Sunteți pe pagina 1din 18

Seminars in Cancer Biology 40–41 (2016) 82–99

Contents lists available at ScienceDirect

Seminars in Cancer Biology


journal homepage: www.elsevier.com/locate/semcancer

Review

Dietary phytochemicals as epigenetic modifiers in cancer: Promise


and challenges
Eswar Shankar a,b , Rajnee Kanwal a,b , Mario Candamo c , Sanjay Gupta a,b,d,e,∗
a
Department of Urology, Louis Stokes Cleveland Veterans Affairs Medical Center, Cleveland, OH 44106, USA
b
Department of Urology, Case Western Reserve University, University Hospitals Case Medical Center, Cleveland, OH 44106, USA
c
Department of Biology, School of Undergraduate Studies, Case Western Reserve University, Cleveland, OH 44106, USA
d
Department of Nutrition, Case Western Reserve University, Cleveland, OH 44106, USA
e
Division of General Medical Sciences, Case Comprehensive Cancer Center, Cleveland, OH 44106, USA

a r t i c l e i n f o a b s t r a c t

Article history: The influence of diet and environment on human health has been known since ages. Plant-derived nat-
Received 27 October 2015 ural bioactive compounds (phytochemicals) have acquired an important role in human diet as potent
Received in revised form 8 April 2016 antioxidants and cancer chemopreventive agents. In past few decades, the role of epigenetic alterations
Accepted 18 April 2016
such as DNA methylation, histone modifications and non-coding RNAs in the regulation of mammalian
Available online 23 April 2016
genome have been comprehensively addressed. Although the effects of dietary phytochemicals on gene
expression and signaling pathways have been widely studied in cancer, the impact of these dietary com-
Keywords:
pounds on mammalian epigenome is rapidly emerging. The present review outlines the role of different
Cancer chemoprevention
Dietary agents
epigenetic mechanisms in the regulation and maintenance of mammalian genome and focuses on the role
Epigenetics of dietary phytochemicals as epigenetic modifiers in cancer. Above all, the review focuses on summariz-
DNA methylation ing the progress made thus far in cancer chemoprevention with dietary phytochemicals, the heightened
Histone modification interest and challenges in the future.
microRNA © 2016 Elsevier Ltd. All rights reserved.
Plant polyphenols

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2. Mechanisms of epigenetic modifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.1. DNA methylation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.2. Histone modifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
2.3. Non-coding RNAs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3. Epigenetic mechanisms regulated by dietary phytochemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.1. Apigenin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.2. Curcumin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.3. Green tea polyphenols and epigallocatechin-3-gallate (EGCG) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.4. Genistein and soy isoflavones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.5. Indole-3-carbinol and diindolylmethane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.6. Lycopene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.7. Organosulfur compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.8. Phenethyl isothiocyanate (PEITC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.9. Quercetin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.10. Resveratrol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.11. Sulforaphane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

∗ Corresponding author at: Department of Urology Case, Western Reserve University, 10900 Euclid Avenue, Cleveland, Ohio 44106, USA.
E-mail address: sanjay.gupta@case.edu (S. Gupta).

http://dx.doi.org/10.1016/j.semcancer.2016.04.002
1044-579X/© 2016 Elsevier Ltd. All rights reserved.
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 83

4. Conclusion, limitation and future direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93


Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

1. Introduction methylation is a chemical modification that involves the transfer


of a methyl (CH3 ) moiety from the donor S-adenosyl methionine
The term ‘Epigenetics’ was coined by British developmental biol- (SAM) to the 5 position of cytosine residue that precedes gua-
ogist Conrad Waddington, derived from the combination of word nine in the CG dinucleotide sequence, forming 5-methyl cytosine
‘epigenesis’ and ‘genetics’ [1]. Epigenetics is defined as the branch and S-adenosyl-l-homocysteine (SAH) [17,20–22]. The mammalian
of biology which studies the casual interactions between genes genome has been reported to harbor 3 × 107 methylated cytosine
and their products, where ‘Epi’ means above or over and ‘genetics’ residues mostly within CG dinucleotide sequences [21]. Although
reflects the involvement of genes and heredity. Today epigenetics CG sequences are unevenly distributed throughout the human
may be defined as the study of chemical changes to the genome genome, they are frequently enriched in gene promoters (often
that are heritable and modulate gene expression or cellular phe- referred as CpG islands) and large repetitive sequences such as LINE
notype through mechanisms involving how DNA is packaged and and ALU retrotransposon elements [23]. In mammals, most CpG
expressed without any alterations in the gene nucleotide sequence islands are methylated on cytosine residues by a group of enzymes
itself [1,2]. Epigenetic mechanisms have been shown to be essen- known as DNA methyltransferases (DNMTs), whereas CpG islands
tial in regulating normal cellular functions and play important role within gene promoters tend to be protected from methylation
during developmental stages. In mammals, DNA methylation is [17,23]. In cancer, increased DNA methylation has been observed in
associated with several key processes such as genomic imprinting, the CpG islands in the promoter region of some genes that function
X-chromosome inactivation and tissue-specific gene expression as a ‘switch’ resulting in gene silencing. Although no mutation or
[3–6]. Histone methylation and acetylation patterns were demon- deficiency in any DNMTs has been identified as causally linked to
strated to be closely associated with cognitive functions such as tumor development, most likely because of their critical role during
long term memory formation and storage [7–9]. Epigenetic alter- embryogenesis. So far, there are three major DNA methyltrans-
ations have been implicated in several pathologies such as cancer, ferases (DNMT1, DNMT3a and DNMT3b) identified in mammals.
metabolic syndrome, Alzheimer’s disease and other neurological DNMT1 enzyme has been demonstrated to have a 5–30 fold more
disorders [10–13]. Unlike genetic changes in the genome such preference for hemi-methylated substrates and therefore popu-
as mutation, epigenetic modifications are potentially reversible larly designated as maintenance methyltransferase. It preserves
and can be modified by environmental, dietary and lifestyle fac- the existing methylation patterns in the daughter DNA strands by
tors. Epigenetic mechanisms have been implicated in physiological adding methyl groups to hemi-methylated CpG sequences follow-
responses to intrinsic and extrinsic environmental stimuli such as ing replication. DNMT1 has been demonstrated to be involved in
nutrition, radiation, exposure to chemicals, toxins and hormones de novo methylation activity in embryo lysates and its sequence
[14,15]. The epigenetic diet, or the control of epigenetic modi- specificity was shown to be confined to 5 -CpG-3 dinucleotide
fiers through consumption of dietary phytochemicals, is of extreme sequence with little dependence on sequence context or density
interest. Several studies suggest that dietary phytochemicals are [24]. DNMT3a and DNMT3b enzymes are essential for global de
not only an essential source of nutrients but also important in the novo methylation as they preferentially target unmethylated CpG
elimination of cancer as a life-threatening disease [14,15]. In this sequences [25]. Although DNMT3L, the fourth family member,
review, we focused on the role of dietary phytochemicals as epi- lacks intrinsic DNMT activity by itself, it co-localizes with DNMT3a
genetic modifiers. We discussed in detail different mechanisms of and DNMT3b to establish genomic imprints in maternal germ line
epigenetic modifications in mammals and present a comprehen- and facilitate methylation of retroposons [26–28]. DNMT2, another
sive overview of the current state of knowledge on dietary bioactive enigmatic member of DNMT family, was found to lack any bio-
compounds and their influence on various epigenetic mechanisms chemically detectable DNMT activity. Evidences from phenotypic
highlighting their role as potential anticancer agents. analyses of mice with mutant DNMT genes have provided use-
ful mechanistic insights into the role and establishment of DNA
2. Mechanisms of epigenetic modifications methylation patterns during development [15,21].
Hypermethylation of CpG islands is usually associated with gene
Although a number of epigenetic mechanisms have now been silencing. There are multiple routes through which DNA methyla-
identified, in mammals there are three major epigenetic mech- tion can suppress transcription. A general mechanism is to exclude
anisms which are known to regulate gene expression. These binding of proteins that modulate transcription through their DNA
include DNA methylation, alteration of chromatin structure by binding domains [29]. This mechanism has been demonstrated
post-translational modification of histone or non-histone proteins, to be essential for imprinting of Igf2 gene [30]. CpG methylation
and small non-coding microRNAs (miRNAs) that modulate gene has also been shown to block the binding of several other tran-
expression by either inhibiting translation or causing targeted scription factors, however their biological consequences remain
degradation of specific mRNAs [15,16]. unknown [31]. Another mechanism for DNA methylation medi-
ated gene repression involve binding of specialized DNA binding
2.1. DNA methylation proteins to the methylated CpG stretches, which form repressor
complexes with histone deacetylases (HDACs) and cause chro-
Methylation of cytosine residues within the dinucleotide matin compaction [32–34]. In mammals six methyl-CpG-binding
sequence-CG is one of the most widely studied epigenetic mod- proteins have been characterized till date, which include MeCp2,
ifications in mammals [17]. Forming an essential component of MBD1-4 and Kaiso. Studies demonstrate that all of them (except
the cellular epigenetic machinery, DNA methylation in cooperation mammalian MBD3) possess a domain that specifically targets them
with histone modification regulates gene expression by modu- to methylated CpG regions in vitro and in vivo [33,35,36]. Several
lating DNA packaging and chromatin architecture [18,19]. DNA detailed reviews are available discussing about DNA methylation,
84 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

their role in cancer and development as biomarkers, which is cur- include microRNA (miRNA), small interfering RNA (siRNA), piwi-
rently beyond the scope of this review [37–39]. interacting RNA (piRNA) which accounts for majority of transcripts,
representing approximately 98% of all human transcriptional out-
2.2. Histone modifications put [56,57]. Based on the size, ncRNA can be classified into small
ncRNA which are generally less than 200 nucleotides in length
In addition to DNA methylation, post-translational modifica- and long ncRNA (lncRNA) transcripts that are more than 200
tion of N-terminal histone tails play a significant role in epigenetic nucleotides in length. Small ncRNAs particularly miRNAs regu-
regulation of gene expression [40,41]. A typical nucleosome unit late key epigenetic mechanisms. MiRNAs are known to regulate
consists of ∼146 bp of DNA wrapped around an octamer of his- various components of cellular epigenetic machinery particularly
tones (H2A, H2B, H3 and H4) represents the fundamental building polycomb complexes and thus affect multiple downstream effects
unit of eukaryotic chromatin. A diverse array of covalent chem- [53,58–60]. One such example include miR-214 which downregu-
ical modification of less structured, protruding N-terminal tails of lates EZH2 expression by targeting its 3 -UTR region and accelerates
core histones by methylation, acetylation, ubiquitination, phospho- skeletal muscle differentiation and transcription of developmental
rylation, sumoylation and ADP-ribosylation dictate the dynamics regulators in embryonic stem cells [61]. There are other miRNAs
of chromatin state [42]. Euchromatin is lightly packed form of which have been implicated in the repression of Bmi1, a component
chromatin where DNA is accessible for transcription whereas hete- of PRC1 complex [62–64]. DNA methylation has also been shown
rochromatin represents tightly packed chromatin state inaccessible to be modulated by miRNAs. DNMT1 and 3 have been reported to
to cellular transcriptional machinery. Most of the chemical modi- be targeted by the miR-29 family in lung cancer and leukemia cells
fications occur at lysine (K), arginine (R) and serine (S) residues [65,66]. In summary, recent evidences suggest that ncRNAs have
within the histone tails. These distinct histone modifications on emerged has key regulators of epigenetic mechanisms and also,
one or more histone tails (often referred to as ‘histone code’) may that the modulation of these RNA transcripts by the same epige-
act sequentially or in combination are recognized by other proteins netic processes may lead to major consequences. Several excellent
that signal further downstream events. publications are available on each group of ncRNAs, their mecha-
A number of enzymes have been implicated in catalyzing (addi- nism of action affecting gene regulation, which is beyond the scope
tion or removal) various histone modifications. Examples include of this review [67,68].
histone acetyltransferases (HATs), histone deacetylases (HDACs),
histone methyltransferases (HMTs), histone demethylases (HDMs),
3. Epigenetic mechanisms regulated by dietary
histone kinases etc. In brief, HATs catalyze the addition of acetyl
phytochemicals
group on the ␧-amino group of lysine residues in the N-terminal
tail of histones, which neutralize the positive charge, relax the
Unlike genetic modifications, epigenetic states of genes are
chromatin and facilitate binding of transcriptional machinery to
reversible and can be altered by certain intrinsic and extrin-
the DNA [43]. Till date, 25 HATs have been characterized which
sic factors. This characteristic of epigenetic mechanism may lead
are divided into four families. Examples include GNAT (hGCN5,
to the development of abnormal phenotype as well as regulate
PCAF), MYST (MYST, Tip60), p300/CBP (p300/CBP, SRC) (SRC-1) and
physiological response to some environmental stimuli, diet or
TAFII250 families (TAFII250) [44,45]. In contrast, HDACs catalyze
therapeutic intervention [15]. In past two decades, accumulated
the removal of acetyl groups from K residues resulting in the com-
evidences show that dietary phytochemicals present in abundance
paction of chromatin configuration repressing gene transcription
in fruits, vegetables and beverages which were earlier known
[45]. HDACs are classified into four groups. HDAC-1, −2, −3 and
only for their antioxidant or chemopreventive effects are also
−8 are members of class I HDAC family while HDAC-4, -5, -6, -
potent epigenetic regulators, which could target or revert abnormal
7, -9 and −10 belong to class II HDAC family. HDAC-11 belongs
epigenetic modifications in various human pathologies including
to class IV HDAC group. Sirtuins which require NAD+ as cofac-
cancer [20,69–77]. Here we describe some common dietary phyto-
tor for their activity and are structurally unrelated to other HDAC
chemicals, their structure and molecular weight, which are potent
classes, constitute class III HDAC family [46,47]. HMTs catalyze the
epigenetic modifiers (Table 1). These dietary phytochemicals have
addition of methyl groups to K or R residues while HDMs act to
shown potential as DNMT inhibitors, histone modifiers and vary
remove them [45–47]. Examples of histone lysine methyltrans-
miRNA level altering gene expression and restoring the expres-
ferase include EZH2 (enhancer of zeste homolog 2) and that of
sion of various tumor suppressor genes which are summarized in
histone lysine demethylase include LSD1 (lysine specific demethy-
Tables 2–4 . We also discuss the progress made on some potential
lase 1) [48,49]. Depending on the site of lysine methylation (K4, K9,
dietary phytochemicals in clinical trials (Table 5). Some popular
K27 in histone H3) and methylation status (mono, di or tri methyla-
dietary phytochemicals are discussed in detail below.
tion), histone methylation may have activating or repressive effect
on gene expression [41,49]. H3K4, H3K36 and H3K79 methylation
have activating effects on gene transcription whereas methyla- 3.1. Apigenin
tion of H3K9, H3K27 and H4K20 is generally associated with gene
silencing or transcriptional repression [41,47,50]. Several excellent Apigenin is a dietary plant flavonoid, chemically known as
reviews are available on each group of histone modifying enzymes, 4 ,5,7,-trihydroxyflavone, present in common fruits and vegeta-
their mechanism of action and various histone modifications for bles such as parsley, onions, oranges, tea, chamomile and wheat
biomarker and assay development, which is beyond the scope of sprouts [78]. Previous studies have confirmed that apigenin possess
this review [51,52]. antioxidant, anticancer, anti-inflammatory and anti-mutagenic
properties [79–84]. However, the effect of apigenin on epigenetic-
2.3. Non-coding RNAs related enzymes and their mechanisms was not recognized until
recently. Apigenin treatment (20 and 50 ␮M) was shown to cause
Recent evidences indicate that non-coding RNA (ncRNA) tran- a marked decrease in DNMT activity in vitro [85]. Studies from our
scripts play fundamental role in epigenetic regulation of gene laboratory demonstrated that apigenin-mediated growth arrest
expression and have been implicated in various epigenetic mech- and apoptosis in human prostate cancer PC-3 and 22Rv1 cells was
anisms such as transposon silencing, X-chromosome inactivation, due to inhibition of class I HDACs [86]. In vivo studies using PC-
DNA imprinting, and paramutation [53–55]. In humans, ncRNAs 3 xenografts in athymic nude mice further confirmed that oral
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 85

Table 1
Structure, IUPAC name and molecular weight of some common dietary phytochemicals.

Dietary Agent Structure IUPAC Name Molecular Weight (g/mol)

Apigenin 5,7-dihydroxy-2-(4-hydroxyphenyl)chromen-4-one 270.23

Curcumin (1E,6E)-1,7-bis(4-hydroxy-3-methoxyphenyl)hepta-1,6- 368.37


diene-3,5-dione
Green tea polyphenols

Epicatechin (2R,3R)-2-(3,4-dihydroxyphenyl)-3,4-dihydro-2H- 290.26


chromene-3,5,7-triol

Epigallocatechin (2R,3R)-2-(3,4,5-trihydroxyphenyl)-3,4-dihydro-2H- 306.27


chromene-3,5,7-triol

Epicatechin gallate [(2R,3R)-2-(3,4-dihydroxyphenyl)-5,7-dihydroxy-3,4- 442.37


dihydro-2H-chromen-3-yl]
3,4,5-trihydroxybenzoate

Epigallocatechin gallate [(2R,3R)-5,7-dihydroxy-2-(3,4,5-trihydroxyphenyl)-3,4- 458.37


dihydro-2H-chromen-3-yl]
3,4,5-trihydroxybenzoate

Soy isoflavones

Biochanin A 5,7-dihydroxy-3-(4-methoxyphenyl)chromen-4-one 284.26

Daidezein 7-hydroxy-3-(4-hydroxyphenyl)chromen-4-one 254.23

Genistein 5,7-dihydroxy-3-(4-hydroxyphenyl)chromen-4-one 270.23

Lycopene (6E,8E,10E,12E,14E,16E,18E,20E,22E,24E,26E)- 536.87


2,6,10,14,19,23,27,31-octamethyldotriaconta-
2,6,8,10,12,14,16,18,20,22,24,26,30-tridecaene

Quercetin 2-(3,4-dihydroxyphenyl)-3,5,7-trihydroxychromen-4-one 302.23

Resveratrol 5-[(E)-2-(4-hydroxyphenyl)ethenyl]benzene-1,3-diol 228.24

Sulforaphane 1-isothiocyanato-4-methylsulfinylbutane 177.28

intake of 20 and 50 ␮g/day apigenin over an 8-week period reduces cells. In a recent study by Paredes-Gonzalez et al. [87], apigenin was
tumor burden, HDAC activity and HDAC-1 and HDAC-3 protein lev- shown to reactivate Nrf2 gene which encodes a key transcription
els. HDAC-1 and HDAC-3 mRNA and protein levels were found to factor known to regulate anti-oxidative defense system and skin
be significantly decreased in these cells after 20–40 ␮M apigenin homeostasis, in mouse skin epidermal JB6 P + cells via epigenetic
treatment with consequential increase in global histone H3 and mechanisms. Hypermethylation of 15CpG sites in Nrf2 promoter
H4 acetylation. A corresponding elevation in p21/waf1 and Bax was demonstrated to be reversed by apigenin treatment in a dose-
levels was observed in both in vitro and in vivo studies, which dependent manner. Apigenin exposure also resulted in decrease
resulted in induction of apoptosis and cell cycle arrest in tumor expression of DNMT1, DNMT3A, DNMT3B and HDAC 1–8 levels.
86 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

Table 2
Dietary phytochemicals altering DNA methylation.

Dietary Phytochemical Molecular mechanism(s) Target genes Preclinical Model Disease Type Dose/Concentration Ref.

Apigenin DNMT inhibitor NA Esophageal cells NA 20–50 ␮M [85]


DNMT inhibitor Decreases CpG Skin cancer (mouse Cancer (1.56–50 ␮M) [87]
hypermethylation in Nrf2 skin)
promoter

Curcumin DNMT inhibitor NA Esophageal Cancer 3–50 ␮M [85]


Leukemia

Catechin DNMT inhibitor RARˇ Breast 5–50 ␮M [204]


Epicatechin DNMT inhibitor Breast 50 ␮M [204,113]
Epicatechin-gallate DNMT inhibitor Esophageal 20–50 ␮M [204,113]
Epigalocatechin-3- gallate DNMT inhibitor RARˇ, MGMT, MLH1, Esophageal, Oral, Cancer 20–100 ␮M: [109,205–208]
CDKN2A, RECK,TERT, Prostate,
RXR˛,CDX2, GSTP1, Lung, Colon cancer 0.3–06%
W1F1 cells

Biochanin A DNMT inhibitor NA Esophageal, Cancer 20–100 ␮M [85]


Prostate

Daidezein DNMT inhibitor NA Breast 20–40 ␮M [113,124,126]

Genistein DNMT inhibitor RARˇ, MGMT, CDKN2A, Esophageal 50–300 mg/kg [85,124,126–129,
GSTP1, HMGNS, BTG3 Prostate Tumors 3.75–100 ␮M 131,135,209–211]
RXR˛, CDX2, GSTP1, (Mice)
W1F1
Lycopene Unknown GSTP1, RARˇ, HIN-1 Breast Cancer 2 ␮M [126]
DNMT inhibitor Activation of GSTP1 Prostate 5–40 ␮M
promoter

Quercetin DNMT inhibitor CDKN2A Esophageal Cancer 5–20 ␮M [165]


Breast
Colon

Resveratrol DNMT inhibitor NA Breast, Lungs Cancer 20–40 ␮M [212,213]

Sulforaphane DNMT inhibitor NA Esophageal, Colon Cancer 50 ␮M [212]

NA, information not available.

Furthermore, nuclear localization of Nrf2 was enhanced along with p300/CREB binding protein (CBP), HDAC-1 and HDAC-3 levels after
increase expression of Nrf2 and its target gene NQO1 after apigenin exposure of Raji cells to curcumin. This decrease in HDACs and CBP
treatment. levels were implicated in curcumin-mediated repression of NF-␬B
and Notch1, which resulted in decreased proliferation of Raji cells.
3.2. Curcumin Studies revealed curcumin as a specific inhibitor of p300/CBP HAT,
which emerged as a novel target for cancer treatment [98–100].
Curcumin [1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6- Curcumin treatment caused proteasomal degradation of p300
heptadiene-3,5-dione] is the major active component of popular and other closely related CBP proteins with no effect on HATs
South Asian spice turmeric (Curcuma longa), belonging to Zingiber- such as GCN5 and PCAF [100]. In addition, curcumin-mediated
aceae (ginger) family. The majority of health-benefits associated inhibition of p300/CBP was found to block HDAC inhibitor MS-275
with turmeric intake has been attributed to curcumin which induced histone hyper-acetylation in prostate cancer PC3-M cells
include anti-microbial, antioxidant, anti-inflammatory, anti- and peripheral blood lymphocytes. Curcumin has been shown to
allergic and anti-carcinogenic effects [88–91]. At the molecular epigenetically up-regulate DLEC1 and reduce CpG methylation
level, curcumin has been shown to modulate multiple intracellular in HT29 cells, lowering protein expression of DNMTs and HDACs
pathways associated with proliferation, survival, invasion, apopto- attenuating clonogenicity of human colorectal adenocarcinoma
sis and inflammation [92]. In the context of epigenetic pathways, HT29 cells [101]. In another study, curcumin has been shown to
several studies have reported curcumin to be a potent modulator act as dual agent in causing apoptosis and reversal of promoter
of DNMTs, histone modifying enzymes including HDACs and HATs methylation of p15 gene in ALL leukemia cells [102].
and miRNAs [93]. In silico molecular docking studies of curcumin The chemotherapeutic and chemopreventive effects of cur-
with DNMT1 revealed its ability to block and/or inhibit the cat- cumin has also been closely linked to its ability to modulate miRNAs
alytic thiol group of C1226 binding site in the enzyme resulting in cancer cells. A microarray based study of the effect of curcumin
in decreased DNMT activity [93]. This study was further validated (10 ␮M) on the miRNA profile in pancreatic cancer PxBC-3 cells
by in vitro experimental studies which showed curcumin to be a showed significant changes in the expression of 29 miRNAs (11
potent DNA hypo-methylating agent [94]. Curcumin was reported upregulated and 18 downregulated) after 72 h treatment [103].
to be an effective HDAC inhibitor in HeLa nuclear extracts with an Further studies confirmed that miR-22, which has tumor suppres-
IC50 value of 115 ␮M. Docking studies performed with curcumin sive function, was upregulated after exposure to curcumin and its
binding to HDAC-8 revealed it to be more potent inhibitor of HDAC downstream target genes SP1 and ESR1 were suppressed in these
than known pharmacological inhibitors such as sodium butyrate cells. Ali et al. [104] demonstrated that treatment of pancreatic
and valproic acid [95]. Another study reported that curcumin cancer cells with curcumin and its analogue CDF could induce
treatment to B-NHL Raji cells reduces HDAC-1, HDAC-3 and HDAC- gemcitabine sensitivity via induction of miR-200 and inactivation
8 protein expression in a dose-dependent manner and increases of miR-21 expression. These studies underline the importance of
H4 acetylation in these cells [96]. Consistence with earlier find- curcumin-mediated modulation of miRNAs as an important mech-
ings, studies by Chen et al. [97] reported significant reduction in anism for its biological effects. Curcumin has been reported to
Table 3
Dietary phytochemicals and histone modification.

Dietary Agent Molecular mechanism(s) Target genes In vitro model In vivo model Disease Type Dose/Concentration Ref.

Apigenin HDAC inhibitor p21 PC-3 cells Mouse xenograft Prostate cancer 20–50 mM [86]
(20–50 mg/day)
HDAC inhibitor NA Skin cancer cells (mouse) 1.56–50 mM [87]

Curcumin HAT & HDAC inhibitor H3/H4 deacetylation Cervix, HIV, Hepatoma Plasmodium falciparum; Malaria 0.3–75 mg/kg [96–106]
GATA4, EOMES, GZMB, Leukemia, Prostate Herpes Virus Herpes (6.25–135 mM) [214–222]
PRF1
Brain, Lymphoma Mouse: Rat Epilepsy
Diabetes and Heart failure

Catechin HAT inhibitor NA Lymphocytes NA 100 mM [223]


Epicatechin HAT inhibitor NA Lymphocytes NA 100 mM [223]
Epicatechin-gallate HAT inhibitor NA Lymphocytes, Colon NA 100 mM [223]
Epigalocatechin-3-gallate HAT inhibitor H3/H4 acetylation Lymphocytes, Colon NA Cancer 5–100 mM [223]

E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99


HMT inhibitor H3K27 trimethylation Keratinocytes, Prostate NA
NF-kB, IL-6, BMI-1, EZH2,
SUZ12

Equol HDAC inhibitor H2A/H2B/H3/H4 Drosophila NA 12.8 mM [130]


Genistein HAT inhibitor Acetylation NA Cancer 5–100 mM [124]
H2A/H2B/H3/H4 Esophageal, Prostate, [127–131]
Breast, Renal
HADC inhibitor Acetylation NA
p21, p16, PTEN, p53
FOXA3, SIRT1, hTERT
Daidzein HDAC inhibitor Histone acetylation Esophageal, Prostate NA Cancer 12.8–100 mM [130,210]

Indole-3 carbinol HDAC inhibitor IFN-␥, TNF-␣, IL-6, IL-2, Prostate cancer Sprague-Dawley rats 20–80 mg/rat [82,141,143–150]
Diindolylmethane NF-kB, COX-2, iNOS, IL-1␤,
IL-12
Colon Cancer Nude mice Cancer 15–100 mM
Spleenocytes (CD3+ cells) Nude mice 15–100 mM
100 ␮M
40 mg/kg

Allicin unknown H4 acetylation Erythroleukemia NA 2–200 mM [157–159]


Allyl mercaptan SIRT1 inhibitor H3/H4 acetylation Erythroleukemia Rat: liver Cancer 2–500 mM [157–159,224]
P21/WAF1 Liver, Colon (100 mg/kg)
S-allymercapto-Cysteine HAT inhibitor H3/H4 acetylation Erythroleukemia NA 20–250 mM [156,161]
Diallyl disulphide HDAC inhibitor H3/H4 acetylation Erythroleukemia, Rat Cancer 20–200 mM [156–159,161,224]
P21/WAF1 Leukemia, Liver, Colon, (42–200 mg/kg)
Prostate, Fibroblast

Quercetin SIRTI activator IP-10, MIP-2 Cervix, Drosophila Cancer Inflammat-ory diseases 100 mM [164–175,181,182]
HAT inhibitor Small intestine Mouse
Bowel inflammation

Resveratrol SIRTi activator TNF-˛, IL-8, RBP Cervix, Endothelial Yeast, Drosophila Colon cancer 10–200 mM [177–191]
Embryonic kidney, Mouse Lung cancer
Macrophages, Lungs, Liver, Rats
Cardiomyocytes

Sulforaphane HDAC inhibitor H3/H4 acetylation Prostate, Breast, Colon, Mouse, prostate cancer Cancer 15–25 mM [193–203,225–227]
RAR␤, HBD-2, p21, Bax Kidney xenografts (443 mg/kg, 68 g)

87
Table 4

88
Dietary phytochemicals and changes in miRNA levels.

Dietary Agent Disease Type miRNA Concentration Ref.

Apigenin Lung cancer cells miR138, miR-125a-5p 25–100 ␮M [230]


Neuroblastoma cells miR-138 100 ␮M [229]
Curcumin Bladder cancer cells miR-203 1–20 ␮M [231]
Lung cancer cells miR-320, miR-26a, let-7i, miR-130a, mir-16, miR-125b,miR-23a, miR-27b, miR-155, miR-625, miR-576-3p, miR-186, miR-9 10–50 ␮M [232]
Leukemia miR-15a, miR-16-1 5–20 ␮M [233]
Pancreatic cancer cells miR-103, miR-140, miR-146b, miR-148, miR-15b,miR-181a, miR-181b, miR-181d, miR-195, miR-196a,miR-199a, 10 ␮mol/L [234]
miR-19a,miR-204, miR-20a, miR-21, miR-22miR-23a,b, miR-24, miR-25, miR-26a, miR-27a, miR-34a,miR-374, miR-510, miR-7,
miR-92, miR-93, miR-98

EGCG Neuroblastoma cells miR-92, miR-93,miR 106b, miR-34a, miR-99a 25–100 ␮M [228]
Hepatocellular miR-467b, miR-487b, miR-197, miR-805, miR-374, let-7f, miR-350, miR-24-1, miR-137, miR-335-3p, let-7a, miR-222, miR-26b, 100 ␮M [235]
carcinoma cells miR-30c-1, let-7d, miR-98, miR-30c, miR-30b, miR-32, miR-674, miR-532-5p, let–7 g, miR-18a, miR-192, miR-302d, miR-30b,
miR-802, let-7e,miR-322, miR-720, miR-146b, miR-340-3p, miR-185, miR-425, miR-10a, miR-126-5p, miR-101a, miR-30e, let-7c,
miR-141, miR-33, miR-29a, miR-199b, miR-450a-5p, miR-21, miR-23a, miR-101b, miR-148a, miR-193, miR-23b, miR-107,
miR-140, miR-551b, miR-466c-5p, miR-106a, miR-590-3p, miR-875-3p, miR-224, miR-292-5p, miR-678, miR-469, let-7b,
miR-463miR-574-3p, miR-201, miR-290-3p, miR-181a, miR-302a, miR-429, miR-133a, miR-190b, miR-710, miR-135b,

E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99


miR-296-5p, miR-191, miR-188-5p, miR-298, miR-181a-1, miR–466 g, miR-26b, miR-466f-3p, miR-29b, miR-1224, miR-291b-5p,
miR-324-5p, miR-486, miR-128, miR-450b-3p, miR-135a, miR-294, miR-671-5p, miR-878-3p, miR-801, miR-370, miR-1,
miR-494, miR-133b
Lung carcinoma let-7c, miR210 100 ␮M [235]
Pancreatic cancer cells let-7c 100 ␮M [235]
Colon carcinoma cells miR-27a, miR-20A, miR-17-5P, miR-21 5–30 ␮M [236]

Genistein Ovarian cancer cells miR-100, miR-122a, miR-125b, miR-126, miR-135, miR-135b, miR-136, miR-137, miR-141, miR-152,miR-190, miR-196a, 5 ␮M [136]
miR-196b, miR-204, miR-205, miR-206, miR-217, miR-22, miR-296, miR-30a-3p, miR-30a-5p, miR-331, miR-335, miR-342,
miR-362, miR-449b, miR-454, miR-497, miR-500, miR-501, miR-503, miR-515, miR-517c, miR-532, miR-565, miR-578, miR-584,
miR-585, miR-590, miR-595, miR-625, miR-647, miR-7, miR-765, miR-766
Pancreatic cancer cells miR-200 50 ␮M [137]

Indole-3 carbinol Breast cancer cells miR34a 25 ␮M [137]


Diindolylmethane
Prostate cancer cells miR34a 10 ␮M [238]
miR-21 30–60 ␮M for cells [237]
5 mg/kg (mice)
Mouse lung tumors miR-21, mir-31, miR-130a, miR-146b and miR-377 100–150 ␮M [239]

Lycopene Hepa 1–6 cells (hepatic miR-21 0.05% (fed orally to [240]
cells) mice)

Diallyl disulphide Endothelial progenitor miR-221 0–10 ␮M [241]


cells (10 mg/kg/day)
Breast cancer miR34a 25–400 ␮M [242]

Resveratrol Prostate cancer cells miR-1224-5p, miR-1228, miR-1231, miR-1246, miR-1260, miR-1267, miR-1268, miR-129, miR-1290, miR-1308, miR-1469, 50 ␮M [243]
from lymph node miR-149, miR-150, miR-152, miR-15a, miR-17, miR-1825, miR-185, miR-18b, miR-1908, miR-1915, miR-197, miR-1972,
miR-1973, miR-1974, miR-1975, miR-1977, miR-1979, miR-20a, miR-20b, miR-24, miR-296-5p, miR-483-5p, miR-513a-5p,
miR-548q, miR-572, miR-575, miR-612, miR-638, miR-654-5p, miR-659, miR-671-5p, miR-7, miR-762, miR-764, miR-874,
miR-92b, miR-939
Prostate cancer cells miR-101, miR-106a, miR-106b, miR-1274b, miR-136, miR-141, miR-145, miR-17, miR-182, miR-1826, miR-200b, miR-200c, 5–100 ␮M [244]
miR-20a, miR-20b, miR-21, miR-214, miR-221, miR-222, miR-302c, miR-375, miR-378, miR-720, miR-768-3, miR-93
Colorectal cancer cells miR-1, miR-100-1/2, miR-102, miR-103-1, miR-103-2, miR-146a, miR-146b-5p, miR-16-0, miR-17, miR-181a2, miR-194-2, 50 ␮M [245]
miR-196a1, miR-205, miR-206, miR-21, miR-23a, miR-23b, miR-25, miR-26a, miR-29c, miR-30a-3p, miR-30c-1, miR-30d,
miR-30e-5p, miR-323, miR-340, miR-363-5p, miR-424, miR-494, miR-497, miR-560, miR-560, miR-565, miR-565, miR-572,
miR-574, miR-594, miR-615, miR-622, miR-629, miR-631, miR-638, miR-639, miR-657, miR-659, miR-663, miR-801, miR-92a-2

Luteolin Prostate cancer cells miR-630 0–10 ␮M [246]


Gastric cancer cells miR-34a 5–50 ␮M [247]

Phenethyl Prostate cancer cells miR-141 0–20 ␮M [248]


isothiocyanate
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 89

Table 5
Clinical trials conducted with dietary phytochemicals.

Dietary Agent Molecular Diseaseconditions Dose Endpoints/Outcome Ref.


mechanism

Curcumin DNMT inhibitor Prostate cancer 0.1 g/day Curcumin modulates PSA production in [249]
combination with other isoflavones
DNMT inhibitor Colorectal cancer 0.036-0.18 g/day Pilot study shows safe administration in [250]
patients up to an equivalent dose of 180 mg. It
has low bioavailability in humans and
undergoes intestinal metabolism.

Catechin DNMT inhibitor Prostate cancer NA May decrease risk for advanced prostate [251]
cancer.
Epicatechin DNMT inhibitor 25.5 mg Studies finds low bioavailability and/or [252]
bioaccumulation of green tea polyphenols in
prostate tissue with statistically insignificant
changes in systemic and tissue biomarkers.
Green tea polyphenol activity may be indirect
Epicatechin-gallate DNMT inhibitor 39.8 mg or need evaluation with longer intervention
Epigalocatechin-3- DNMT inhibitor 200 mg durations, repeated dosing, or in patients at
gallate earlier stages of the disease.

Genistein DNMT inhibitor Prostate cancer 30 mg Genistein intake by early prostate cancer [253]
patients modulates several biomarkers and has
an inhibitory effect on androgen related
biomarker.
Decreases MBD1, May induce cell cycle arrest and inhibit [254]
MBD4, MeCP2 proliferation
expression

Daidzein HDAC inhibitor Prostate cancer 0.02 mg Safety and tolerability of isoflavone intake are [254]
proven. The results support the value of
isoflavone treatment for prostate cancer risk
reduction.

Indole-3 carbinol HDAC inhibitor Prostate cancer 225 mg BR-DIM was well tolerated in prostate cancer [255]
patients. Showed variable results on PSA levels.

Lycopene DNMT inhibitor Prostate cancer 20–25 ␮M Prostatic lycopene levels were low after [256]
supplementation in prostate cancer patients
and no change in rate of HGPIN progression to
prostate cancer.

Quercetin DNMT inhibitor Renal cancer 1400 mg/m2 Plasma levels of quercetin inhibited [257]
lymphocyte tyrosine kinase activity, with
evidence of antitumor activity.

Sulforaphane Decreases DNMT Prostate cancer 200 ␮moles/day PSA levels were not affected significantly. [258]
expression

NA, information not available.

reduce DNMT3B activation leading to hypomethylation of RARˇ dose-dependent manner. Furthermore, EGCG treatment (5–50 ␮M
resulting in upregulation of its expression [105]. Dendrasomal cur- for 12–144 h) was found to effectively reactivate methylation-
cumin has been shown to induce DNA hypomethylation in part via silenced genes viz. p16 (INK4a), retinoic acid receptor beta (RARˇ),
up-regulation of epi-miRs, leading to the re-expression of silenced O(6)-methylguanine methyltransferase (MGMT), and human mutL
tumor suppressor genes including MEG3, and suppress oncogenes homologue 1 (hMLH1), in human esophageal cancer KYSE 510 cells.
including HOTAIR, targeting histone modifying complexes in hep- EGCG was also reported to inhibit HDACs and increase permis-
atocellular carcinoma HCC cells [106]. sive or active histone modifications such as histone acetylation
at the target gene promoters. Studies from our laboratory by
Pandey et al. [110] showed that exposure of prostate cancer cells
3.3. Green tea polyphenols and epigallocatechin-3-gallate (EGCG) to green tea polyphenols (GTP) caused re-expression of epige-
netically silenced glutathione-S transferase pi (GSTP1) gene which
EGCG, the major constituent present in green tea, is widely correlated with the promoter demethylation due to DNMT1 inhi-
known for its health-related benefits. Other major constituents bition and histone modification at the promoter region. However,
present in green tea are epicatechin-3-gallate, epigallocatechin, GTP treatment did not show any global hypomethylation effect
and epicatechin. These polyphenols have been shown to possess which could result in genomic instability assessed by the methyla-
multifaceted effect on various cellular targets at the molecular tion status of LINE-1 promoter remained unaffected. GTP treatment
level including epigenetic pathways or enzymes [107,108]. Several decreased mRNA and protein levels of MBD1, MBD4, MeCP2, HDAC
studies have demonstrated that EGCG is a potent demethylat- 1–3 whereas acetylated histone H3 (LysH9/18) and H4 were found
ing agent which inhibits enzymes involved in DNA methylation to be elevated. Another study from our group by Thakur et al. [111],
and is an effective histone modifying agent [109–116]. It is well demonstrated that GTP treatment caused cell cycle arrest and apo-
known that CpG island hypermethylation at the promoter region ptosis by inducing proteasomal degradation of class I HDACs in
leads to epigenetic repression of several critical tumor suppres- human prostate cancer cells. Studies by Li et al. [112] demonstrated
sor genes during tumorigenesis. Fang et al. [109] reported that that EGCG in combination with class I HDAC inhibitor, tricho-
EGCG acts as a competitive inhibitor of DNMT (Ki = 6.89 ␮M), statin A (TSA) could synergistically reactivate ER␣ expression in
which binds to the catalytic pocket and inhibit DNMT activity in a
90 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

ER␣ negative MDA-MB-231 breast cancer cells by modulating his- inhibitor as compared to biochanin A or daidzein. Fang et al. [124]
tone methylation and acetylation pattern at the gene promoter. reported that genistein (2–20 ␮M) could reactivate methylation-
In addition, they also reported that treatment with EGCG and/or silenced genes such as retinoic acid receptor beta (RARˇ), p16INK4a,
TSA contributes to transcriptional activation of ER␣ by causing a and MGMT in esophageal squamous carcinoma cells KYSE 510 and
decreased binding of transcription repressor complex, Rb/p130- prostate cancer LNCaP and PC3 cells. Reactivation of such genes
E2F4/5-HDAC1-SUV39H1-DNMT1 to the regulatory region of the by direct inhibition of DNMT may be one of the factors contribut-
gene. Nandkumar et al. [113] demonstrated that EGCG reactivates ing to the chemopreventive effects of soy isoflavones. In another
silenced tumor suppressor genes Cip1/p21 and p16INK4a by reduc- study, Li et al. [125] demonstrated that genistein treatment in
tion in DNA methylation and increase in histone acetylation in breast MCF10AT benign cells and MCF-7 cancer cells depletes tel-
human skin cancer cells. omerase (hTERT) activity through epigenetic modification which
EGCG has been reported to modulate polycomb proteins such involves genistein-mediated decrease in DNMT1, DNMT3a and
as Bmi-1 and EZH2 [114,115]. EGCG alone or in combination with DNMT3b levels. Site-specific hypomethylation in hTERT promoter
3-Deazaneplanocin A (DZNep), an EZH2 inhibitor was shown to was found to increase the binding of E2F-1 transcription factor.
decrease polycomb group proteins (PcG) proteins including EZH2, Furthermore, genistein was shown to repress hTERT promoter by
EED, SUZ12, MEL18 and Bmi-1 via a mechanism involving pro- chromatin remodeling which involved increase in trimethyl-H3K9
teasome associated degradation. The reduction in PcG proteins enrichment with a concomitant decrease in dimetyl-H3K4 chro-
correlated with a decrease in repressive chromatin marks viz. matin marks. Studies by King-Batoon et al. [126] showed that low,
H3K27me3 and H2AK119ub and HDAC-1 whereas accumulation non-toxic dose of genistein (3.125 ␮M, re-supplemented every 48 h
of acetylated H3 levels was found to be elevated in skin cancer for 1 week) could partially demethylate GSTP1, a tumor suppres-
SCC-13 and human epidermoid carcinoma A431 cells [114,115]. In sor gene, in MCF-7, MDA-MB-468 and MCF10A breast cells. Similar
a recent study form our group, Deb et al. [116] reported that EGCG in vitro studies on other cancer types provide evidence that genis-
and/or GTP treatment to breast cancer cells induced expression of tein is a potent demethylating as well as histone modifying agent,
epigenetically repressed TIMP-3 gene, which is mediated by mod- which could reverse the silenced state of critical tumor suppressor
ulating epigenetic mechanisms involving EZH2 and class I HDACs genes [127–129]. Genistein, daidzein, daidzein metabolite equol
independent of the promoter DNA methylation. After EGCG or GTP and AglyMax have been reported to induce ER␣-mediated his-
treatment, protein levels of class I HDACs and EZH2 were signifi- tone acetylation via modulation of HAT activity [130]. Majid et al.
cantly reduced. Interestingly, transcriptional activation of TIMP-3 [131] demonstrated that genistein-mediated epigenetic induction
was associated with decreased EZH2 localization and H3K27me3 at of p21waf1/Cip1 and p16INK4a tumor suppressor genes in human
the promoter with a concomitant elevation in H3K9/18 acetylation prostate cancer cells involve chromatin modification via upreg-
levels. Time-dependent exposure to EGCG resulted in reactivation ulation of HATs expression. In another study, Basak et al. [132]
of known tumor-suppressor genes in HeLa cells due to marked demonstrated that androgen receptor (AR) down-regulation in
changes in the methylation of the promoter regions of these genes prostate cancer LNCaP cells by genistein was attributed to the inhi-
[117]. In cystic and adenoid cystic carcinoma SACC83 cells, EGCG bition of HDAC6-Hsp90 co-chaperone function, which is required
significantly enhanced the protein and mRNA expression levels for AR protein stabilization. Genistein and daidzein treatment in
of RECK, and markedly reduced the invasive ability of these cells prostate cancer cells caused alterations in the methylation pro-
[118]. Jha et al. [119] demonstrated the dual inhibitory effect of files of 58 genes [133]. Both these agents caused modifications
EGCG in oral squamous carcinoma cells. In this study, EGCG has in the methylation frequencies of MAD1-like 1 (MAD1L1), tumor
been reported to efficaciously reduce DNA methylation catalyzed necrosis factor receptor-associated factor 7 (TRAF7), lysine (K)-
by SssI methyltransferase and human DNMT1 and inhibited DNMT specific demethylase 4B (KDM4B), and human telomerase reverse
activity by increasing the level of SAH in the methylation reaction transcriptase (hTERT) genes. Furthermore, genistein treatment in
catalyzed by catechol-O-methyl transferase (COMT). Additionally, prostate cancer cells increased ER-␤ levels reducing its promoter
EGCG also interacted directly with the catalytic site of human methylation that is required in preventing cell proliferation [134].
DNMT1. Chen et al. [120] performed a genome wide methyla- However, in vivo clinical studies were inconclusive and did not fall
tion and mRNA expression screen in oral squamous carcinoma in line with the studies performed in cell line models. In a ran-
OSCC cells after EGCG treatment that revealed a total of 761 dif- domized trial study involving thirty four healthy pre-menopausal
ferentially methylated gene loci. Comparison of gene expression women to investigate the estrogenic and methylation effect of soy
profiling revealed 184 transcripts with a significant difference and isoflavones (40 mg or 140 mg daily) through one menstrual cycle,
fold change. Gene ontology analysis of differentially methylated Qin et al. [135] reported that considering all subjects or by dose
loci and functional annotation of differentially expressed genes there were no significant changes for any of the cancer related
indicated that the main pathways involved were metabolic, mito- genes (p16INK4a, RASSF1A, RARˇ2, ER, and CCND2). In fact, RARˇ2
gen activated protein kinase (MAPK), Wnt signaling, and cell cycle methylation was shown to increase post-treatment at circulating
pathways suggesting EGCG can affect the methylation status and genistein levels greater than 600 ng/ml unlike the effect observed
gene expression. for genistein levels below 600 ng/ml. Similarly, CCND2 methylation
was found to increase in subjects with post-treatment genistein
3.4. Genistein and soy isoflavones levels above 200 ng/ml as opposed to the effects observed below
this genistein levels. Changes in the methylation of other genes
Genistein is a phyto-estrogen belonging to isoflavone group. p16INK4a, RASSF1A and ER did not correlate with post-treatment
Isoflavones such as genistein, biochanin A and daidzein occur genistein levels.
naturally in plant sources such as soybeans, fava beans, lupine, Genistein and other soy isoflavones have shown to module
kudzu, psoralea etc. [121,122]. Till date, numerous epidemiological miRNAs as well. Parker et al. [136] performed miRNA profiling of
and experimental studies have demonstrated the chemopreventive genistein treated and untreated ovarian cancer UL-3A and UL-3B
effects of genistein and other isoflavones on various cancer types cells and found 53 differentially expressed miRNAs. Upregulation
including breast and prostate cancer [123]. In recent years, the role of miR-200 and let-7 by isoflavones was shown to down-regulate
of genistein and other soy isoflavones as epigenetic modulators ZEB1, slug and vimentin and therefore cause reversal of epithelial
regulating gene expression has been widely reported by several to mesenchymal transition in gemcitabine resistant pancreatic can-
studies. Genistein has been shown to be a more potent DNMT cer cells [137]. In human uveal melanoma cells, genistein treatment
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 91

demonstrated significant growth inhibition by targeting miR-27a carrot, watermelon, papaya, cherries etc. It is a potent anti-oxidant
and its target ZBTB10 [138]. In colorectal cancer cells, genistein has and has been shown to modulate multiple genes involved in
been shown to reduce miR-95, Akt and SGK1 to inhibit cell growth DNA repair, cell cycle control and apoptosis in breast cancer cells
while in breast cancer cells it increased the levels of miR-23b to [152,153]. Studies by King-Batoon et al. [126] reported the effects of
attenuate proliferation [139,140]. lycopene on GSTP1 gene in breast cancer cells. Lycopene (2 ␮M for
1 week) was shown to induce GSTP1 expression and demethylate
3.5. Indole-3-carbinol and diindolylmethane GSTP1 promoter in MDA-MB-468 cell line but not in MCF-7 breast
cancer cells. The expression of other genes such as RARˇ2 and HIN1
Indole-3-carbinol (I3C) is a hydrolysis product derived from a remained unaltered by lycopene treatment in MCF-7 and MDA-
glucosinolate catalyzed by the plant enzyme myrosinase, which MB-468 breast cancer cells. Although this study has shown that
is further metabolized under the acidic conditions in the stom- lycopene may be DNA methylating agent but further studies are
ach to form dimer 3,3 -diindolylmethane (DIM) [141–143]. These needed to decipher the lycopene-mediated effects on epigenetic
phytochemicals are present in cruciferous vegetables particularly mechanisms. The protective effect of lycopene on the prostate is
Brassica genus which include broccoli, cabbage, cauliflower, brussel different between androgen-responsive and androgen-refractory
sprouts, mustard, radish etc. [144]. Both these phytochemicals have derived prostate cancer cells. Fu et al. [154] demonstrated that
been shown to exert anticancer effects by down-regulating multi- lycopene treatment significantly decreased methylation levels of
ple cellular targets and inducing cell cycle arrest and apoptosis in GSTP1 promoter and increased the mRNA and protein levels of
various cancer types [144,145]. A recent study by Li et al. [146] GSTP1 in an androgen-refractory PC-3 cells. It also caused decrease
demonstrated that DIM selectively induce proteasomal degrada- in DNMT3A expression in PC-3 cells while no such change was
tion of class I HDACs (HDAC-1, -2, -3 and -8) with no effect on observed in LNCaP cells.
class II HDACs, both in human colon cancer cells in vitro and in
vivo in tumor xenografts. Furthermore, depletion of class I HDACs 3.7. Organosulfur compounds
was shown to relieve transcription repression of cyclin-dependent
kinase inhibitors p21/waf1 and p27/Kip2, which ultimately resulted Vegetables such as onion, garlic, shallots etc. belonging to
in G2 arrest and DNA damage triggered apoptosis. In another study, allium family possess various water and fat soluble organosul-
Beaver et al. [147] examined the effects of I3C and DIM on prostate fur compounds (OSCs) which have been widely implicated for
cancer LNCaP and PC-3 cells, which differ in their androgen recep- numerous health-related benefits including cancer prevention,
tor status. DIM was shown to inhibit HDAC activity in both cell improved immunity, cardiovascular health and anti–microbial
lines; however I3C demonstrated modest inhibition of HDAC activ- activity [155,156]. Conversion of alliin to allicin and other alkyl
ity in androgen-responsive LNCaP cells with no significant effect alkane-thiosulfinates by the enzyme alliinase generate OSCs, which
on androgen-refractory PC-3 cells. A significant decrease in HDAC- further decompose to more stable sulfur compounds such as dial-
2 protein levels was reported in both cell lines after DIM treatment lyl sulfide (DAS), diallyl disulfide (DADS) and diallyl trisulfide
but the levels of HDAC-1, HDAC-3, HDAC-4, HDAC-6, and HDAC-8 (DATS). S-allylmercaptocysteine (SAMC) is the active metabolite
remained unaltered. Interestingly, a recent study by Busbee et al. derived from DADS, both of which are finally converted to allyl
[148] reported that I3C and DIM may decrease activation, pro- mercaptan (AM) and other metabolites [157]. Accumulated evi-
liferation and cytokine production by staphylococcal enterotoxin dences suggest that in human cancer cells these OSCs increase
B-induced T cells in vitro and in vivo by acting as class I HDAC histone (H3/H4) acetylation, highlighting HDACs as their potential
inhibitors. targets [157–159]. In colon cancer Caco-2 and HT-29 cells, DADS
In a study to examine the effect of DIM on miRNA expression was reported to inhibit cell proliferation and cause G2 cell cycle
profile between gemcitabine-sensitive and gemcitabine-resistant arrest by inhibiting HDAC activity, inducing histone hyperacety-
pancreatic cancer cells, Li et al. [149] reported that DIM treatment lation as well as p21 (Waf1/cip1) expression [160]. Furthermore,
caused significant upregulation of members of miR-200 and let-7 treatment of human colon cancer cells with AM caused accumu-
families. Furthermore, re-expression of miR-200 in gemcitabine- lation of acetylated histones in cellular chromatin and facilitated
resistant cells was found to be associated with a corresponding binding of transcription factor SP3, followed by recruitment of p53
downregulation of mesenchymal markers ZEB1, slug and vimentin at the p21/waf1 promoter. Similarly, a study by Nian et al. [161]
and upregulation of epithelial marker E-cadherin, which lead to the where several garlic derived OSCs were tested or screened in vitro
reversal of EMT phenotype to epithelial morphology. In another for their ability to inhibit HDACs. They reported that AM was the
study, DIM treatment was demonstrated to increase the expres- most potent HDAC inhibitor. Further experiments which include
sion of miR-146a which correlated with reduced expression of molecular modeling, structure-activity and enzyme kinetic studies
EGFR, MTA-2, IRAK-1, and NF-␬B, ultimately resulting in the inhi- confirmed competitive inhibition of HDAC-8 by AM (Ki = 24 ␮M).
bition of pancreatic cancer cell invasion [150]. In human breast
cancer MCF-7 (estrogen receptor positive) and MDA-MB-468 (p53 3.8. Phenethyl isothiocyanate (PEITC)
mutant) cells, DIM treatment was demonstrated to augment G2/M
cell cycle arrest by down-regulating the expression of Cdk-2, Cdk-4 Isothiocyanates are metabolites of glucosinolates which occur
and Cdc25A [151]. In addition DIM treatment was shown to cause naturally in cruciferous vegetables [19]. PEITC is a metabolite
upregulation of miR-21 which negatively regulates Cdc25A result- of gluconasturtin obtained from water cress, which has shown
ing in decreased breast cancer cell proliferation. Similar studies significant beneficial health-related effects particularly in cancer
were reported with I3C which targets the PTEN/AKT signaling path- chemoprevention [162]. Like other natural phytochemicals, PEITC
way and downregulates miR-21. Down-regulation of miR-21 was have been demonstrated to regulate epigenetic mechanisms by
shown to further reduce resistance of HCC cells to I3C through modulating DNA methylation and histone modifications. Studies
inhibition of PTEN/AKT pathway in vivo. by Wang et al. [163] reported dual action of PEITC treatment on
DNA methylation as well as histone modifications in epigenetic
3.6. Lycopene re-activation of GSTP1 gene in prostate cancer cells. Exposure of
prostate cancer LNCaP cells to PEITC inhibited HDAC activity and
Lycopene is the bright red carotene pigment belonging to tetra concurrently demethylate the promoter to restore the GSTP1 lev-
terpenoids and a phytochemical which occur naturally in tomatoes, els to that found in normal prostate cells. In another study, exposure
92 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

Fig. 1. Overview of epigenetic mechanisms regulating gene expression. Dietary phytochemicals impart anticancer properties through alteration in DNA methylation patterns,
histone modification, and changes in non-coding (micro) RNAs levels. Up and down, arrows and  inhibition.

of mouse erythroleukemic cells to allyl isothiocyanates was found the SIRT6 histone deacetylase [170]. Quercetin has been demon-
to increase acetylation of histones, with no significant effect on strated to induce histone acetylation (H3K56ac) in cultured cells
histone deacetylation [158,159]. inhibiting SIRT6 that has a critical role in several types of cancer
[171].
3.9. Quercetin Quercetin has shown to module miRNAs in various human can-
cers as well. Using gastric cancer cells, Du et al. [172] demonstrated
Quercetin is a dietary flavonoid abundantly present in citrus that quercetin in the presence of miR-143 caused autophagy inhi-
fruits, onion and buckwheat. Studies have shown that this bio- bition via targeting GABARAPL1. In another study, Sonoki et al.
flavonoid is a potent free radical scavenger or anti-oxidant as well as [173] demonstrated that quercetin decreased claudin-2 expression
a chemopreventive agent with pro-apoptotic and growth inhibitory and instability of mRNA in lung adenocarcinoma cells mediated
effects on cancer cells [164]. Tan et al. [165] investigated the in vitro by upregulation of miR-16 expression. Treatment of hepatocel-
effects of quercetin on human colon cancer RKO cells and showed lular carcinoma cells with quercetin has been shown to involve
that the hypermethylation status of p16INK4a gene could be com- miR-34a to enhance apoptosis signal through a p53/miR-34a/SIRT1
pletely reversed after quercetin treatment in a dose-dependent signal feedback loop [174]. In another study, modulation of miR-
manner. In another study, quercetin was shown to downregulate 217-KRAS axis in osteosarcoma by quercetin enhanced cisplatin
tumor necrosis factor (TNF)-induced interferon-gamma-inducible sensitivity, regulating the resistance of cancer to chemotherapeutic
protein 10 (IP-10) and macrophage inflammatory protein 2 (MIP-2) drugs [175]. In androgen-refractory prostate cancer cells quercetin
gene expression in murine intestinal epithelial cells. At the molec- treatment caused apoptosis, inhibited invasion and migration capa-
ular level, quercetin treatment was found to cease recruitment of bilities and reduced the expression of numerous prostate tumor
cAMP response element binding protein (CBP)/p300 and histone associated miRNAs, including miR-21. Quercetin increased the
H3 phosphorylation/acetylation at the gene promoters as well as expression of tumor suppressor programmed cell death protein
inhibition of histone acetyltransferase activity [166]. Studies by 4(PDCD4), that negatively regulate miR-21 in these cells [176].
Lee et al. [167] demonstrate that in human leukemic HL-60 cells,
quercetin treatment induced Fas-L related apoptosis and activa- 3.10. Resveratrol
tion of the ERK (extracellular signal–regulated kinase) and JNK
(Jun N-terminus kinase) signaling pathways. Quercetin treatment Resveratrol (trans-3,4 ,5-trihydroxystilbene), a plant derived
was found to increase histone H3 acetylation which in turn caused polyphenolic compound present in grapes, wine and root extracts
increased expression of Fas-L. Furthermore, quercetin treatment of the weed Polygonum cuspidatum, has been identified as a potent
resulted in the activation of HAT and inhibition of HDAC; how- anti-aging, anti-inflammatory and chemopreventive agent affect-
ever HAT activation was found to be only associated with ERK and ing various molecular targets [177,178]. Focusing on epigenetic
JNK pathways. Priyadarsini et al. [168] reported a positive corre- pathways related cellular targets, SIRT1 and acetyl transferase p300
lation between HDAC1 and DNMT1 inhibition by quercetin and its were reported to be activated by resveratrol [179–181]. Resvera-
anti-tumor properties such as pro-apoptosis, anti-angiogenesis and trol was identified as a potent dietary activator of SIRT1, which
anti-invasiveness. In a recent study, Ravichandran et al. [169] pro- lowers the Km (Michaelis constant) for both acetylated substrate
posed a refined pharmacophore model for quercetin binding site of and NAD+. It was reported to stimulate SIRT1-dependent p53
SIRT6 protein. This study was in continuation from a previous study deacetylation which ultimately contributes to increased cell sur-
reporting quercetin to be a potent SIRT6 inhibitor and identified a vival [180,181]. In a study by Wood et al. [182], resveratrol was
preliminary pharmacophore for quercetin binding (consisting of shown to activate sirtuins from metazoans − Caenorhabditis elegans
three hydrogen bond donor and one hydrogen bond acceptor) on and Drosophila melanogaster and delay aging without any effect on
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 93

fecundity. The anti-tumor effect of resveratrol was reported to be induced hyper-acetylation was reported to promote the binding of
mediated partly by SIRT1 [183]. Studies by Wang et al. [184] using repressor proteins such as MAD1 and CTCF to the hTERT regula-
SIRT1 mutant mice demonstrated that disruption of SIRT1 func- tory region. In another study, Myzak et al. [196] reported that SFN
tion p53 null background causes tumor formation and resveratrol metabolites–SFN–cysteine and SFN-N-acetylcysteine were more
mediated activation of SIRT1 reduced tumorigenesis. In addition, potent HDAC inhibitors in vitro as compared to SFN or its glu-
resveratrol was shown to have a negative effect on Survivin gene tathione conjugate. Furthermore, SFN treatment in HCT116 human
expression through histone deacetylation at the gene promoter and colorectal cancer cells increased ␤-catenin-responsive reporter
display a more profound inhibitory effect on BRCA-1 mutant cells (TOPflash) activity in a dose dependent manner and inhibited HDAC
both in vitro and in vivo. Using prostate cancer cells, Kai et al. [185] activity. Consequently, there was an induction in acetylated histone
demonstrated that resveratrol was able to cause downregulation of levels bound to p21/waf1 promoter. In human prostate epithelial
MTA1 (metastasis associated protein), destabilize the NuRD (nucle- BpH −1, LNCaP and PC-3 cells, SFN treatment was shown to inhibit
osome remodeling deacetylase) complex and thus allowing p53 HDAC activity, which was accompanied by an increase in acety-
acetylation. Furthermore, activation of p53 was shown to induce lated histone levels by 50–100% and a corresponding induction of
pro-apoptotic pathways causing apoptosis in prostate cancer cells. p21 and Bax expression which lead to downstream events such
In triple negative breast cancer cells, combination of resveratrol as cell cycle arrest and apoptosis [197]. In human breast cancer
and pterostilbene administered at close to physiologically relevant cells, SFN treatment was shown to inhibit HDAC activity but no
doses caused synergistic (CI <1) growth inhibition by downregu- change in H3 or H4 acetylation was observed [198]. In a recent
lating SIRT1. This combination also decreased the expression of study, Wong et al. [199] demonstrated the genome wide effects
␥-H2AX and telomerase expression affecting DNA damage and of SFN and DIM on promoter methylation in prostate epithelial as
repair [186]. Resveratrol treatment of human bladder cancer EJ well as cancer cells using methyl-DNA immunoprecipitation fol-
cells was found to lead to remarkable S-phase arrest and apoptotic lowed by genome-wide DNA methylation array. The study reported
cell death accompanied by loss of phosphorylation of STAT-3, lead- SFN and DIM treatment induce widespread changes in promoter
ing to downregulation of the STAT-3 pathway as well as decreased methylation patterns and reverse the status of several cancer-
nuclear translocation of SIRT1 and p53 [187]. Resveratrol has also associated aberrantly methylated genes. SFN enhanced the nuclear
shown to have a moderate inhibitory effect on DNMTs. Resvera- translocation of Nrf2 and increased the mRNA and protein levels
trol treatment led to decreased DNMT1 and DNMT3b expression of the Nrf2 target genes in 12-O-tetradecanoylphorbol-13-acetate
in vitro. In a rodent model of estrogen dependent mammary car- (TPA) in mouse skin epidermal JB6 (JB6 P + ) cells [200]. Further-
cinoma, resveratrol treatment decreased DNMT3b expression in more, SFN reduced the expression of DNMTs and downregulated
tumor samples but not in the normal tissues [188]. In another the expression of HDAC-1, HDAC-2, HDAC-3 and HDAC-4, thereby
study, the chemopreventive effects of resveratrol were demon- reactivating Nrf2, a transcription factor for anti-oxidant enzymes.
strated in vivo by pre-exposing aromatic hydrocarbon receptor Another recent study by Mileo and Miccadei [201] highlighted
(AhR) agonist-treated pregnant Sprague-Dawley rats with resvera- the role of SFN and its metabolites sulforaphane-N-acetylcysteine
trol, an AhR-antagonist, which led to lesser CpG methylation of the (SFN-NAC) and sulforaphane-cysteine (SFNCys) generated via the
BRCA-1 gene [189]. mercapturic acid pathway in inducing acetylation of histones and
Studies have demonstrated resveratrol induces apoptosis in a inhibiting HDAC activity in the ileum, colon, and prostate C57BL/6J
time-dependent and dose-dependent manner through increased mice tissues. Another study by Myzak et al. [202] provided first
expression level of miR-16-1 and miR-15a in leukemia cells [190]. evidence for inhibition of in vivo HDAC activity and suppression of
In prostate cancer, study with stilbenes has demonstrated miRNA- tumorigenesis in APC-minus mice. Further studies demonstrated
mediated (miR-17-5p and miR-106a-5p) therapeutic strategy, and that daily SFN dose (7.5 ␮M per animal for 21 days) in the diet
proposed circulating miRNAs as predictive biomarkers for clinical significantly reduced the growth of PC-3 cells in male nude mice,
development [191]. which correlated with a decrease in HDAC activity and increased
global acetylation levels. Furthermore, in human subjects, single
3.11. Sulforaphane dose of 68 g broccoli sprouts caused a significant decrease in HDAC
activity in peripheral blood mononuclear cells within 3–6 h after
Sulforaphane (SFN) is an isothiocyanate abundantly present consumption [203].
in broccoli, water crass, broccoli sprouts, cabbage and kale. SFN
has been shown to modulate various physiological processes and
affect various cellular pathways consistent with its chemopreven- 4. Conclusion, limitation and future direction
tive effects which include induction of cell cycle arrest, apoptosis
and xenobiotic metabolism [192,193]. Microarray based transcrip- The role of dietary phytochemicals as epigenetic modifiers
tome analysis of human colon Caco-2 cells treated with SFN at has been well supported by the studies described herein holding
physiological concentrations showed downregulation of DNMT1 great promise in cancer prevention and/or therapy. In addi-
gene as well as other genes known to be associated with carcino- tion, this review article provides a comprehensive summary on
genesis [194]. Studies by Meeran et al. [195] demonstrated that dietary phytochemicals with potential beneficial effects may in part
in MCF-7 and MDA-MB-231 breast cancer cells, SFN treatment be attributable to their epigenetic properties, including changes
cause dose-dependent and time-dependent inhibition of hTERT in DNA methylation patterns, histone modifications, and in the
(human telomerase reverse transcriptase) via an epigenetic mech- expression of miRNAs (Fig. 1). The fundamental role of epi-
anism involving DNA methylation and histone modifications. SFN genetic mechanisms in the regulation of gene expression and
treatment was shown to cause downregulation of DNMT1 and epigenetic aberrations implicated in various pathologies itself
DNMT3a, which induced site-specific demethylation at hTERT gene underlines the importance of compounds targeting them. It is much
first exon facilitating the binding of CTCF associated with hTERT more challenging than it seems to develop dietary phytochemi-
repression. Furthermore, ChIP analysis of hTERT promoter revealed cals as effective chemopreventive and/or chemotherapeutic agents
that active histone chromatin marks such as acetyl-H3, acetyl- despite their widespread recognition and acceptance as evident
H3K9 and acetyl-H4 were increased whereas repressive chromatin from a large number of research publications [204–258]. There is
marks which include trimethyl-H3K9 and trimethyl-H3K27 were still insufficient preclinical and clinical data published on the epi-
reduced after SFN treatment in a dose-dependent manner. The SFN genetic changes induced by some of the dietary polyphenols, which
94 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

remains a challenge and limitation, of the current data to acquire nutrients or bioactive food components on DNA methylation,
better insight on the molecular mechanisms responsible for these miRNA expression histone methylation or chromatin remodeling
effects. The next important step would be to determine the most complexes are largely unknown. In future, with the emergence
effective doses of these dietary phytochemicals in order to obtain of novel technologies such as next generation sequencing for the
beneficial effects in humans. The doses provided in most of the genome-wide assessment of DNA methylation, localization of the
cell culture studies and in vivo experiments remains pharmacolog- histone marks, and development of bioinformatics tools to sys-
ically irrelevant as such high concentrations of these compounds tematically integrate information can facilitate the analysis of
may not be achieved in humans, and might lack clinical benefit. epigenetic studies with dietary phytochemicals.
As such, pre-clinical and clinical studies addressing the relevance
and validity of in vitro experimental outcomes as well as analysis of
safety profile, dose, bio-availability and routes of administration for Conflict of interest
these compounds are required for designing better chemopreven-
tive/therapeutic strategies along with appropriate designing of the The authors have no competing interest.
clinical trial. Based on profound research interest in the emerging
area of epigenetics with dietary phytochemicals several directions Acknowledgements
in future could be pursued:
The original work from author’s laboratory outlined in this
1. Epigenetic remedy with dietary phytochemicals is an emerging review was supported by VA Merit Review 1I01BX002494 and
area of research that could be effective and valuable approach from the United States Public Health Service Grants RO1CA108512,
in fighting cancer at the level of gene expression. Studies con- R21CA193080 and R03CA186179 to SG.
ducted to date have shown significant effects from many dietary
compounds that target epigenetic pathways in preventing can-
cer and additional mechanistic studies are needed, in particular, References
with untested dietary agents.
2. The exciting field of epigenetics and nutrigenomics, in particu- [1] S. Henikoff, M.A. Matzke, Exploring and explaining epigenetic effects, Trends
Genet. 13 (1997) 293–295.
lar, addresses how nutrients affect the epigenome by targeting [2] V. Hughes, Epigenetics: the sins of the father, Nature 507 (2014) 22–24.
DNA methylation, HDAC activity and miRNAs having chemo- [3] B. Payer, J.T. Lee, X chromosome dosage compensation: how mammals keep
preventive potential. The effect of a combination of dietary the balance, Annu. Rev. Genet. 42 (2008) 733–772.
[4] R. Illingworth, A. Kerr, D. DeSousa, H. Jørgensen, P. Ellis, J. Stalker, et al., A
phytochemicals, their course of action and targets is still at an novel CpG island set identifies tissue-specific methylation at developmental
early stage and could be a promising strategy for epigenetic stud- gene loci, PLoS Biol. 6 (2008) e22.
ies. [5] E. Li, C. Beard, R. Jaenisch, Role for DNA methylation in genomic imprinting,
Nature 366 (1993) 362–365.
3. The phenotype as a result of complex gene-environment inter- [6] P.A. Jones, D. Takai, The role of DNA methylation in mammalian epigenetics,
actions in the current, past and ancestral environment results Science 293 (2001) 1068–1070.
in lifelong remodeling of the epigenome. As some of the effects [7] T.J. Jarome, F.D. Lubin, Histone lysine methylation: critical regulator of
memory and behavior, Rev. Neurosci. 24 (2013) 375–387.
of these dietary phytochemicals appear to be cell type or organ
[8] L. Peixoto, T. Abel, The role of histone acetylation in memory formation and
specific therefore, understanding the mechanism(s) of these dif- cognitive impairments, Neuropsychopharmacology 38 (2013) 62–76.
ferences is a key in designing a personalized regimen for cancer [9] F.D. Lubin, S. Gupta, R.R. Parrish, N.M. Grissom, R.L. Davis, Epigenetic
mechanisms: critical contributors to long-term memory formation,
prevention and/or cure.
Neuroscientist 17 (2011) 616–632.
4. Several studies have demonstrated that disruption of epigenetic [10] M. Gos, Epigenetic mechanisms of gene expression regulation in
mechanisms can alter immune function and contribute to var- neurological diseases, Acta Neurobiol. Exp. (Wars) 73 (2013) 19–37.
ious disease phenotypes. Understanding and manipulating the [11] P.A. Jones, S.B. Baylin, The fundamental role of epigenetic events in cancer,
Nat. Rev. Genet. 3 (2002) 415–428.
epigenome with dietary phytochemicals with a focus on modu- [12] R.M. Stilling, A. Fischer, The role of histone acetylation in age-associated
lating DNA and histone methylation of key inflammatory genes memory impairment and Alzheimer’s disease, Neurobiol. Learn Mem. 96
is an effective approach to target immune-related pathways. (2011) 19–26.
[13] K.D. Bruce, F.R. Cagampang, Epigenetic priming of the metabolic syndrome,
5. Epigenetic defects may eventually lead to genetic defects which Toxicol Mech Methods 21 (2011) 353–361.
are not reversible, appropriate time of interventions using [14] M.A. Suter, K.M. Aagaard-Tillery, Environmental influences on epigenetic
dietary phytochemicals which might be effective in slowing profiles, Semin. Reprod. Med. 27 (2009) 380–390.
[15] R. Jaenisch, A. Bird, Epigenetic regulation of gene expression: how the
and/or reversing cancer progression should be pursued as future genome integrates intrinsic and environmental signals, Nat. Genet. 33
research. (2003) 245–254.
6. It is demonstrated that nanomaterials encapsulation technol- [16] S. Choudhuri, From Waddington’s epigenetic landscape to small noncoding
RNA: some important milestones in the history of epigenetics research,
ogy enhances bioavailability and offers significant promise for
Toxicol. Mech. Methods 21 (2011) 252–274.
chemoprevention. Attempts have been made encapsulating var- [17] J.P. Issa, H.M. Kantarjian, Targeting DNA methylation, Clin. Cancer Res. 15
ious dietary phytochemicals such as curcumin, resveratrol, EGCG (2009) 3938–3946.
[18] P. Dehan, G. Kustermans, S. Guenin, J. Horion, J. Boniver, P. Delvenne, DNA
and genistein in nanoparticles to circumvent the issues of poor
methylation and cancer diagnosis: new methods and applications, Exp. Rev.
bioavailability with dietary agents, and may be considered for Mol. Diagn. 9 (2009) 651–657.
epigenetic studies. [19] A. Link, F. Balaguer, A. Goel, Cancer chemoprevention by dietary
7. Studies have highlighted the existence of cancer stem cells in polyphenols: promising role for epigenetics, Biochem. Pharmacol. 80 (2010)
1771–1792.
most solid and nonsolid tumors. These cells are considered [20] C. Gerhauser, Cancer chemoprevention and nutriepigenetics: state of the art
responsible for tumor relapse and resistance to therapy. It has and future challenges, Top Curr. Chem. 329 (2013) 73–132.
been shown that dietary phytochemicals can impact cancer stem [21] T.H. Bestor, The DNA methyltransferases of mammals, Hum. Mol. Genet. 9
(2000) 2395–2402.
cell self-renewal related pathways. Additional epigenetic stud- [22] D.V. Santi, C.E. Garrett, P.J. Barr, On the mechanism of inhibition of
ies are needed to determine the pathways which are altered by DNA-cytosine methyltransferases by cytosine analogs, Cell 33 (1983) 9–10.
this particular pool of cancer stem cells. [23] A. Bird, DNA methylation patterns and epigenetic memory, Genes Dev. 16
(2002) 6–21.
[24] J.A. Yoder, N.S. Soman, G.L. Verdine, T.H. Bestor, DNA
In conclusion, our knowledge regarding nutritional and diet- (cytosine-5)-methyltransferases in mouse cells and tissues: studies with a
based epigenetics is still limited. In particular, the effects of mechanism-based probe, J. Mol. Biol. 270 (1997) 385–395.
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 95

[25] M. Okano, D.W. Bell, D.A. Haber, E. Li, DNA methyltransferases Dnmt3a and [60] M. Szymanski, M.Z. Barciszewska, V.A. Erdmann, J. Barciszewski, A new
Dnmt3b are essential for de novo methylation and mammalian frontier for molecular medicine: noncoding RNAs, Biochim. Biophys. Acta
development, Cell 99 (1999) 247–257. 1756 (2005) 65–75.
[26] D. Bourc’his, G.L. Xu, C.S. Lin, B. Bollman, T.H. Bestor, Dnmt3L and the [61] G. Deb, V.S. Thakur, S. Gupta, Multifaceted role of EZH2 in breast and
establishment of maternal genomic imprints, Science 294 (2001) prostate tumorigenesis: epigenetics and beyond, Epigenetics 8 (2013)
2536–2539. 464–476.
[27] K. Hata, M. Okano, H. Lei, E. Li, Dnmt3L cooperates with the Dnmt3 family of [62] S. Sander, L. Bullinger, K. Klapproth, K. Fiedler, H.A. Kestler, T.F. Barth, et al.,
de novo DNA methyltransferases to establish maternal imprints in mice, MYC stimulates EZH2 expression by repression of its negative regulator
Development 129 (2002) 1983–1993. miR-26a, Blood 112 (2008) 4202–4212.
[28] H. Denis, M.N. Ndlovu, F. Fuks, Regulation of mammalian DNA [63] S. Varambally, Q. Cao, R.S. Mani, S. Shankar, X. Wang, B. Ateeg, et al.,
methyltransferases: a route to new mechanisms, EMBO Rep. 12 (2011) Genomic loss of microRNA-101 leads to overexpression of histone
647–656. methyltransferase EZH2 in cancer, Science 322 (2008) 1695–1699.
[29] F. Watt, P.L. Molloy, Cytosine methylation prevents binding to DNA of a [64] A.H. Juan, R.M. Kumar, J.G. Marx, R.A. Young, V. Sartorelli,
HeLa cell transcription factor required for optimal expression of the Mir-214-dependent regulation of the polycomb protein Ezh2 in skeletal
adenovirus major late promoter, Genes Dev. 2 (1988) 1136–1143. muscle and embryonic stem cells, Mol. Cell. 36 (2009) 61–74.
[30] A.T. Hark, C.J. Schoenherr, D.J. Katz, R.S. Ingram, J.M. Levorse, S.M. Tilghman, [65] S. Liu, M.T. Tetzlaff, R. Cui, X. Xu, miR-200c inhibits melanoma progression
CTCF mediates methylation-sensitive enhancer-blocking activity at the and drug resistance through down-regulation of BMI-1, Am. J. Pathol. 181
H19/Igf2 locus, Nature 405 (2000) 486–489. (2012) 1823–1835.
[31] P.H. Tate, A.P. Bird, Effects of DNA methylation on DNA-binding proteins and [66] J. Godlewski, M.O. Nowicki, A. Bronisz, S. Williams, A. Otsuki, G. Nuovo,
gene expression, Curr. Opin. Genet Dev. 3 (1993) 226–231. et al., Targeting of the Bmi-1 oncogene/stem cell renewal factor by
[32] P.L. Jones, G.J. Veenstra, P.A. Wade, D. Vermaak, S.U. Kass, N. Landsberger, microRNA-128 inhibits glioma proliferation and self-renewal, Cancer Res.
et al., Methylated DNA and MeCP2 recruit histone deacetylase to repress 68 (2008) 9125–9130.
transcription, Nat. Genet. 19 (1998) 187–191. [67] J. Liz, M. Esteller, lncRNAs and microRNAs with a role in cancer
[33] X. Nan, H.H. Ng, C.A. Johnson, C.D. Laherty, B.M. Turner, R.N. Eisenman, et al., development, Biochim. Biophys. Acta 1859 (2016) 169–176.
Transcriptional repression by the methyl-CpG-binding protein MeCP2 [68] L. Fabris, Y. Ceder, A.M. Chinnaiyan, G.W. Jenster, K.D. Sorensen, S. Tomlins,
involves a histone deacetylase complex, Nature 393 (1998) 386–389. et al., The potential of MicroRNAs as prostate cancer biomarkers, Eur. Urol.
[34] Q. Feng, Y. Zhang, The MeCP1 complex represses transcription through (2016), pii: S0302-2838(16)00012-9.
preferential binding, remodeling, and deacetylating methylated [69] U. Wellner, J. Schubert, U.C. Burk, O. Schmalhofer, F. Zhu, A. Sonntag, et al.,
nucleosomes, Genes Dev. 15 (2001) 827–832. The EMT-activator ZEB1 promotes tumorigenicity by repressing
[35] B. Hendrich, A. Bird, Identification and characterization of a family of stemness-inhibiting microRNAs, Nat. Cell Biol. 11 (2009) 1487–1495.
mammalian methyl-CpG binding proteins, Mol. Cell. Biol. 18 (1998) [70] M. Fabbri, R. Garzon, A. Cimmino, Z. Liu, N. Zanesi, E. Callegari, et al.,
6538–6547. MicroRNA-29 family reverts aberrant methylation in lung cancer by
[36] X. Nan, R.R. Meehan, A. Bird, Dissection of the methyl-CpG binding domain targeting DNA methyltransferases 3A and 3B, Proc. Natl. Acad. Sci. U. S. A.
from the chromosomal protein MeCP2, Nucliic Acids Res. 21 (1993) 104 (2007) 15805–15810.
4886–4892. [71] R. Garzon, S. Liu, M. Fabbri, Z. Liu, C.E. Heaphy, E. Callegari, et al.,
[37] L. Roos, J. Van Dongen, C.G. Bell, A. Burri, P. Deloukas, D.I. Boomsma, et al., MicroRNA-29b induces global DNA hypomethylation and tumor suppressor
Integrative DNA methylome analysis of pan-cancer biomarkers in cancer gene reexpression in acute myeloid leukemia by targeting directly DNMT3A
discordant monozygotic twin-pairs, Clin. Epigenetics. 8 (2016) 7. and 3B and indirectly DNMT1, Blood 113 (2009) 6411–6418.
[38] S. Bhat, S.P. Kabekkodu, A. Noronha, K. Satyamoorthy, Biological [72] J. Huang, C. Plass, C. Gerhauser, Cancer chemoprevention by targeting the
implications and therapeutic significance of DNA methylation regulated epigenome, Curr. Drug Targets. 12 (2011) 1925–1956.
genes in cervical cancer, Biochimie 121 (2016) 298–311. [73] G.L. Russo, V. Vastolo, M. Ciccarelli, L. Albano, P.E. Macchia, P. Ungaro,
[39] C.Y. Yen, H.W. Huang, C.W. Shu, M.F. Hou, S.F. Yuan, H.R. Wang, et al., DNA Dietary polyphenols and chromatin remodelling, Crit. Rev. Food Sci. Nutr. 10
methylation, histone acetylation and methylation of epigenetic (September) (2015) (Epub ahead of print).
modifications as a therapeutic approach for cancers, Cancer Lett. 373 (2016) [74] V.S. Thakur, G. Deb, M.A. Babcook, S. Gupta, Plant phytochemicals as
185–192. epigenetic modulators: role in cancer chemoprevention, AAPS J. 16 (2014)
[40] K. Luger, T.J. Richmond, The histone tails of the nucleosome, Curr. Opin 151–163.
Genet Dev. 8 (1998) 140–146. [75] A.T. Hauser, M. Jung, Targeting epigenetic mechanisms: potential of natural
[41] R.D. Kornberg, Y. Lorch, Twenty-five years of the nucleosome, fundamental products in cancer chemoprevention, Planta Med. 74 (2008) 1593–1601.
particle of the eukaryote chromosome, Cell 98 (1999) 285–294. [76] C.D. Davis, S.A. Ross, Evidence for dietary regulation of microRNA expression
[42] B.D. Strahl, C.D. Allis, The language of covalent histone modifications, Nature in cancer cells, Nutr. Rev. 66 (2008) 477–482.
403 (2000) 41–45. [77] Y. Li, D. Kong, Z. Wang, F.H. Sarkar, Regulation of microRNAs by natural
[43] K. Struhl, Histone acetylation and transcriptional regulatory mechanisms, agents: an emerging field in chemoprevention and chemotherapy research,
Genes Dev. 12 (1998) 599–606. Pharm. Res. 27 (2010) 1027–1041.
[44] A.J. Bannister, T. Kouzarides, Regulation of chromatin by histone [78] D. Patel, S. Shukla, S. Gupta, Apigenin and cancer chemoprevention:
modifications, Cell Res. 21 (2011) 381–395. progress, potential and promise (review), Int. J. Oncol. 30 (2007) 233–245.
[45] M.H. Kuo, C.D. Allis, Roles of histone acetyltransferases and deacetylases in [79] D.F. Birt, S. Hendrich, W. Wang, Dietary agents in cancer prevention:
gene regulation, Bioessays 20 (1998) 615–626. flavonoids and isoflavonoids, Pharmacol. Ther. 90 (2001) 157–177.
[46] A.A. Sauve, C. Wolberger, V.L. Schramm, J.D. Boeke, The biochemistry of [80] Y.J. Surh, Cancer chemoprevention with dietary phytochemicals, Nat. Rev.
sirtuins, Annu. Rev. Biochem. 75 (2006) 435–465. Cancer 3 (2003) 768–780.
[47] D. Mottet, V. Castronovo, Histone deacetylases: target enzymes for cancer [81] C. Manach, A. Scalbert, C. Morand, C. Rémésy, L. Jiménez, Polyphenols: food
therapy, Clin. Exp. Metastasis 25 (2008) 183–189. sources and bioavailability, Am. J. Clin. Nutr. 79 (2004) 727–747.
[48] Y. Shi, Histone lysine demethylases: emerging roles in development, [82] D.F. Birt, B. Walker, M.G. Tibbels, E. Bresnick, Anti-mutagenesis and
physiology and disease, Nat. Rev. Genet. 8 (2007) 829–833. anti-promotion by apigenin, robinetin and indole-3-carbinol,
[49] J.C. Rice, C.D. Allis, Histone methylation versus histone acetylation: new Carcinogenesis 7 (1986) 959–963.
insights into epigenetic regulation, Curr. Opin. Cell Biol. 13 (2001) 263–273. [83] H. Wei, L. Tye, E. Bresnick, D.F. Birt, Inhibitory effect of apigenin, a plant
[50] A.K. Upadhyay, X. Cheng, Dynamics of histone lysine methylation: flavonoid, on epidermal ornithine decarboxylase and skin tumor promotion
structures of methyl writers and erasers, Prog. Drug Res. 67 (2011) 107–124. in mice, Cancer Res. 50 (1990) 499–502.
[51] A.R. Mir, S. Moinuddin Islam, Circulating autoantibodies in cancer patients [84] J.S. Choi, M.N. Islam, M.Y. Ali, E.J. Kim, Y.M. Kim, H.A. Jung, Effects of
have high specificity for glycoxidation modified histone H2A, Clin. Chim. C-glycosylation on anti-diabetic, anti-Alzheimer’s disease and
Acta. 453 (2016) 48–55. anti-inflammatory potential of apigenin, Food Chem. Toxicol. 64 (2014)
[52] F. Ma, C.Y. Zhang, Histone modifying enzymes: novel disease biomarkers 27–33.
and assay development, Expert Rev. Mol. Diagn. 11 (2016) 1–10. [85] M. Fang, D. Chen, C.S. Yang, Dietary polyphenols may affect DNA
[53] G. Deb, A.K. Singh, S. Gupta, EZH2: not EZHY (Easy) to deal, Mol. Cancer Res. methylation, J. Nutr. 137 (2007) 223s–228s.
12 (2014) 639–653. [86] M. Pandey, P. Kaur, S. Shukla, A. Abbas, P. Fu, S. Gupta, Plant flavone apigenin
[54] T. Kouzarides, Chromatin modifications and their function, Cell 128 (2007) inhibits HDAC and remodels chromatin to induce growth arrest and
693–705. apoptosis in human prostate cancer cells: in vitro and in vivo study, Mol.
[55] T. Suzuki, N. Miyata, Epigenetic control using natural products and synthetic Carcinog. 51 (2012) 952–962.
molecules, Curr. Med. Chem. 13 (2006) 935–958. [87] X. Paredes-Gonzalez, F. Fuentes, Z.Y. Su, A.N. Kong, Apigenin reactivates Nrf2
[56] F.F. Costa, Non-coding RNAs, epigenetics and complexity, Gene 410 (2008) anti-oxidative stress signaling in mouse skin epidermal JB6 P+ cells through
9–17. epigenetics modifications, AAPS J. 16 (2014) 727–735.
[57] V.J. Peschansky, C. Wahlestedt, Non-coding RNAs as direct and indirect [88] S. Prasad, S.C. Gupta, A.K. Tyagi, B.B. Aggarwal, Curcumin a component of
modulators of epigenetic regulation, Epigenetics 9 (2014) 3–12. golden spice: from bedside to bench and back, Biotechnol. Adv. 32 (2014)
[58] H. Zhou, H. Hu, M. Lai, Non-coding RNAs and their epigenetic regulatory 1053–1064.
mechanisms, Biol. Cell. 102 (2010) 645–655. [89] G. Bar-Sela, R. Epelbaum, M. Schaffer, Curcumin as an anti-cancer agent:
[59] J.S. Mattick, Non-coding RNAs: the architects of eukaryotic complexity, review of the gap between basic and clinical applications, Curr. Med. Chem.
EMBO Rep. 2 (2001) 986–991. 17 (2010) 190–197.
96 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

[90] R.A. Sharma, A.J. Gescher, W.P. Steward, Curcumin: the story so far, Eur. J. [116] G. Deb, V.S. Thakur, A.M. Limaye, S. Gupta, Epigenetic induction of tissue
Cancer. 41 (2005) 1955–1968. inhibitor of matrix metalloproteinase-3 by green tea polyphenols in breast
[91] M.H. Teiten, S. Eifes, M. Dicato, M. Diederich, Curcumin-the paradigm of a cancer cells, Mol. Carcinog. 54 (2014) 485–499.
multi-target natural compound with applications in cancer prevention and [117] M.A. Khan, A. Hussain, M.K. Sundaram, U. Alalami, D. Gunasekera, L. Ramesh,
treatment, Toxins (Basel) 2 (2010) 128–162. et al., (−)-Epigallocatechin-3 gallate reverses the expression of various
[92] S. Fu, R. Kurzrock, Development of curcumin as an epigenetic agent, Cancer tumor suppressor genes by inhibiting DNA methyltransferases and histone
116 (2010) 4670–4676. deacetylases in human cervical cancer cells, Oncol. Rep. 33 (2015)
[93] J.L. Medina-Franco, F. Lopez-Vallejo, D. Kuck, F. Lyko, Natural products as 1976–1984.
DNA methyltransferase inhibitors: a computer-aided discovery approach, [118] X.Q. Zhou, X.N. Xu, L. Li, J.J. Ma, E.M. Zhen, C.B. Han,
Mol. Divers 15 (2011) 293–304. Epigallocatechin-3-gallate inhibits the invasion of salivary adenoid cystic
[94] Z. Liu, Z. Xie, W. Jones, R.E. Pavlovicz, S. Liu, J. Yu, et al., Curcumin is a potent carcinoma cells by reversing the hypermethylation status of the RECK gene,
DNA hypomethylation agent, Bioorg Med. Chem. Lett. 19 (2009) 706–709. Mol. Med. Rep. 12 (2015) 6031–6036.
[95] G. Bora-Tatar, D. Dayangac-Erden, A.S. Demir, S. Dalkara, K. Yelekçi, H. [119] M. Jha, R. Aggarwal, A.K. Jha, A. Shrivastava, Natural compounds DNA
Erdem-Yurter, Molecular modifications on carboxylic acid derivatives as methyltransferase inhibitors in oral squamous cell carcinoma, Appl.
potent histone deacetylase inhibitors: activity and docking studies, Bioorg Biochem. Biotechnol. 177 (2015) 577–594.
Med. Chem. 17 (2009) 5219–5228. [120] L.L. Chen, W.F. Han, Y. Geng, J.S. Su, A genome-wide study of DNA
[96] H.L. Liu, Y. Chen, G.H. Cui, J.F. Zhou, Curcumin a potent anti-tumor reagent, methylation modified by epigallocatechin-3-gallate in the CAL-27 cell line,
is a novel histone deacetylase inhibitor regulating B-NHL cell line Raji Mol. Med. Rep. 12 (2015) 5886–5890.
proliferation, Acta Pharmacol. Sin. 26 (2005) 603–609. [121] L. Coward, N.C. Barnes, K.D.R. Setchell, S. Barnes, Genistein, daidzein, and
[97] Y. Chen, W. Shu, W. Chen, Q. Wu, H. Liu, G. Cui, Curcumin, both histone their ␤-glycoside conjugates: antitumor isoflavones in soybean foods from
deacetylase and p300/CBP-specific inhibitor, represses the activity of American and Asian diets, J. Agric. Food Chem. 41 (1993) 1961–1967.
nuclear factor kappa B and Notch 1 in Raji cells, Basic Clin. Pharmacol. [122] P.B. Kaufman, J.A. Duke, H. Brielmann, J. Boik, J.E. Hoyt, A comparative
Toxicol. 101 (2007) 427–433. survey of leguminous plants as sources of the isoflavones, genistein and
[98] F.J. Dekker, H.J. Haisma, Histone acetyl transferases as emerging drug daidzein: implications for human nutrition and health, J. Altern.
targets, Drug Discov. Today 14 (2009) 942–948. Complement Med. 3 (1997) 7–12.
[99] M.G. Marcu, Y.J. Jung, S. Lee, E.J. Chung, M.J. Lee, J. Trepel, et al., Curcumin is [123] S. Banerjee, Y. Li, Z. Wang, F.H. Sarkar, Multi-targeted therapy of cancer by
an inhibitor of p300 histone acetylatransferase, Med. Chem. 2 (2006) genistein, Cancer Lett. 269 (2008) 226–242.
169–174. [124] M.Z. Fang, D. Chen, Y. Sun, Z. Jin, J.K. Christman, C.S. Yang, Reversal of
[100] K. Balasubramanyam, R.A. Varier, M. Altaf, V. Swaminathan, N.B. Siddappa, hypermethylation and reactivation of p16INK4a, RARbeta, and MGMT genes
U. Ranga, et al., Curcumin, a novel p300/CREB-binding protein-specific by genistein and other isoflavones from soy, Clin. Cancer Res. 11 (2005)
inhibitor of acetyltransferase, represses the acetylation of 7033–7041.
histone/nonhistone proteins and histone acetyltransferase-dependent [125] Y. Li, L. Liu, L.G. Andrews, T.O. Tollefsbol, Genistein depletes telomerase
chromatin transcription, J. Biol. Chem. 279 (2004) 51163–51171. activity through cross-talk between genetic and epigenetic mechanisms, Int.
[101] Y. Guo, L. Shu, C. Zhang, Z.Y. Su, A.N. Kong, Curcumin inhibits J. Cancer 125 (2009) 286–296.
anchorage-independent growth of HT29 human colon cancer cells by [126] A. King-Batoon, J.M. Leszczynska, C.B. Klein, Modulation of gene methylation
targeting epigenetic restoration of the tumor suppressor gene DLEC1, by genistein or lycopene in breast cancer cells, Environ. Mol. Mutagen 49
Biochem. Pharmacol. 9 (2015) 69–78. (2008) 36–45.
[102] V. Sharma, A.K. Jha, A. Kumar, A. Bhatnagar, G. Narayan, J. Kaur, [127] S. Majid, A.A. Dar, A.E. Ahmad, H. Hirata, K. Kawakami, V. Shahryari, et al.,
Curcumin-Mediated reversal of p15 gene promoter methylation: BTG3 tumor suppressor gene promoter demethylation, histone modification
implication in anti-Neoplastic action against acute lymphoid leukaemia cell and cell cycle arrest by genistein in renal cancer, Carcinogenesis 30 (2009)
line, Folia Biol. (Praha) 61 (2015) 81–89. 662–670.
[103] M. Sun, Z. Estrov, Y. Ji, K.R. Coombes, D.H. Harris, R. Kurzrock, Curcumin [128] S. Majid, A.A. Dar, V. Shahryari, H. Hirata, A. Ahmad, S. Saini, et al., Genistein
(diferuloylmethane) alters the expression profiles of microRNAs in human reverses hypermethylation and induces active histone modifications in
pancreatic cancer cells, Mol. Cancer Ther. 7 (2008) 464–473. tumor suppressor gene B-Cell translocation gene 3 in prostate cancer,
[104] S. Ali, A. Ahmad, S. Banerjee, S. Padhye, K. Dominiak, J.M. Schaffert, et al., Cancer 116 (2010) 66–76.
Gemcitabine sensitivity can be induced in pancreatic cancer cells through [129] J.K. Day, A.M. Bauer, C. DesBordes, Y. Zhuang, B.E. Kim, L.G. Newton, et al.,
modulation of miR-200 and miR-21 expression by curcumin or its analogue Genistein alters methylation patterns in mice, J. Nutr. 132 (2002)
CDF, Cancer Res. 70 (2010) 3606–3617. 2419–2423.
[105] A. Jiang, X. Wang, X. Shan, Y. Li, P. Wang, P. Jiang, Q. Feng, Curcumin [130] T. Hong, T. Nakagawa, W. Pan, M.Y. Kim, W.L. Kraus, T. Ikehara, et al.,
reactivates silenced tumor suppressor gene RAR␤ by reducing DNA Isoflavones stimulate estrogen receptor-mediated core histone acetylation,
methylation, Phytother. Res. 29 (2015) 1237–1245. Biochem. Biophys. Res. Commun. 317 (2004) 259–264.
[106] M. Zamani, M. Sadeghizadeh, M. Behmanesh, F. Najafi, Dendrosomal [131] S. Majid, N. Kikuno, J. Nelles, E. Noonan, Y. Tanaka, K. Kawamoto, et al.,
curcumin increases expression of the long non coding RNA gene MEG3 via Genistein induces the p21WAF1/CIP1 and p16INK4a tumor suppressor
up-regulation of epi-miRs in hepatocellular cancer, Phytomedicine 22 genes in prostate cancer cells by epigenetic mechanisms involving active
(2015) 961–967. chromatin modification, Cancer Res. 68 (2008) 2736–2744.
[107] M.S. Butt, M.T. Sultan, Green tea: nature’s defense against malignancies, Crit. [132] S. Basak, D. Pookot, E.J. Noonan, R. Dahiya, Genistein down-regulates
Rev. Food Sci. Nutr. 49 (2009) 463–473. androgen receptor by modulating HDAC6-Hsp90 chaperone function, Mol.
[108] S. Shankar, S. Ganapathy, R.K. Srivastava, Green tea polyphenols: biology Cancer Ther. 7 (2008) 3195–3202.
and therapeutic implications in cancer, Front. Biosci. 12 (2007) 4881–4899. [133] S. Karsli-Ceppioglu, M. Ngollo, M. Adjakly, A. Dagdemir, G. Judes, A. Lebert,
[109] M.Z. Fang, Y. Wang, N. Ai, Z. Hou, Y. Sun, H. Lu, et al., Tea polyphenol et al., Genome-wide DNA methylation modified by soy phytoestrogens: role
(−)-epigallocatechin-3-gallate inhibits DNA methyltransferase and for epigenetic therapeutics in prostate cancer? OMICS 19 (2015) 209–219.
reactivates methylation-silenced genes in cancer cell lines, Cancer Res. 63 [134] A.M. Mahmoud, U. Al-Alem, M.M. Ali, M.C. Bosland, Genistein increases
(2003) 7563–7570. estrogen receptor beta expression in prostate cancer via reducing its
[110] M. Pandey, S. Shukla, S. Gupta, Promoter demethylation and chromatin promoter methylation, J. Steroid Biochem. Mol. Biol. 152 (2015) 62–75.
remodeling by green tea polyphenols leads to re-expression of GSTP1 in [135] W. Qin, W. Zhu, H. Shi, J.E. Hewett, R.L. Ruhlen, R.S. MacDonald, et al., Soy
human prostate cancer cells, Int. J. Cancer. 126 (2010) 2520–2533. isoflavones have an antiestrogenic effect and alter mammary promoter
[111] V.S. Thakur, K. Gupta, S. Gupta, Green tea polyphenols causes cell cycle hypermethylation in healthy premenopausal women, Nutr. Cancer 61
arrest and apoptosis in prostate cancer cells by suppressing class I histone (2009) 238–244.
deacetylases, Carcinogenesis 33 (2012) 377–784. [136] L.P. Parker, D.D. Taylor, J. Kesterson, D.S. Metzinger, C. Gercel-Taylor,
[112] Y. Li, Y.Y. Yuan, S.M. Meeran, T.O. Tollefsbol, Synergistic epigenetic Modulation of microRNA associated with ovarian cancer cells by genistein,
reactivation of estrogen receptor-alpha (ERalpha) by combined green tea Eur. J. Gynaecol. Oncol. 30 (2009) 616–621.
polyphenol and histone deacetylase inhibitor in ERalpha-negative breast [137] Y. Li, T.G. VandenBoom, D. Kong, Z. Wang, S. Ali, P.A. Philip, et al.,
cancer cells, Mol. Cancer 9 (2010) 274. Up-regulation of miR-200 and let-7 by natural agents leads to the reversal of
[113] V. Nandakumar, M. Vaid, S.K. Katiyar, (−)-Epigallocatechin-3-gallate epithelial-to-mesenchymal transition in gemcitabine-resistant pancreatic
reactivates silenced tumor suppressor genes, Cip1/p21 and p16INK4a, by cancer cells, Cancer Res. 69 (2009) 6704–6712.
reducing DNA methylation and increasing histones acetylation in human [138] Q. Sun, R. Cong, H. Yan, H. Gu, Y. Zeng, N. Liu, et al., Genistein inhibits growth
skin cancer cells, Carcinogenesis 32 (2011) 537–544. of human uveal melanoma cells and affects microRNA-27a and target gene
[114] S. Balasubramanian, G. Adhikary, R.L. Eckert, The Bmi-1 polycomb protein expression, Oncol. Rep. 22 (2009) 563–567.
antagonizes the (−)-epigallocatechin-3-gallate-dependent suppression of [139] J. Qin, J.X. Chen, Z. Zhu, J.A. Teng, Genistein inhibits human colorectal cancer
skin cancer cell survival, Carcinogenesis 31 (2010) 496–503. growth and suppresses miR-95, Akt and SGK1, Cell Physiol. Biochem. 35
[115] S.R. Choudhury, S. Balasubramanian, Y.C. Chew, B. Han, V.E. Marquez, R.L. (2015) 2069–2077.
Eckert, (−)-Epigallocatechin-3-gallate and DZNep reduce polycomb protein [140] C.B. Avci, S.Y. Susluer, H.O. Caglar, T. Balci, D. Aygunes, Y. Dodurga, et al.,
level via a proteasome-dependent mechanism in skin cancer cells, Genistein-induced mir-23b expression inhibits the growth of breast cancer
Carcinogenesis 32 (2011) 1525–1532. cells, Contemp Oncol (Pozn) 19 (2015) 32–35.
[141] E.G. Rogan, The natural chemopreventive compound indole-3-carbinol:
state of the science, In Vivo 20 (2006) 221–228.
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 97

[142] D.M. Minich, J.S. Bland, A review of the clinical efficacy and safety of [169] S. Ravichandran, N. Singh, D. Donnelly, M. Migliore, P. Johnson, C. Fishwick,
cruciferous vegetable phytochemicals, Nutr. Rev. 65 (2007) 259–267. et al., Pharmacophore model of the quercetin binding site of the SIRT6
[143] A. Acharya, I. Das, S. Singh, T. Saha, Chemopreventive properties of protein, J. Mol. Graph Model 49 (2014) 38–46.
indole-3-carbinol, diindolylmethane and other constituents of cardamom [170] N. Singh, S. Ravichandran, D.D. Norton, S.D. Fugmann, R. Moaddel, Synthesis
against carcinogenesis, Recent Pat. Food Nutr. Agric. 2 (2010) 166–177. and characterization of a SIRT6 open tubular column: predicting
[144] B.B. Aggarwal, H. Ichikawa, Molecular targets and anticancer potential of deacetylation activity using frontal chromatography, Anal. Biochem. 436
indole-3-carbinol and its derivatives, Cell Cycle 4 (2005) 1201–1215. (2013) 78–83.
[145] S. Banerjee, D. Kong, Z. Wang, B. Bao, G.G. Hillman, F.H. Sarkar, Attenuation [171] P. Kokkonen, M. Rahnasto-Rilla, P. Mellini, E. Jarho, M. Lahtela-Kakkonen, T.
of multi-targeted proliferation-linked signaling by 3,3 -diindolylmethane Kokkola, Studying SIRT6 regulation using H3K56 based substrate and small
(DIM): from bench to clinic, Mutat. Res. 728 (2011) 47–66. molecules, Eur. J. Pharm. Sci. 63 (2014) 71–76.
[146] Y. Li, X. Li, B. Guo, Chemopreventive agent 3,3 -diindolylmethane selectively [172] F. Du, Y. Feng, J. Fang, M. Yang, MicroRNA-143 enhances chemosensitivity of
induces proteasomal degradation of class I histone deacetylases, Cancer Res. Quercetin through autophagy inhibition via target GABARAPL1 in gastric
70 (2010) 646–654. cancer cells, Biomed. Pharmacother. 74 (2015) 169–177.
[147] L.M. Beaver, T.W. Yu, E.I. Sokolowski, D.E. Williams, R.H. Dashwood, E. Ho, [173] H. Sonoki, T. Sato, S. Endo, T. Matsunaga, M. Yamaguchi, Y. Yamazaki, et al.,
3,3 -Diindolylmethane, but not indole-3-carbinol, inhibits histone Quercetin decreases claudin-2 expression mediated by up-Regulation of
deacetylase activity in prostate cancer cells, Toxicol. Appl. Pharmacol. 263 microRNA miR-16 in lung adenocarcinoma A549Cells, Nutrients 7 (2015)
(2012) 345–351. 4578–4592.
[148] P.B. Busbee, M. Nagarkatti, P.S. Nagarkatti, Natural indoles, [174] G. Lou, Y. Liu, S. Wu, J. Xue, F. Yang, H. Fu, et al., The p53/miR-34a/SIRT1
indole-3-carbinol and 3,3’diindolymethane, inhibit T cell activation by positive feedback loop in quercetin induced apoptosis, Cell Physiol.
staphylococcal enterotoxin B through epigenetic regulation involving HDAC Biochem. 35 (2015) 2192–2202.
expression, Toxicol. Appl. Pharmacol. 274 (2014) 7–16. [175] X. Zhang, Q. Guo, J. Chen, Z. Chen, Quercetin enhances cisplatin sensitivity of
[149] Y. Li, T.G. Vandenboom, Z. Wang, D. Kong, S. Ali, P.A. Philip, et al., miR-146a human osteosarcoma cells by modulating microRNA-217-KRAS axis, Mol.
suppresses invasion of pancreatic cancer cells, Cancer Res. 70 (2010) Cells 38 (2015) 638–642.
1486–1495. [176] F.Q. Yang, M. Liu, W. Li, J.P. Che, G.C. Wang, Zheng JH.Combination of
[150] Y. Jin, X. Zou, X. Feng, 3,3 -Diindolylmethane negatively regulates Cdc25A quercetin and hyperoside inhibits prostate cancer cell growth and
and induces a G2/M arrest by modulation of microRNA 21 in human breast metastasis via regulation of microRNA-21, Mol. Med. Rep. 11 (2015)
cancer cells, Anticancer Drugs 21 (2010) 814–822. 1085–1092.
[151] F. Wang, S. Li, L. Wang, H. Zong, Q. Fan, DATS suppresses growth of [177] M. Athar, J.H. Back, L. Kopelovich, D.R. Bickers, A.L. Kim, Multiple molecular
esophageal squamous cell carcinoma by regulation of ERK1/2, Clin. Lab. 61 targets of resveratrol: anti- carcinogenic mechanisms, Arch. Biochem.
(2015) 315–322. Biophys. 486 (2009) 95–102.
[152] N. Chalabi, S. Satih, L. Delort, Y.J. Bignon, D.J. Bernard-Gallon, Expression [178] J.F. Savouret, M. Quesne, Resveratrol and cancer: a review, Biomed.
profiling by whole-genome microarray hybridization reveals differential Pharmacother. 56 (2002) 84–87.
gene expression in breast cancer cell lines after lycopene exposure, Biochim. [179] R.H. Wang, K. Sengupta, C. Li, H.S. Kim, L. Cao, C. Xiao, et al., Impaired DNA
Biophys. Acta 1769 (2007) 124–130. damage response, genome instability, and tumorigenesis in SIRT1 mutant
[153] N. Chalabi, L. Delort, L. Le Corre, S. Satih, Y.J. Bignon, D. Bernard-Gallon, Gene mice, Cancer Cell 14 (2008) 312–323.
signature of breast cancer cell lines treated with lycopene, [180] K.T. Howitz, K.J. Bitterman, H.Y. Cohen, D.W. Lamming, S. Lavu, J.G. Wood,
Pharmacogenomics 7 (2006) 663–672. et al., Small molecule activators of sirtuins extend Saccharomyces cerevisiae
[154] L.J. Fu, Y.B. Ding, L.X. Wu, C.J. Wen, Q. Qu, X. Zhang, et al., The effects of lifespan, Nature 425 (2003) 191–196.
lycopene on the methylation of the GSTP1 promoter and global methylation [181] A. Bishayee, Cancer prevention and treatment with resveratrol: from rodent
in prostatic cancer cell lines PC3 and LNCaP, Int. J. Endocrinol. 2014 (2014) studies to clinical trials, Cancer Prev. Res. (Phila) 2 (2009) 409–418.
620165. [182] J.G. Wood, B. Rogina, S. Lavu, K. Howitz, S.L. Helfand, M. Tatar, et al., Sirtuin
[155] M. Iciek, I. Kwiecien, L. Wlodek, Biological properties of garlic and activators mimic caloric restriction and delay ageing in metazoans, Nature
garlic-derived organosulfur compounds, Environ. Mol. Mutagen. 50 (2009) 430 (2004) 686–689.
247–265. [183] G. Boily, X.H. He, B. Pearce, K. Jardine, M.W. McBurney, SirT1-null mice
[156] H. Nian, B. Delage, E. Ho, R.H. Dashwood, Modulation of histone deacetylase develop tumors at normal rates but are poorly protected by resveratrol,
activity by dietary isothiocyanates and allyl sulfides: studies with Oncogene 28 (2009) 2882–2893.
sulforaphane and garlic organosulfur compounds, Environ. Mol. Mutagen. [184] R.H. Wang, Y. Zheng, H.S. Kim, X. Xu, L. Cao, T. Luhasen, et al., Interplay
50 (2009) 213–221. among BRCA1, SIRT1, and survivin during BRCA1-associated tumorigenesis,
[157] M.A. Lea, V.M. Randolph, M. Patel, Increased acetylation of histones induced Mol. Cell. 32 (2008) 11–20.
by diallyl disulfide and structurally related molecules, Int. J. Oncol. 15 (1999) [185] L. Kai, S.K. Samuel, A.S. Levenson, Resveratrol enhances p53 acetylation and
347–352. apoptosis in prostate cancer by inhibiting MTA1/NuRD complex, Int. J.
[158] M.A. Lea, V.M. Randolph, J.E. Lee, C. desBordes, Induction of histone Cancer 126 (2010) 1538–1548.
acetylation in mouse erythroleukemia cells by some organosulfur [186] R. Kala, H.N. Shah, S.L. Martin, T.O. Tollefsbol, Epigenetic-based
compounds including allyl isothiocyanate, Int. J. Cancer. 92 (2001) 784–789. combinatorial resveratrol and pterostilbene alters DNA damage response by
[159] M.A. Lea, M. Rasheed, V.M. Randolph, F. Khan, A. Shareef, C. desBordes, affecting SIRT1 and DNMT enzyme expression, including SIRT1-dependent
Induction of histone acetylation and inhibition of growth of mouse ␥-H2AX and telomerase regulation in triple-negative breast cancer, BMC
erythroleukemia cells by S-allylmercaptocysteine, Nutr. Cancer 43 (2002) Cancer 15 (2015) 672.
90–102. [187] M.L. Wu, H. Li, L.J. Yu, X.Y. Chen, Q.Y. Kong, X. Song, et al., Shortterm
[160] N. Druesne, A. Pagniez, C. Mayeur, M. Thomas, C. Cherbuy, P.H. Duée, et al., resveratrol exposure causes in vitro and in vivo growth inhibition and
Diallyl disulfide (DADS) increases histone acetylation and p21(waf1/cip1) apoptosis of bladder cancer cells, PLoS One 9 (2014) e89806.
expression in human colon tumor cell lines, Carcinogenesis 25 (2004) [188] W. Qin, K. Zhang, K. Clarke, T. Weiland, E.R. Sauter, Methylation and miRNA
1227–1236. effects of resveratrol on mammary tumors vs normal tissue, Nutr. Cancer 66
[161] H. Nian, B. Delage, J.T. Pinto, R.H. Dashwood, Allyl mercaptan a garlic-derived (2014) 270–277.
organosulfur compound, inhibits histone deacetylase and enhances Sp3 [189] A.J. Papoutsis, O.I. Selmin, J.L. Borg, D.F. Romagnolo, Gestational exposure to
binding on the P21WAF1 promoter, Carcinogenesis 29 (2008) 1816–1824. the AhR agonist 2,3,7,8-tetrachlorodibenzo-p-dioxin induces BRCA-1
[162] K.L. Cheung, A.N. Kong, Molecular targets of dietary phenethyl promoter hypermethylation and reduces BRCA-1 expression in mammary
isothiocyanate and sulforaphane for cancer chemoprevention, AAPS J. 12 tissue of rat offspring:Preventive effects of resveratrol, Mol. Carcinog. 54
(2010) 87–97. (2015) 261–269.
[163] L.G. Wang, A. Beklemisheva, X.M. Liu, A.C. Ferrari, J. Feng, J.W. Chiao, Dual [190] A. Azimi, M.F. Hagh, M. Talebi, B. Yousefi, A.A. Hossein Pour Feizi, B.
action on promoter demethylation and chromatin by an isothiocyanate Baradaran, et al., Time – and concentration – dependent effects of
restored GSTP1 silenced in prostate cancer, Mol. Carcinog. 46 (2007) 24–31. resveratrol on miR 15a and miR16-1Expression and apoptosis in the
[164] L. Gibellini, M. Pinti, M. Nasi, J.P. Montagna, S. De Biasi, E. Roat, et al., CCRF-CEM acute lymphoblastic leukemia cell line, Asian Pac. J. Cancer Prev.
Quercetin and cancer chemoprevention, Evid. Based Complement Alternat 16 (2015) 6463–6468.
Med. 2011 (2011) 591356. [191] S. Dhar, A. Kumar, A.M. Rimando, X. Zhang, A.S. Levenson, Resveratrol and
[165] S. Tan, C. Wang, C. Lu, B. Zhao, Y. Cui, X. Shi, et al., Quercetin is able to pterostilbene epigenetically restore PTEN expression by targeting oncomiRs
demethylate the p16INK4a gene promoter, Chemotherapy 55 (2009) 6–10. of the miR-17 family in prostate cancer, Oncotarget 6 (2015) 27214–27226.
[166] P.A. Ruiz, A. Braune, G. Holzlwimmer, L. Quintanilla-Fend, D. Haller, [192] J.D. Clarke, R.H. Dashwood, E. Ho, Multi-targeted prevention of cancer by
Quercetin inhibits TNF-induced NF-kappaB transcription factor recruitment sulforaphane, Cancer Lett. 269 (2008) 291–304.
to proinflammatory gene promoters in murine intestinal epithelial cells, J. [193] J. Tomczyk, A. Olejnik, Sulforaphane–a possible agent in prevention and
Nutr. 137 (2007) 1208–1215. therapy of cancer, Postepy Hig. Med. Dosw. 64 (2010) 590–603.
[167] W.J. Lee, Y.R. Chen, T.H. Tseng, Quercetin induces FasL-related apoptosis, in [194] M. Traka, A.V. Gasper, J.A. Smith, C.J. Hawkey, Y. Bao, R.F. Mithen,
part, through promotion of histone H3 acetylation in human leukemia Transcriptome analysis of human colon Caco-2 cells exposed to
HL-60 cells, Oncol. Rep. 25 (2011) 583–591. sulforaphane, J. Nutr. 135 (2005) 1865–1872.
[168] R.V. Priyadarsini, G. Vinothini, R.S. Murugan, P. Manikandan, S. Nagini, The [195] S.M. Meeran, S.N. Patel, T.O. Tollefsbol, Sulforaphane causes epigenetic
flavonoid quercetin modulates the hallmark capabilities of hamster buccal repression of hTERT expression in human breast cancer cell lines, PLoS One
pouch tumors, Nutr. Cancer. 63 (2011) 218–226. 5 (2010) e11457.
98 E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99

[196] M.C. Myzak, P.A. Karplus, F.L. Chung, R.H. Dashwood, A novel mechanism of acetyltransferase activity and prevents heart failure in rats, J. Clin. Invest.
chemoprotection by sulforaphane: inhibition of histone deacetylase, Cancer 118 (2008) 868–878.
Res. 64 (2004) 5767–5774. [222] H.L. Liu, Y. Chen, G.H. Cui, J.F. Zhou, Curcumin a potent anti-tumor reagent,
[197] M.C. Myzak, K. Hardin, R. Wang, R.H. Dashwood, E. Ho, Sulforaphane inhibits is a novel histone deacetylase inhibitor regulating B-NHL cell line Raji
histone deacetylase activity in BpH −1, LnCaP and PC-3 prostate epithelial proliferation, Acta Pharmacol. Sin. 26 (2005) 603–609.
cells, Carcinogenesis 27 (2006) 811–819. [223] K.C. Choi, M.G. Jung, Y.H. Lee, J.C. Yoon, S.H. Kwon, H.B. Kang, et al.,
[198] A. Pledgie-Tracy, M.D. Sobolewski, N.E. Davidson, Sulforaphane induces cell Epigallocatechin-3-gallate, a histone acetyltransferase inhibitor, inhibits
type-specific apoptosis in human breast cancer cell lines, Mol. Cancer Ther. EBV-induced B lymphocyte transformation via suppression of RelA
6 (2007) 1013–1021. acetylation, Cancer Res. 69 (2009) 583–592.
[199] C.P. Wong, A. Hsu, A. Buchanan, Z. Palomera-Sanchez, L.M. Beaver, Andres [224] M.A. Lea, V.M. Randolph, Induction of histone acetylation in rat liver and
Houseman E, et al Effects of sulforaphane and 3,3 -diindolylmethane on hepatoma by organosulfur compounds including diallyl disulfide,
genome-wide promoter methylation in normal prostate epithelial cells and Anticancer Res. 21 (2001) 2841–2845.
prostate cancer cells, PLoS One 9 (2014) e86787. [225] M. Schwab, V. Reynders, S. Loitsch, D. Steinhilber, O. Schroder, J. Stein, The
[200] Z.Y. Su, C. Zhang, J.H. Lee, L. Shu, T.Y. Wu, T.O. Khor, et al., Requirement and dietary histone deacetylase inhibitor sulforaphane induces human
epigenetics reprogramming of Nrf2 in suppression of tumor promoter beta-defensin-2 in intestinal epithelial cells, Immunology 125 (2008)
TPA-induced mouse skin cell transformation by sulforaphane, Cancer Prev. 241–251.
Res. 7 (2014) 319–329. [226] R. Hu, T.O. Khor, G. Shen, W.S. Jeong, V. Hebbar, C. Chen, et al., Cancer
[201] A.M. Mileo, S. Miccadei, Polyphenols as modulator of oxidative stress in chemoprevention of intestinal polyposis in ApcMin/+ mice by sulforaphane,
cancer disease: new therapeutic strategies, Oxid. Med. Cell. Longevity 2016 a natural product derived from cruciferous vegetable, Carcinogenesis 27
(2016) 6475624, http://dx.doi.org/10.1155/2016/6475624. (2006) 2038–2046.
[202] M.C. Myzak, W.M. Dashwood, G.A. Orner, E. Ho, R.H. Dashwood, [227] A. Pledgie-Tracy, M.D. Sobolewski, N.E. Davidson, Sulforaphane induces cell
Sulforaphane inhibits histone deacetylase in vivo and suppresses type-specific apoptosis in human breast cancer cell lines, Mol. Cancer Ther.
tumorigenesis in Apc-minus mice, FASEB J. 20 (2006) 506–508. 6 (2007) 1013–1021.
[203] M.C. Myzak, P. Tong, W.M. Dashwood, R.H. Dashwood, E. Ho, Sulforaphane [228] M. Chakrabarti, M. Khandkar, N.L. Banik, S.K. Ray, Alterations in expression
retards the growth of human PC-3 xenografts and inhibits HDAC activity in of specific microRNAs by combination of 4-HPR and EGCG inhibited growth
human subjects, Exp. Biol. Med. (Maywood) 232 (2007) 227–234. of human malignant neuroblastoma cells, Brain Res. 1454 (2012) 1–13.
[204] W.J. Lee, J.Y. Shim, B.T. Zhu, Mechanisms for the inhibition of DNA [229] M. Chakrabarti, N.L. Banik, S.K. Ray, miR-138 overexpression is more
methyltransferases by tea catechins and bioflavonoids, Mol. Pharmacol. 68 powerful than hTERT knockdown to potentiate apigenin for apoptosis in
(2005) 1018–1030. neuroblastoma in vitro and in vivo, Exp. Cell Res. 319 (2013) 1575–1585.
[205] H.C. Huang, T.D. Way, C.L. Lin, J.K. Lin, EGCG stabilizes p27kip1 in [230] L. Jiang, Q. Huang, J. Chang, E. Wang, X. Qiu, MicroRNA HSA-miR-125a-5p
E2-stimulated MCF-7 cells through down-regulation of the Skp2 protein, induces apoptosis by activating p53 in lung cancer cells, Exp. Lung Res. 37
Endocrinology 149 (2008) 5972–5983. (2011) 387–398.
[206] J.B. Berletch, C. Liu, W.K. Love, L.G. Andrews, S.K. Katiyar, T.O. Tollefsbol, [231] S. Saini, S. Arora, S. Majid, V. Shahryari, Y. Chen, G. Deng, et al., Curcumin
Epigenetic and genetic mechanisms contribute to telomerase inhibition by modulates microRNA-203-mediated regulation of the Src-Akt axis in
EGCG, J. Cell Biochem. 103 (2008) 509–519. bladder cancer, Cancer Prev. Res. (Phila.) 4 (2011) 1698–1709.
[207] E. Navarro-Peran, J. Cabezas-Herrera, L.S. Campo, J.N. Rodriguez-Lopez, [232] J. Zhang, Y. Du, C. Wu, X. Ren, X. Ti, J. Shi, et al., Curcumin promotes
Effects of folate cycle disruption by the green tea polyphenol apoptosis in human lung adenocarcinoma cells through miR-186n signaling
epigallocatechin-3-gallate, Int. J. Biochem. Cell Biol. 39 (2007) 2215–2225. pathway, Oncol. Rep. 24 (2010) 1217–1223.
[208] K. Kato, N.K. Long, H. Makita, M. Toida, T. Yamashita, D. Hatakeyama, et al., [233] S.M. Gao, J.J. Yang, C.Q. Chen, J.J. Chen, L.P. Ye, L.Y. Wang, et al., Pure
Effects of green tea polyphenol on methylation status of RECK gene and curcumin decreases the expression of WT1 by upregulation of miR-15a and
cancer cell invasion in oral squamous cell carcinoma cells, Br. J. Cancer. 99 miR-16-1 in leukemic cells, J. Exp. Clin. Cancer Res. 31 (2012) 27.
(2008) 647–665. [234] M. Sun, Z. Estrov, Y. Ji, K.R. Coombes, D.H. Harris, R. Kurzrock, et al.,
[209] W.Y. Tang, R. Newbold, K. Mardilovich, W. Jefferson, R.Y. Cheng, M. Curcumin (diferuloylmethane) alters the expression profiles of microRNAs
Medvedovic, et al., Persistent hypomethylation in the promoter of in human pancreatic cancer cells, Mol. Cancer Ther. 7 (2008) 464–473.
nucleosomal binding protein 1 (Nsbp1) correlates with overexpression of [235] W.P. Tsang, T.T. Kwok, Epigallocatechin gallate up-regulation of miR-16 and
Nsbp1 in mouse uteri neonatally exposed to diethylstilbestrol or genistein, induction of apoptosis in human cancer cells, J. Nutr. Biochem. 21 (2010)
Endocrinology 149 (2008) 5922–5931. 140–146.
[210] M.Z. Fang, Z. Jin, Y. Wang, J. Liao, G.Y. Yang, L.D. Wang, et al., Promoter [236] S.U. Gandhy, K. Kim, L. Larsen, R.J. Rosengren, S. Safe, Curcumin and
hypermethylation and inactivation of O(6)-methylguanine-DNA synthetic analogs induce reactive oxygen species and decreases specificity
methyltransferase in esophageal squamous cell carcinomas and its protein (Sp) transcription factors by targeting microRNAs, BMC Cancer 12
reactivation in cell lines, Int. J. Oncol. 26 (2005) 615–622. (2012) 564.
[211] M.B. Vandegehuchte, F. Lemière, L. Vanhaecke, W. Vanden Berghe, C.R. [237] Y. Jin, 3,3 -Diindolylmethane inhibits breast cancer cell growth via
Janssen, Direct and transgenerational impact on Daphnia magna of miR-21-mediated Cdc25A degradation, Mol. Cell Biochem. 358 (2011)
chemicals with a known effect on DNA methylation, Comp. Biochem. 345–354.
Physiol. C Toxicol. Pharmacol. 151 (2010) 278–285. [238] D. Kong, E. Heath, W. Chen, M. Cher, I. Powell, L. Heilbrun, et al., Epigenetic
[212] J. Paluszczak, V. Krajka-Kuźniak, W. Baer-Dubowska, The effect of dietary silencing of miR-34a in human prostate cancer cells and tumor tissue
polyphenols on the epigenetic regulation of gene expression in MCF7 breast specimens can be reversed by BR-DIM treatment, Am. J. Transl. Res. 4 (2012)
cancer cells, Toxicol. Lett. 192 (2010) 119–125. 14–23.
[213] B. Stefanska, K. Rudnicka, A. Bednarek, K. Fabianowska-Majewska, [239] T. Melkamu, X. Zhang, J. Tan, Y. Zeng, F. Kassie, Alteration of microRNA
Hypomethylation and induction of retinoic acid receptor beta 2 by expression in vinyl carbamate-induced mouse lung tumors and modulation
concurrent action of adenosine analogues and natural compounds in breast by the chemopreventive agent indole-3-carbinol, Carcinogenesis 31 (2010)
cancer cells, Eur. J. Pharmacol. 638 (2010) 47–53. 252–258.
[214] N. Singh, K. Misra, Computational screening of molecular targets in [240] J. Ahn, H. Lee, C.H. Jung, T. Ha, Lycopene inhibits hepatic steatosis via
Plasmodium for novel non-resistant anti-malarial drugs, Bioinformation 3 microRNA-21-induced downregulation of fatty acid-binding protein 7 in
(2009) 255–262. mice fed a high-fat diet, Mol. Nutr. Food Res. 56 (2012) 1665–1674.
[215] J. Kang, J. Chen, Y. Shi, J. Jia, Y. Zhang, Curcumin-induced histone [241] E.P. Chiang, S.C. Chiu, M.H. Pai, Y.C. Wang, F.Y. Wang, Y.H. Kuo, et al.,
hypoacetylation: the role of reactive oxygen species, Biochem. Pharmacol. Organosulfur garlic compounds induce neovasculogenesis in human
69 (2005) 1205–1213. endothelial progenitor cells through a modulation of MicroRNA 221 and the
[216] L. Cui, J. Miao, T. Furuya, X. Li, X.Z. Su, L. Cui, PfGCN5-mediated histone H3 PI3-K/Akt signaling pathways, J. Agric. Food Chem. 61 (2013) 4839–4849.
acetylation plays a key role in gene expression in Plasmodium falciparum, [242] X. Xiao, B. Chen, X. Liu, P. Liu, G. Zheng, F. Ye, et al., Diallyl disulfide
Eukaryot. Cell 6 (2007) 1219–1227. suppresses SRC/Ras/ERK signaling-mediated proliferation and metastasis in
[217] J.C. Sng, H. Taniura, Y. Yoneda, Histone modifications in kainate-induced human breast cancer by up-regulating miR-34a, PLoS One 9 (2014) e112720.
status epilepticus, Eur. J. Neurosci. 23 (2006) 1269–1282. [243] S. Dhar, C. Hicks, A.S. Levenson, Resveratrol and prostate cancer: promising
[218] J. Chiu, Z.A. Khan, H. Farhangkhoee, S. Chakrabarti, Curcumin prevents role for microRNAs, Mol. Nutr. Food Res. 55 (2011) 1219–1229.
diabetes-associated abnormalities in the kidneys by inhibiting p300 and [244] S. Sheth, S. Jajoo, T. Kaur, D. Mukherjea, K. Sheehan, L.P. Rybak, et al.,
nuclear factor-kappaB, Nutrition 25 (2009) 964–972. Resveratrol reduces prostate cancer growth and metastasis by inhibiting the
[219] K. Tikoo, R.L. Meena, D.G. Kabra, A.B. Gaikwad, Change in post-translational Akt/MicroRNA-21 pathway, PLoS One 7 (2012) e51655.
modifications of histone H3, heat-shock protein-27 and MAP kinase p38 [245] E. Tili, J.J. Michaille, H. Alder, S. Volinia, D. Delmas, N. Latruffe, C.M. Croce,
expression by curcumin in streptozotocin-induced type I diabetic Resveratrol modulates the levels of microRNAs targeting genes encoding
nephropathy, Br. J. Pharmacol. 153 (2008) 1225–1231. tumor-suppressors and effectors of TGFbeta signaling pathway in SW480
[220] H.L. Li, C. Liu, G. de Couto, M. Ouzounian, M. Sun, A.B. Wang, et al., Curcumin cells, Biochem. Pharmacol. 80 (2010) 2057–2065.
prevents and reverses murine cardiac hypertrophy, J. Clin. Invest. 118 [246] M.A. Sakurai, Y. Ozaki, D. Okuzaki, Y. Naito, T. Sasakura, A. Okamoto, et al.,
(2008) 879–893. Gefitinib and luteolin cause growth arrest of human prostate cancer PC-3
[221] T. Morimoto, Y. Sunagawa, T. Kawamura, T. Takaya, H. Wada, A. Nagasawa, cells via inhibition of cyclin G-associated kinase and induction of miR-630,
et al., The dietary compound curcumin inhibits p300 histone PLoS One 9 (2014) e100124.
E. Shankar et al. / Seminars in Cancer Biology 40–41 (2016) 82–99 99

[247] H. Wu, M. Huang, Y. Liu, Y. Shu, P. Liu, Luteolin induces apoptosis by expression in patients with localised prostate cancer before radical
up-regulating miR-34a in human gastric cancer cells, Technol. Cancer Res. prostatectomy, Br. J. Nutr. 108 (2012) 2138–2147.
Treat. 14 (2015) 747–755. [254] N. Miyanaga, H. Akaza, S. Hinotsu, T. Fujioka, S. Naito, M. Namiki, et al.,
[248] J. Xiao, A.Y. Gong, A.N. Eischeid, D. Chen, C. Deng, C.Y. Young, et al., Prostate cancer chemoprevention study: an investigative randomized
miR-141modulates androgen receptor transcriptional activity in human control study using purified isoflavones in men with rising prostate-specific
prostate cancer cells through targeting the small heterodimer partner antigen, Cancer Sci. 103 (2012) 125–130.
protein, Prostate 72 (2012) 1514–1522. [255] E.I. Heath, L.K. Heilbrun, J. Li, U. Vaishampayan, F. Harper, P. Pemberton,
[249] H. Ide, S. Tokiwa, K. Sakamaki, K. Nishio, S. Isotani, S. Muto, et al., Combined et al., A phase I dose-escalation study of oral BR-DIM in castrate-resistant,
inhibitory effects of soy isoflavones and curcumin on the production of non-metastatic prostate cancer, Am. J. Transl. Res. 2 (2010) 402–411.
prostate-specific antigen, Prostate 70 (2010) 1127–1133. [256] S. Mariani, L. Lionetto, M. Cavallari, A. Tubaro, D. Rasio, C. De Nunzio, et al.,
[250] R.A. Sharma, H.R. McLelland, K.A. Hill, C.R. Ireson, S.A. Euden, M.M. Manson, Low prostate concentration of lycopene is associated with development of
et al., Pharmacodynamic and pharmacokinetic study of oral Curcuma extract prostate cancer in patients with high-grade prostatic intraepithelial
in patients with colorectal cancer, Clin. Cancer Res. 7 (2001) 1894–1900. neoplasia, Int. J. Mol. Sci. 15 (2014) 1433–1440.
[251] J.J. Johnson, H.H. Bailey, H. Mukhtar, Green tea polyphenols for prostate [257] D.M. Goldberg, J. Yan, G.J. Soleas, Absorption of three wine-related
cancer chemoprevention: a translational perspective, Phytomedicine 17 polyphenols in three different matrices by healthy subjects, Clin. Biochem.
(2010) 3–13. 36 (2003) 79–87.
[252] M.M. Nguyen, F.R. Ahmann, R.B. Nagle, C.H. Hsu, J.A. Tangrea, H.L. Parnes, [258] J.J. Alumkal, R. Slottke, J. Schwartzman, G. Cherala, M. Munar, J.N. Graff,
et al., Randomized, double-blind, placebo-controlled trial of polyphenon E in et al., A phase II study of sulforaphane-rich broccoli sprout extracts in men
prostate cancer patients before prostatectomy: evaluation of potential with recurrent prostate cancer, Investig. New Drugs 33 (2015) 480–489.
chemopreventive activities, Cancer Prev. Res. 5 (2012) 290–298.
[253] B. Lazarevic, C. Hammarstrom, J. Yang, H. Ramberg, L.M. Diep, S.J. Karlsen,
et al., The effects of short-term genistein intervention on prostate biomarker

S-ar putea să vă placă și