopp, M.; Ueno, Y.; Zhang, L.; Zhang, R.L.; Morris, D.; Zhang, Y.; Zhang, Z.G. Regulation of
serum response factor by mirna-200 and mirna-9 modulates oligodendrocyte progenitor cell differentiation.
Glia 2012,60, 1906–1914. [CrossRef] [PubMed]
58.
Liu, X.S.; Chopp, M.; Pan, W.L.; Wang, X.L.; Fan, B.Y.; Zhang, Y.; Kassis, H.; Zhang, R.L.; Zhang, X.M.;
Zhang, Z.G. Microrna-146a promotes oligodendrogenesis in stroke. Mol. Neurobiol.
2017
,54, 227–237.
[CrossRef] [PubMed]
59.
Greenberg, D.A.; Jin, K. From angiogenesis to neuropathology. Nature
2005
,438, 954–959. [CrossRef]
[PubMed]
60.
Caporali, A.; Emanueli, C. Microrna regulation in angiogenesis. Vasc. Pharmacol.
2011
,55, 79–86. [CrossRef]
[PubMed]
61.
Zhang, R.; Zhang, Z.; Wang, L.; Wang, Y.; Gousev, A.; Zhang, L.; Ho, K.L.; Morshead, C.; Chopp, M. Activated
neural stem cells contribute to stroke-induced neurogenesis and neuroblast migration toward the infarct
boundary in adult rats. J. Cereb. Blood Flow Metab. 2004,24, 441–448. [CrossRef] [PubMed]
62.
Zhang, R.L.; Zhang, Z.G.; Chopp, M. Neurogenesis in the adult ischemic brain: Generation, migration,
survival, and restorative therapy. Neuroscientist 2005,11, 408–416. [CrossRef] [PubMed]
63.
Liu, X.S.; Chopp, M.; Zhang, R.L.; Zhang, Z.G. Micrornas in cerebral ischemia-induced neurogenesis.
J. Neuropathol. Exp. Neurol. 2013,72, 718–722. [CrossRef] [PubMed]
64.
Liu, F.J.; Lim, K.Y.; Kaur, P.; Sepramaniam, S.; Armugam, A.; Wong, P.T.; Jeyaseelan, K. Micrornas involved in
regulating spontaneous recovery in embolic stroke model. PLoS ONE
2013
,8, e66393. [CrossRef] [PubMed]
65.
Liu, X.S.; Zhang, Z.G.; Zhang, R.L.; Gregg, S.; Morris, D.C.; Wang, Y.; Chopp, M. Stroke induces gene
profile changes associated with neurogenesis and angiogenesis in adult subventricular zone progenitor cells.
J. Cereb. Blood Flow Metab. 2007,27, 564–574. [CrossRef] [PubMed]
66.
Mendell, J.T. Miriad roles for the mir-17-92 cluster in development and disease. Cell
2008
,133, 217–222.
[CrossRef] [PubMed]
67.
Bian, S.; Hong, J.; Li, Q.; Schebelle, L.; Pollock, A.; Knauss, J.L.; Garg, V.; Sun, T. Microrna cluster mir-17-92
regulates neural stem cell expansion and transition to intermediate progenitors in the developing mouse
neocortex. Cell Rep. 2013,3, 1398–1406. [CrossRef] [PubMed]
Genes 2017,8, 89 10 of 10
68.
Budde, H.; Schmitt, S.; Fitzner, D.; Opitz, L.; Salinas-Riester, G.; Simons, M. Control of oligodendroglial cell
number by the mir-17-92 cluster. Development 2010,137, 2127–2132. [CrossRef] [PubMed]
69.
Zhang, R.; Chopp, M.; Zhang, Z.G. Oligodendrogenesis after cerebral ischemia. Front. Cell. Neurosci.
2013
,7,
201. [CrossRef] [PubMed]
70.
Barca-Mayo, O.; Lu, Q.R. Fine-tuning oligodendrocyte development by micrornas. Front. Neurosci.
2012
,6,
13. [CrossRef] [PubMed]
71.
Vijayan, M.; Reddy, P.H. Peripheral biomarkers of stroke: Focus on circulatory micrornas.
Biochim. Biophys. Acta
2016,1862, 1984–1993. [CrossRef] [PubMed]
72.
Li, W.Y.; Jin, J.; Chen, J.; Guo, Y.; Tang, J.; Tan, S. Circulating micrornas as potential non-invasive biomarkers
for the early detection of hypertension-related stroke. J. Hum. Hypertens. 2014,28, 288–291. [CrossRef] [PubMed]
73.
Tan, K.S.; Armugam, A.; Sepramaniam, S.; Lim, K.Y.; Setyowati, K.D.; Wang, C.W.; Jeyaseelan, K. Expression
profile of micrornas in young stroke patients. PLoS ONE 2009,4, e7689. [CrossRef] [PubMed]
74.
Long, G.; Wang, F.; Li, H.; Yin, Z.; Sandip, C.; Lou, Y.; Wang, Y.; Chen, C.; Wang, D.W. Circulating mir-30a,
mir-126 and let-7b as biomarker for ischemic stroke in humans. BMC Neurol.
2013
,13, 178. [CrossRef]
[PubMed]
75.
Liu, Y.; Zhang, J.; Han, R.; Liu, H.; Sun, D.; Liu, X. Downregulation of serum brain specific microrna is
associated with inflammation and infarct volume in acute ischemic stroke. J. Clin. Neurosci. Off. J. Neurosurg.
Soc. Australas. 2015,22, 291–295. [CrossRef] [PubMed]
76.
Nam, J.W.; Rissland, O.S.; Koppstein, D.; Abreu-Goodger, C.; Jan, C.H.; Agarwal, V.; Yildirim, M.A.;
Rodriguez, A.; Bartel, D.P. Global analyses of the effect of different cellular contexts on microrna targeting.
Mol. Cell 2014,53, 1031–1043. [CrossRef] [PubMed]
77.
Hengst, U.; Jaffrey, S.R. Function and translational regulation of mrna in developing axons. Semin. Cell
Dev. Biol. 2007,18, 209–215. [CrossRef] [PubMed]
78.
McNeill, E.; Van Vactor, D. Micrornas shape the neuronal landscape. Neuron
2012
,75, 363–379. [CrossRef]
[PubMed]
79.
Bourassa, M.W.; Ratan, R.R. The interplay between micrornas and histone deacetylases in neurological
diseases. Neurochem. Int. 2014,77, 33–39. [CrossRef] [PubMed]
80.
Hunsberger, J.G.; Fessler, E.B.; Wang, Z.; Elkahloun, A.G.; Chuang, D.M. Post-insult valproic acid-regulated
micrornas: Potential targets for cerebral ischemia. Am. J. Transl. Res. 2012,4, 316–332. [PubMed]
81.
Williams, A.H.; Valdez, G.; Moresi, V.; Qi, X.; McAnally, J.; Elliott, J.L.; Bassel-Duby, R.; Sanes, J.R.; Olson, E.N.
Microrna-206 delays als progression and promotes regeneration of neuromuscular synapses in mice. Science
2009,326, 1549–1554. [CrossRef] [PubMed]
82.
Jovicic, A.; Zaldivar Jolissaint, J.F.; Moser, R.; Silva Santos Mde, F.; Luthi-Carter, R. Microrna-22 (mir-22)
overexpression is neuroprotective via general anti-apoptotic effects and may also target specific huntington’s
disease-related mechanisms. PLoS ONE 2013,8, e54222. [CrossRef] [PubMed]
83.
Gao, J.; Wang, W.Y.; Mao, Y.W.; Graff, J.; Guan, J.S.; Pan, L.; Mak, G.; Kim, D.; Su, S.C.; Tsai, L.H. A novel
pathway regulates memory and plasticity via sirt1 and mir-134. Nature
2010
,466, 1105–1109. [CrossRef]
[PubMed]
©
2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... Thus, the upregulation of potent gene expression for neuroplasticity and neuroprotection in the brain, including neurotrophic and growth factors, is expected after stroke (Bejot et al., 2011;Guo et al., 2017;Zhang and Liao, 2020). Epigenetic regulation is associated with gene expression for neuroplasticity and neuroprotection during functional recovery after stroke (Schweizer et al., 2013;Felling and Song, 2015;Kassis et al., 2017). Epigenetics is a regulatory mechanism of gene expression that does not alter DNA sequences and is regulated by histone modifications and DNA methylation (Ma et al., 2010). ...
... In particular, the acetylation of histones H3 and H4 is essential for transcriptional activation (Li et al., 2007) because it generally enhances gene expression. It has been recognized that histone acetylation is generally reduced in the brain after ischemic stroke (Kassis et al., 2017). Considering the general reduction in histone acetylation after stroke, researchers have focused on pharmacological treatments using HDAC inhibitors to increase and adjust histone acetylation as a therapeutic intervention (Schweizer et al., 2013). ...
... ICH accounts for 20% of clinical stroke cases, and a large number of patients with ICH are encountered in clinical practice (van Asch et al., 2010;Donkor, 2018). However, contrary to many studies reporting epigenetic alterations in ischemic stroke models (Schweizer et al., 2013;Felling and Song, 2015;Kassis et al., 2017), modifications after ICH are poorly understood. Our previous study using ICH model rats targeting internal capsule (IC) lesions showed that total histone H4 was deacetylated in the motor cortex of the affected hemisphere at the late recovery stage 4 weeks after ICH (Maejima et al., 2023), which is in line with the findigs based on ischemic stroke models (Schweizer et al., 2013;Felling and Song, 2015;Kassis et al., 2017). ...
Article
Epigenetic regulation is involved in post-stroke neuroplasticity. We investigated the effects of intracerebral hemorrhage (ICH) on histone acetylation and gene expression related to neuronal plasticity in the bilateral sensorimotor cortices, which may affect post-stroke sensorimotor function. Wistar rats were randomly divided into the SHAM and ICH groups. We performed ICH surgery stereotaxically based on the microinjection of a collagenase solution in the ICH group. Foot fault and cylinder tests were performed to evaluate motor functions at 4-time points, including pre-ICH surgery. The amount of acetyl histones and the mRNA expression of neurotrophic factors crucial to neuroplasticity in the bilateral sensorimotor cortices were analyzed approximately 2 weeks after ICH surgery. Sensorimotor functions of the ICH group were inferior to those of the SHAM group during 2 weeks post-ICH. ICH increased the acetylation of histone H3 and H4 over the sham level in the ipsilateral and contralateral cortices. ICH increased the mRNA expression of IGF-1, but decreased the expression of BDNF compared with the sham level in the ipsilateral cortex. The present study suggests that histone acetylation levels are enhanced in bilateral sensorimotor cortices after ICH, presenting an altered epigenetic platform for gene expressions related to neuronal plasticity.
... MicroRNAs are 22-nucleotide non-coding RNAs that can downregulate gene products by translational repression when partially complementary sequences are present in the 30-un-translated regions of the target mRNAs or by directing mRNA degradation. MiRNAs are expressed in a tissue-specific manner and are considered to play important roles in cell proliferation, apoptosis and differentiation [69]. Wang and colleagues demonstrated that in an ischemic stroke model, inhibition of miR-148b enhances proliferation and differentiation of NSCs via Wnt signaling pathway which contributes to functional recovery after stroke [70]. ...
Article
Full-text available
A stroke is a complicated neurological illness that occurs when there is a disruption in the blood flow to the brain. This disruption results in the damage of neurons, which then leads to functional abnormalities. The Wnt signalling pathway, which is already well-known for its important function in development and tissue homeostasis, has recently been recognised as a critical factor in the pathophysiology of stroke. Recent studies have shown the Wnt pathway’s roles in stroke-related events. The complex-interactions between the Wnt pathway and stroke emphasising the pathway’s contributions to neuro-protection and synaptic plasticity. The Wnt pathway’s influence on neuro-genesis and synaptic plasticity underscores its potential for driving stroke recovery and rehabilitation strategies. The current review discusses about the Wnt signalling pathway in brain pathophysiology and stroke with special emphasis on the various pathways involved in the positive and negative modulation of Wnt pathway namely Phosphoinositide 3-kinase (PI3-K), Glycogen synthase kinase-3β (GSK-3β), Mitogen-activated protein kinase (MAPK) and nuclear factor erythroid 2-related factor 2 (Nrf2) pathway.
... Epigenetics may provide unique insights to this problem (Qureshi and Mehler, 2010;Kassis et al., 2017;Ng et al., 2018). DNA methylation is one of the most intensively studied epigenetic mechanisms involved in the pathogenesis and prognosis of stroke. ...
Article
Full-text available
Inflammation is closely related to stroke prognosis, and high inflammation status leads to poor functional outcome in stroke. DNA methylation is involved in the pathogenesis and prognosis of stroke. However, the effect of DNA methylation on stroke at high levels of inflammation is unclear. In this study, we constructed a hyperinflammatory cerebral ischemia mouse model and investigated the effect of hypomethylation and hypermethylation on the functional outcome. We constructed a mouse model of transient middle cerebral artery occlusion and treated the mice with lipopolysaccharide to induce a hyperinflammatory state. To investigate the effect of DNA methylation on stroke, we used small molecule inhibitors to restrain the function of key DNA methylation and demethylation enzymes. 2,3,5-Triphenyltetrazolium chloride staining, neurological function scores, neurobehavioral tests, enzyme-linked immunosorbent assay, quantitative reverse transcription PCR and western blot assay were used to evaluate the effects after stroke in mice. We assessed changes in the global methylation status by measuring DNA 5-mc and DNA 5-hmc levels in peripheral blood after the use of the inhibitor. In the group treated with the DNA methylation inhibitor, brain tissue 2,3,5-triphenyltetrazolium chloride staining showed an increase in infarct volume, which was accompanied by a decrease in neurological scores and worsening of neurobehavioral performance. The levels of inflammatory factors interleukin 6 and interleukin-1 beta in ischemic brain tissue and plasma were elevated, indicating increased inflammation. Related inflammatory pathway exploration showed significant overactivation of nuclear factor kappa B. These results suggested that inhibiting DNA methylation led to poor functional outcome in mice with high inflammation following stroke. Further, the effects were reversed by inhibition of DNA demethylation. Our findings suggest that DNA methylation regulates the inflammatory response in stroke and has an important role in the functional outcome of hyperinflammatory stroke.
... In this phase, microglia of M2 phenotype are engaged in repair of the brain region damaged by ischemia via production of anti-inflammatory cytokines (IL-4, IL-10), enhancement of phagocytosis of cellular debris and promoting neuroplasticity. The involvement of epigenetic processes, including DNA methylation, post-translational modifications of histone proteins and microRNAs in post-ischemic brain repair and neuroplasticity has also been postulated [46,47]. ...
Article
Full-text available
Ischemic stroke is one of the major causes of death and permanent disability worldwide. The only efficient treatment to date is anticoagulant therapy and thrombectomy, which enable restitution of blood flow to ischemic tissues. Numerous promising neuroprotectants have failed in clinical trials. Given the complex pathomechanism of stroke, a multitarget pharmacotherapy seems a more rational approach in stroke prevention and treatment than drugs acting on single molecular targets. Recently, vitamin D3 has emerged as a potential treatment adjunct for ischemic stroke, as it interferes with the key prosurvival pathways and shows neuroprotective, anti-inflammatory, regenerative and anti-aging properties in both neuronal and vascular tissue. Moreover, the stimulatory effect of vitamin D3 on brain-derived neurotrophic factor (BDNF) signaling and neuroplasticity may play a role not only in the recovery of neurological functions, but also in ameliorating post-stroke depression and anxiety. This narrative review presents advances in research on the biochemical mechanisms of stroke-related brain damage, and the genomic and non-genomic effects of vitamin D3 which may interfere with diverse cell death signaling pathways. Next, we discuss the results of in vitro and in vivo experimental studies on the neuroprotective potential of 1alpha,25-dihydroxyvitamin D3 (calcitriol) in brain ischemia models. Finally, the outcomes of clinical trials on vitamin D3 efficiency in ischemic stroke patients are briefly reviewed. Despite the mixed results of the clinical trials, it appears that vitamin D3 still holds promise in preventing or ameliorating neurological and psychiatric consequences of ischemic stroke and certainly deserves further study.
... While both genetics and epigenetics play a role in stroke risk factors, the potential for modification of epigenetic changes by lifestyle choices may also allow these alterations to serve as markers for disease progression [12]. The development of pharmaceuticals or cell-based therapies that target epigenetic aberrations may provide the potential to develop therapies that can promote plasticity to improve long-term recovery [14]. ...
Article
Full-text available
Epigenetic changes in stroke may revolutionize cell-based therapies aimed at reducing ischemic stroke risk and damage. Epigenetic changes are a novel therapeutic target due to their specificity and potential for reversal. Possible targets for epigenetic modification include DNA methylation and demethylation, post-translational histone modification, and the actions of non-coding RNAs such as microRNAs. Many of these epigenetic modifications have been reported to modulate atherosclerosis development and progression, ultimately contributing to stroke pathogenesis. Furthermore, epigenetics may play a major role in inflammatory responses following stroke. Stem cells for stroke have demonstrated safety in clinical trials for stroke and show therapeutic benefit in pre-clinical studies. The efficacy of these cell-based interventions may be amplified with adjunctive epigenetic modifications. This review advances the role of epigenetics in atherosclerosis and inflammation in the context of stroke, followed by a discussion on current stem cell studies modulating epigenetics to ameliorate stroke damage.
... Epigenetics is that it can change gene expression or cell phenotype without changing the DNA sequence, mainly including DNA methylation/hydroxymethylation, histone modification, microRNA (miRNAs) post-transcriptional regulation, and non-encoding RNA interference. More and more studies show that epigenetics plays a vital role in nervous system diseases with the continuous innovation of experimental techniques, such as Alzheimer's disease (AD) (Nikolac Perkovic et al. 2021), Parkinson's disease (PD) (Pavlou and Outeiro 2017), depression (Penner-Goeke and Binder 2019), and stroke (Kassis et al. 2017). ...
Article
Full-text available
5-Hydroxymethylcytosine (5hmC) has recently been found that plays an important role in many diseases; however, there are still few studies in the field of stroke. The purpose of this review is to introduce the influence and function of 5hmC in stroke, in order for more people can study it. In this review, we introduced the role of 5hmC in ischemia and hemorrhage stroke, and summarized the possible therapeutic prospects of 5hmC in stroke. In conclusion, we suggest that 5hmC may serve as a biomarker or therapeutic target for the treatment of stroke.
Article
Full-text available
Background Stroke is a major cause of global mortality and disability. Currently, the treatment of acute ischaemic stroke through reperfusion has posed several challenges, raising the need for complementary options to protect the ischaemic penumbra. Recent investigations have indicated that certain epigenetic factors, specifically, histone deacetylases (HDACs) and sirtuins, can be promising for ischaemic stroke therapy, with recent studies suggesting that inhibitors of HDACs or sirtuins may provide neuronal protection after ischaemic stroke. However, the impact of specific HDAC/sirtuin isoforms on the survival of neuronal cells following stroke is still uncertain. This study aims to provide a comprehensive overview of the function of HDACs and their modulators in the treatment of acute ischaemic stroke. Methods This systematic review and meta-analysis will encompass animal intervention studies that explore the efficacy of modulation of HDACs and sirtuins in the acute phase of ischaemic stroke. The review will be reported according to the Preferred Reporting Items for Systematic Reviews and Meta-Analyses guidelines. Electronic searches will be conducted in PubMed, Web of Science and Scopus, with subsequent screening by independent reviewers based on the established eligibility criteria. Methodological quality will be evaluated using the SYRCLE risk of bias tool. The primary outcomes will be infarct volume and functional response, with the secondary outcomes established a priori. Data pertaining to infarct volume will be used for random-effects meta-analysis. Additionally, a descriptive summary will be conducted for the functional response and secondary outcomes. Discussion No systematic review and meta-analysis on the treatment of ischaemic stroke through HDAC modulation has been conducted to date. A comprehensive analysis of the available literature on the relevant preclinical investigations can yield invaluable insights in discerning the most effective trials and in further standardisation of preclinical studies. Systematic review registration This systematic review has been recorded in the International Prospective Register of Systematic Reviews (PROSPERO), with the assigned reference number: CRD42023381420
Article
Objectives: Post-stroke cognitive impairment (PSCI) interferes with neurorehabilitation in patients with stroke. Epigenetic regulation of the hippocampus has been targeted to ameliorate cognitive function. In particular, the acetylation level of histones is modulated by exercise, a potent therapy for patients with stroke. Materials and methods: We examined the effects of exercise and pharmacological inhibition of histone deacetylase (HDAC) using sodium butyrate (NaB) on cognitive function and epigenetic factors in the hippocampus after intracerebral hemorrhage (ICH) to seek beneficial neuronal conditioning against PSCI. Forty rats were randomly assigned to five groups: sham, control, NaB, exercise, and NaB plus exercise groups (n=8 in each group). Except for those in the sham group, all rats underwent stereotaxic ICH surgery with a microinjection of collagenase solution. Intraperitoneal administration of NaB (300 mg/kg) and treadmill exercise (11 m/min for 30 min) were conducted for approximately 4 weeks starting 3 days post-surgery. Results: ICH reduced cognitive function, as detected by the object location test, accompanied by enhanced activity of HDACs. Although exercise did not modulate HDAC activity or cognitive function, repetitive NaB administration increased HDAC activity and ameliorated cognitive impairment induced by ICH. Conclusions: This study suggests that pharmacological treatment with an HDAC inhibitor could potentially present an enriched epigenetic platform in the hippocampus and ameliorate PSCI for neurorehabilitation following ICH.
Article
Epigenetic regulation is expected to provide an enriched platform for neurorehabilitation of post-stroke patients. Acetylation of specific lysine residues in histones is a potent epigenetic target essential for transcriptional regulation. Exercise modulates histone acetylation and gene expression in neuroplasticity in the brain. This study sought to examine the effect of epigenetic treatment using a histone deacetylase (HDAC) inhibitor, sodium butyrate (NaB), and exercise on epigenetic markers in the bilateral motor cortex after intracerebral hemorrhage (ICH) to identify a more enriched neuronal condition for neurorehabilitation. Forty-one male Wistar rats were randomly divided into five groups: sham (n=8), control (n=9), NaB, exercise (n=8), and NaB and exercise (n=8). Intraperitoneal administration of an HDAC inhibitor (300 mg/kg NaB) and treadmill exercise (11 m/min for 30 min) was conducted five days a week for approximately four weeks. ICH specifically decreased the acetylation level of histone H4 in the ipsilateral cortex, and HDAC inhibition with NaB increased the acetylation level of histone H4 over the sham level, accompanied by an improvement in motor function as assessed by the cylinder test. Exercise increased the acetylation levels of histones (H3 and H4) in the bilateral cortex. Synergistic effects of exercise and NaB were not observed during histone acetylation. Pharmacological treatment with a HDAC inhibitor and exercise can provide an enriched epigenetic platform for neurorehabilitation in an individual manner.
Article
Diabetes is the most prevalent metabolic disturbance disease and has been regarded globally as one of the principal causes of mortality. Diabetes is accompanied by several macrovascular complications, including stroke, coronary artery disease (CAD), and cardiomyopathy as a consequence of atherosclerosis. The onset of type 2 diabetes is closely related to insulin resistance (IR). miRNAs have been linked to various metabolic processes, including glucose homeostasis, regulation of lipid metabolism, gluconeogenesis, adipogenesis, glucose transporter type 4 expression, insulin sensitivity, and signaling. Consequently, miRNA dysregulation mediates IR in some target organs, comprising liver, muscle, and adipose tissue. Moreover, miRNAs are crucial in developing diabetes and its associated macrovascular complications through their roles in several signaling pathways implicated in inflammation, apoptosis, cellular survival and migration, the proliferation of vascular smooth muscle cells, neurogenesis, angiogenesis, autophagy, oxidative stress, cardiac remodeling, and fibrosis. Therefore, the purpose of this review is to clarify the role of miRNAs in hepatic, muscle, and adipose tissue IR and explain their roles in the pathogenesis of macrovascular diabetic complications, including stroke, CAD, and cardiomyopathy. Also, explain their roles in gestational diabetes mellitus (GDM). Besides, this review discusses the latest updates on the alteration of miRNA expression in diabetic macrovascular complications.
Article
Full-text available
Histone deacetylase (HDAC) 6 exists exclusively in cytoplasm and deacetylates cytoplasmic proteins such as α-tubulin. HDAC6 dysfunction is associated with several pathological conditions in the central nervous system. This study investigated the beneficial effects of tubastatin A (TubA), a novel specific HDAC6 inhibitor, in a rat model of transient middle cerebral artery occlusion (MCAO) and an in vitro model of excitotoxicity. Post-ischemic TubA treatment robustly improved functional outcomes, reduced brain infarction, and ameliorated neuronal cell death in MCAO rats. These beneficial effects lasted at least three days after MCAO. Notably, when given at 24 hours after MCAO, TubA still exhibited significant protection. Levels of acetylated α-tubulin were decreased in the ischemic hemisphere on Days 1 and 3 after MCAO, and were significantly restored by TubA. MCAO markedly downregulated fibroblast growth factor-21 (FGF-21) and TubA significantly reversed this downregulation. TubA also mitigated impaired FGF-21 signaling in the ischemic hemisphere, including up-regulating β-Klotho, and activating ERK and Akt/GSK-3β signaling pathways. In addition, both TubA and exogenous FGF-21 conferred neuroprotection and restored mitochondrial trafficking in rat cortical neurons against glutamate-induced excitotoxicity. Our findings suggest that the neuroprotective effects of TubA likely involve HDAC6 inhibition and the subsequent up-regulation of acetylated α-tubulin and FGF-21.
Article
Full-text available
Stroke induces new myelinating oligodendrocytes that are involved in ischemic brain repair. Molecular mechanisms that regulate oligodendrogenesis have not been fully investigated. MicroRNAs (miRNAs) are small non-coding RNA molecules that post-transcriptionally regulate gene expression. MiR-146a has been reported to regulate immune response, but the role of miR-146a in oligodendrocyte progenitor cells (OPCs) remains unknown. Adult Wistar rats were subjected to the right middle cerebral artery occlusion (MCAo). In situ hybridization analysis with LNA probes against miR-146a revealed that stroke considerably increased miR-146a density in the corpus callosum and subventricular zone (SVZ) of the lateral ventricle of the ischemic hemisphere. In vitro, overexpression of miR-146a in neural progenitor cells (NPCs) significantly increased their differentiation into O4(+) OPCs. Overexpression of miR-146a in primary OPCs increased their expression of myelin proteins, whereas attenuation of endogenous miR-146a suppressed generation of myelin proteins. MiR-146a also inversely regulated its target gene-IRAK1 expression in OPCs. Attenuation of IRAK1 in OPCs substantially increased myelin proteins and decreased OPC apoptosis. Collectively, our data suggest that miR-146a may mediate stroke-induced oligodendrogenesis.
Article
Full-text available
Each year, the American Heart Association (AHA), in conjunction with the Centers for Disease Control and Prevention, the National Institutes of Health, and other government agencies, brings together the most up-to-date statistics related to heart disease, stroke, and other cardiovascular and metabolic diseases and presents them in its Heart Disease and Stroke Statistical Update. The Statistical Update represents a critical resource for the lay public, policy makers, media professionals, clinicians, healthcare administrators, researchers, and others seeking the best available data on these conditions. Together, cardiovascular disease (CVD) and stroke produce immense health and economic burdens in the United States and globally. The Statistical Update brings together in a single document up-to-date information on the core health behaviors (including diet, physical activity [PA], smoking, and energy balance) and health factors (including blood pressure, cholesterol, and glucose) that define cardiovascular health; a range of …
Article
Epigenetic dysfunction is implicated in many neurological and psychiatric diseases, including Alzheimer's disease and schizophrenia. Consequently, histone deacetylases (HDACs) are being aggressively pursued as therapeutic targets. However, a fundamental knowledge gap exists regarding the expression and distribution of HDACs in healthy individuals for comparison to disease states. Here, we report the first-in-human evaluation of neuroepigenetic regulation in vivo. Using positron emission tomography with [¹¹C]Martinostat, an imaging probe selective for class I HDACs (isoforms 1, 2, and 3), we found that HDAC expression is higher in cortical gray matter than in white matter, with conserved regional distribution patterns within and between healthy individuals. Among gray matter regions, HDAC expression was lowest in the hippocampus and amygdala. Through biochemical profiling of postmortem human brain tissue, we confirmed that [¹¹C]Martinostat selectively binds HDAC isoforms 1, 2, and 3, the HDAC subtypes most implicated in regulating neuroplasticity and cognitive function. In human stem cell-derived neural progenitor cells, pharmacologic-level doses of Martinostat induced changes in genes closely associated with synaptic plasticity, including BDNF (brain-derived neurotrophic factor) and SYP (synaptophysin), as well as genes implicated in neurodegeneration, including GRN (progranulin), at the transcript level, in concert with increased acetylation at both histone H3 lysine 9 and histone H4 lysine 12. This study quantifies HDAC expression in the living human brain and provides the foundation for gaining unprecedented in vivo epigenetic information in health and disease. Copyright
Article
Stroke is the second leading cause of death in the world. Stroke occurs when blood flow stops, and that stoppage results in reduced oxygen supply to neurons in the brain. The occurrence of stroke increases with age, but anyone at any age can suffer from stroke. Recent research has implicated multiple cellular changes in stroke patients, including oxidative stress and mitochondrial dysfunction, inflammatory responses, and changes in mRNA and proteins. Recent research has also revealed that stroke is associated with modifiable and non-modifiable risk factors. Stroke can be controlled by modifiable risk factors, including diet, cardiovascular, hypertension, smoking, diabetes, obesity, metabolic syndrome, depression and traumatic brain injury. Stroke is the major risk factor for vascular dementia (VaD) and Alzheimer’s disease (AD). The purpose of this article is to review the latest developments in research efforts directed at identifying 1) latest developments in identifying biomarkers in peripheral and central nervous system tissues, 2) changes in microRNAs (miRNAs) in patients with stroke, 3) miRNA profile and function in animal brain, and 4) protein biomarkers in ischemic stroke. This article also reviews research investigating circulatory miRNAs as peripheral biomarkers of stroke.
Article
Each year, the American Heart Association (AHA), in conjunction with the Centers for Disease Control and Prevention, the National Institutes of Health, and other government agencies, brings together the most up-to-date statistics related to heart disease, stroke, and other cardiovascular and metabolic diseases and presents them in its Heart Disease and Stroke Statistical Update. The Statistical Update represents a critical resource for the lay public, policy makers, media professionals, clinicians, healthcare administrators, researchers, and others seeking the best available data on these conditions. Together, cardiovascular disease (CVD) and stroke produce immense health and economic burdens in the United States and globally. The Statistical Update brings together in a single document up-to-date information on the core health behaviors and health factors that define cardiovascular health; a range of major clinical disease conditions (including stroke, congenital heart disease, rhythm disorders, subclinical atherosclerosis, coronary heart disease, …
Article
We have previously demonstrated that stroke induces nuclear shuttling of class IIa histone deacetylase 4 (HDAC4). Stroke-induced nuclear shuttling of HDAC4 is positively and significantly correlated with improved indices of neuronal remodeling in the peri-infarct cortex. In this study, using a rat model for middle cerebral artery occlusion (MCAO), we tested the effects of selective inhibition of class IIa HDACs on functional recovery and neuronal remodeling when administered 24hr after stroke. Adult male Wistar rats (n = 15-17/group) were subjected to 2 h MCAO and orally gavaged with MC1568 (a selective class IIa HDAC inhibitor), SAHA (a non-selective HDAC inhibitor), or vehicle-control for 7 days starting 24 h after MCAO. A battery of behavioral tests was performed. Lesion volume measurement and immunohistochemistry were performed 28 days after MCAO. We found that stroke increased total HDAC activity in the ipsilateral hemisphere compared to the contralateral hemisphere. Stroke-increased HDAC activity was significantly decreased by the administration of SAHA as well as by MC1568. However, SAHA significantly improved functional outcome compared to vehicle control, whereas selective class IIa inhibition with MC1568 increased mortality and lesion volume and did not improve functional outcome. In addition, MC1568 decreased microtubule associated protein 2 (MAP2, dendrites), phosphorylated neurofilament heavy chain (pNFH, axons) and myelin basic protein (MBP, myelination) immunoreactivity in the peri-infarct cortex. Quantitative RT-PCR of cortical neurons isolated by laser capture microdissection revealed that MC1568, but not SAHA, downregulated CREB and c-fos expression. Additionally, MC1568 decreased the expression of phosphorylated CREB (active) in neurons. Taken together, these findings demonstrate that selective inhibition of class IIa HDACs impairs neuronal remodeling and neurological outcome. Inactivation of CREB and c-fos by MC1568 likely contributes to this detrimental effect.
Article
Mounting evidence suggests that epigenetic modifications play critical roles in the survival/death of stressed neurons. Chief among these modifications is the deacetylation of histones within the chromatin by histone deacetylases (HDACs). HDAC4 is highly expressed in neurons and is usually trapped in cytosol. However, tightly regulated signal-dependent shuttling of this molecule between cytosol and nucleus occurs. Here, we studied the intracellular trafficking of HDAC4 and regulatory mechanisms during stroke. HDAC4 translocated from the cytosol into the nucleus of neurons in response to stroke induced by middle cerebral artery occlusion (MCAO) in mice. Similar translocation was seen after oxygen-glucose deprivation (OGD) in cultured mouse neurons. Expression of nuclear-restricted HDAC4 increased neuronal death after OGD and worsened infarcts and functional deficits in mice following MCAO; however, expression of cytosolic-restricted HDAC4 did not affect outcome after ischemia. In contrast, HDAC4 knockdown with siRNA improved neuronal survival after OGD. Furthermore, expression of nuclear-restricted HDAC4 reduced the acetylation of histones 3 and 4 as well as the levels of pro-survival downstream molecules after OGD. Finally, genetic deletion of calcium/calmodulin-dependent protein kinase IV (CaMKIV) increased the nuclear accumulation of HDAC4 in MCAO model, while overexpression of CaMKIV reduced the levels of nuclear HDAC4 following OGD. When HDAC4 was inhibited, the neuroprotection provided by CaMKIV overexpression was absent during OGD. Our data demonstrate a detrimental role of the nuclear accumulation of HDAC4 following stroke and identify CaMKIV as a key regulator of neuronal intracellular HDAC4 trafficking during stroke.
Article
Atherosclerotic risk factors can be divided into two main categories-genetical and environmental issues. The latter ones include habitual factors since human habits manifest as environmental factors at the cellular level. Environmental issues can govern human health via epigenetic modification of chromatin structure. This review discusses the recent findings linking general epigenetic mechanisms of chromatin modifications to atherosclerosis development. © 2012 Springer Science+Business Media New York. All rights are reserved.