Вы находитесь на странице: 1из 13

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/223994740

Epigenetic impact of dietary polyphenols in


cancer chemoprevention: Lifelong remodeling
of our epigenomes
Article in Pharmacological Research March 2012
DOI: 10.1016/j.phrs.2012.03.007 Source: PubMed

CITATIONS

READS

96

389

1 author:
Wim Vanden Berghe
University of Antwerp
168 PUBLICATIONS 8,735 CITATIONS
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Pharmacology of Natural Product / Neuroprotection View project

All content following this page was uploaded by Wim Vanden Berghe on 25 February 2014.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.

Pharmacological Research 65 (2012) 565576

Contents lists available at SciVerse ScienceDirect

Pharmacological Research
journal homepage: www.elsevier.com/locate/yphrs

Review

Epigenetic impact of dietary polyphenols in cancer chemoprevention:


Lifelong remodeling of our epigenomes
Wim Vanden Berghe a,b,
a
Laboratory of Protein Science, Proteomics and Epigenetic Signaling (PPES), Department of Biomedical Sciences, University Antwerp, Campus Drie Eiken, Universiteitsplein 1, Wilrijk,
Belgium
b
Laboratory of Eukaryotic Gene Expression and Signal Transduction (LEGEST), Department of Physiology, Ghent University, K.L. Ledeganckstraat 35, Gent, Belgium

a r t i c l e

i n f o

Article history:
Received 10 February 2012
Received in revised form 10 March 2012
Accepted 13 March 2012
Keywords:
Cancer
Chemoprevention
Diet
Polyphenol
Epigenome
Epigenetics

a b s t r a c t
Cancer, as one of the non-communicable diseases, remains one of the leading causes of death around the
world. Recently, epigenetic changes in DNA methylation patterns at CpG sites (epimutations) or deregulated chromatin states of tumor promoting genes and noncoding RNAs emerged as major governing
factors in tumor progression and cancer drug sensitivity. Furthermore, various environmental factors such
as nutrition, behavior, stress, and toxins remodel our epigenomes lifelong in a benecial or detrimental
way. Since epigenetic marks (epimutations) are reversible in contrast to genetic defects, chemopreventive
nutritional polyphenols (soy, genistein, resveratrol, catechin, curcumin) are currently evaluated for their
ability to reverse adverse epigenetic marks in cancer (stem) cells to attenuate tumorigenesis-progression,
prevent metastasis or sensitize for drug sensitivity. Although polyphenols in fruit and vegetables may
help to reduce the risk of cancer, few protective effects have been rmly established, presumably because
of inappropriate timing or dosing of diet exposure or due to confounding factors such as smoking and
alcohol. In this review will discuss the possible epigenetic contributions of dietary polyphenols in cancer
chemoprevention.
2012 Elsevier Ltd. All rights reserved.

Contents
1.
2.
3.
4.
5.
6.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Chromatin states in the epigenomic cancer landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Dietary chemoprevention of cancer-inammation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Nutri-epigenomics: lifelong remodeling of our epigenomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Epigenetic targets of bioactive dietary components for cancer prevention and therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusion and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Cancer remains a major health problem and is responsible for
one in eight deaths worldwide. Genome-wide association studies
have identied hundreds of genetic variants associated with complex human diseases and traits, and have provided valuable insights
into their genetic architecture. Despite the success of genome-wide
association studies in identifying loci associated with cancer, a substantial proportion of the causality remains unexplained, leaving

Correspondence address: Laboratory of Protein Science, Proteomics and Epigenetic Signaling (PPES), Department of Biomedical Sciences, University Antwerp,
Campus Drie Eiken, Universiteitsplein 1, Wilrijk, Belgium.
E-mail address: wim.vandenberghe@ua.ac.be
1043-6618/$ see front matter 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.phrs.2012.03.007

565
566
568
570
571
572
572
572

many questions how the remaining missing heritability can be


explained, although polygenic disease traits may account for some
of this limitations [13]. Only a minority of cancers are caused by
germline mutations, whereas the vast majority (90%) are linked
to somatic mutations and environmental factors [4]. Also, an estimated 55% increase in cancer incidence is expected by the year 2020
[5]. A recent survey of the global incidence of cancer shows that the
age-adjusted cancer incidence in the Western world is above 300
cases per 100,000 population, whereas that in Asian countries is
less than 100 cases per 100,000. Observational studies have suggested that lifestyle risk factors such as tobacco, obesity, alcohol,
sedentary lifestyle, high-fat diet, radiation, and infections are major
contributors in cancer causes, which is further emphasized by the
increase in cancer cases among immigrants from Asian to Western

566

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

Fig. 1. Interplay of environmental oxidative stress, ROS, RNS, diet chemoprevention and epigenetics in Cancer. Oxidative damage of CpG motifs can result in cytosine
hydroxymethylation and formation of 8-oxoguanine, which may interfere with epigenetic dynamics in cancer biology, metastasis and drug sensitivity.

countries [4,6,7]. Reciprocally, a reasonable good fraction of cancer


deaths maybe prevented by modifying the diet composition (i.e.
content of ber, polyphenols, fat/oil, protein, spices, cereals, etc.)
and regular physical exercise [4,810]. Rather than the chemical
conversion of food to energy and body matter of classic metabolism,
food is now also a conditioning environment that shapes the activity of the (epi)genome and determines stress adaptative responses,
metabolism, immune homeostasis and the physiology of the body
[11,12]. The contribution of epigenetic changes (epimutations) to
human disease is probably underestimated. Epigenetics encompasses several extra-genetic processes such as DNA methylation
(methylation of cytosines within CpG dinucleotides), histone tail
modications (including acetylation, phosphorylation, methylation, sumoylation, ribosylation and ubiquitination), non-coding
RNA functions, regulation of polycomb group proteins and the epigenetic cofactor modiers, all of which may alter gene expression
but do not involve changes in the DNA sequence itself [1317]
(Fig. 1).
2. Chromatin states in the epigenomic cancer landscape
DNA methylation is the best-known epigenetic mark [18,19].
It is catalyzed by two types of DNMTs: DNMT1 is a maintenance methyltransferase, whereas both DNMT3A and DNMT3B are
de novo methyltransferases [20,21]. The role of DNMT2 in DNA
methylation is minor, its enzymology being largely directed to
tRNA. DNA methylation is normally associated with gene inactivation and it usually occurs in CpG dinucleotides. Alternatively,
DNA methylation of transcription factor binding sites which prevents binding of repressor proteins, may paradoxically induce
gene activation. CpGs are normally methylated when scattered
throughout the genome, but are mostly unmethylated when they
are clustered as CpG islands at 5 ends of many genes. Hypermethylation of CpG-rich promoters triggers local histone code
modications resulting in a cellular camouage mechanism that
sequesters gene promoters away from transcription factors and
results in stable silencing of gene expression. DNA methylation at
CpG dinucleotides occurs upon transfer of S-adenosylmethionine
(SAM) on cytosine by DNMTs. Whereas DNMT3A/B are responsible for DNA methylation during development (differentiation),
DNMT1 is in charge of maintaining DNA methylation patterns
in DNA replication during cell division. In mammalian cells, the
delity of maintenance of methylation is 9799.9% per mitosis,
whereas de novo methylation is as high as 35% per mitosis, thus
creating possibilities for epigenetic changes. Although in most cases

DNA methylation is a stable epigenetic mark, reduced levels of


methylation can also be observed during development. This net
loss of methylation can either occur passively by replication in
the absence of funtional maintenance methylation pathways, or
actively, by removing methylated cytosines. In mammals, coupling of 5-methylcytosine deaminase and thymine DNA glycosylase
activities maybe responsible for DNA demethylation. Alternatively, a role for the 5-hydroxymethylcytosine modication in
mammalian DNA demethylation has also been proposed as an
intermediate in an active DNA demethylation pathway involving
DNA repair and 5-hydroxymethylcytosine-specic DNA glycosylase activity [21]. Of particular interest, ROS and oxidative stress
may affect DNA demethylation by DNA oxidation or Tet mediated
hydroxymethylation [22,23].
In general, DNA is wrapped around nucleosomes, which are
arranged as regularly spaced beads (146 bp DNA/nucleosome)
along the DNA. Typically, nucleosomes consist of a histone octamer
of histones (H)2A/B, H3 and H4. The DNA bridging two adjacent
nucleosomes is normally bound by the linker histone H1 and is
termed linker DNA. While the core histones are bound relatively
tightly to DNA, chromatin is largely maintained by the dynamic
association with its architectural proteins. Before most activators
of a gene access their DNA-binding sites, a transition from a condensed heterochromatin (solenoid-like ber) to a decondensed
euchromatin (beads on a string) structure appears to take place.
Conversely, the acquisition of a more condensed heterochromatin
structure is often associated with gene silencing [15]. This structural restriction of silenced chromatin on gene expression can be
overcome by chromatin writer, reader and eraser enzyme complexes, which remodel nucleosomes along the DNA or reversibly
modify (acetylation, phosphorylation, ubiquitylation, glycosylation, sumoylation) histones on lysine, arginine, serine or threonine
residues of amino-terminal histone tails and establish specic chromatin states involved in transcription [15,24,25].
One critical facet of histone and DNA modifying enzymes
is that their activity also depends on intracellular levels of
essential metabolites (acetyl-coA, Fe, ketoglutarate, NAD+ , Sadenosylmethionine) of which the concentrations are tighly linked
to global cellular metabolism and energy levels [2631]. Gene
regulation is thus linked to the metabolic status of cells (Fig. 2).
To maintain uncontrolled cell proliferation of cancer cells, energy
metabolism needs to be adjusted in order to fuel cell growth and
vision. In contrast to healthy cells which mainly generate energy
from oxidative breakdown of pyruvate, cancer cell reprogram their
glucose metabolism, limiting their energy metabolism largely to

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

567

Fig. 2. Once critical facet of histone modications is that they are elicited by specic enzymatic activities that depend on the intracellular levels of essential metabolites:
these metabolites sense ROS, RNS, cellular metabolism, nutrients and energy levels in the cell. NAD, acetyl-coenzyme A (Acetyl-coA) and S-adenosylmethionine (SAM)
are elemental for epigenetic control of transcription including methylation of DNA and posttranslational modications of histones and non-histone chromatin factors (not
shown). NAD contributes to transcriptional control mainly via the activity of the protein deacetylase sirtuin (SIRT), which uses NAD as one of the substrates. Sirtuins are also
important for maintaining the activity of the acetyl-coA acetyltransferases. Ac-coA is synthesized by acetyl-coA-synthetase and ATP-citrate lyase that use acetate and citrate
as the precursor, respectively. Citrate is an intermediate/product of the TCA cycle. SAM is the methyl donor for DNA, RNA, histones and non-histone protein methylation.
SAH generated in each round of methylation reaction is a potent inhibitor of methyltransferases and has to be cleared by SAH hydrolase. NAD is an essential coenzyme for
SAHH. Synthesis of methionine from homocystein is achieved through extracting the methyl group from betaine, derived from choline, or 5-methyl-THF, a derivative of
folic acid. Metabolism of phospholipids and folic acid may thus indirectly contribute to epigenetic regulation. Likewise, the abundance of NAD and citrate is linked to the
cellular energy ux, e.g. the TCA cycle. Changes in the expression of certain genes may therefore be inuenced signicantly. Abbreviations used: Ac-coA, acetyl-coenzyme A;
ACS, acetyl-coA-synthetase; AC-ACS acetylated-ACS; Ado, adenosine; HAT, histone acetyltransferase; Hcy homocysteine; MTases, methyltransferases; NAD, Nicotinamide
adenine dinucleotide; ROS, reactive oxygen species; RNS, reactive nitrogen species; SAH, S-adenosyl homocysteine; TCA tricarboxylic cycle; THF, tetrahydrofolate.

glycolysis. The fundamental difference in ratio of glycolysis to mitochondrial respiration between normal and cancerous cells is also
known as the Warburg effect. As such, dynamic changes in energy
levels and metabolite concentrations in the inammatory tumor
microenvironment can have signicant epigenetic changes through
epigenetic cofactor enzymes [3237]. Specic sets of histone modications and/or variants are associated with genes that are actively
transcribed or are repressed, a phenomenon dened as the histone
code [15]. Based on coexisting histone marks and genomewide
ChIP-seq data available within the ENCODE consortium, principal component analysis allowed to reduce the complexity of the
histone code into 9 different chromatin states, linked to different
functional regulatory features [24,25].
The combinatorial nature of DNA methylation and histone modications signicantly extends the information potential of the
genetic cancer code [38]. The most studied epigenetic lesion, which
is DNA hypermethylation at the promoter region of many genes
[18,39], is proved to be responsible for silencing of more than 600
cancer-related genes and this number is still rising. They include
tumor suppressor genes, oncogenes as well as genes involved
in the cell-cycle regulation, DNA repair, angiogenesis, metastasis
and apoptosis [40]. Oxidative stress (ROS, RNS) and inammatory
damage play an important role in epigenetic reprogramming of

expression of cytokines, oncogenes and tumor suppressor genes,


thereby setting up a ground for chronic inammatory diseases and
carcinogenesis [4144]. On the other hand, global hypomethylation
of the DNA is said to activate endoparasitic sequences and causes
the global chromosome instability leading to various mutations and
cancer progression [45].
There is good evidence that also noncoding RNAs regulate chromatin architecture [16,4651]. The term non-coding RNA (ncRNA)
is commonly employed for RNA that does not encode a protein.
Although it has been generally assumed that most genetic information is transacted by proteins, recent evidence suggests that the
majority of the genomes of mammals and other complex organisms
is in fact transcribed into ncRNAs, many of which are alternatively spliced and/or processed into smaller products. Besides tRNA
and rRNA, these ncRNAs include long-noncodingRNAs (lncRNAs),
microRNAs (miRNAs) and tinyRNAs (tiRNAs) as well as several other classes of, sometimes yet-to-be-discovered [4648,51].
These RNAs (including those derived from introns) appear to
comprise a hidden layer of internal signals that control various levels of gene expression in physiology and development,
including chromatin architecture/epigenetic memory, transcription (enhancer function), RNA splicing, editing, translation and
turnover [16,17,5254]. RNA regulatory networks may determine

568

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

Fig. 3. Dietary reversal of epigetic changes in cancer cells. Changes in DNA methylation have been recognized as one of the most common molecular alterations in human
neoplasia and hypermethylation of gene-promoter regions is being revealed as one of the most frequent mechanisms of loss of gene function. This gure summarizes how
changes in DNA-methylation (epimutations) contribute to the 6 hallmarks of a cancer cell, i.e. limitless replicative potential, self-sufciency in growth signals, insensitivity to
growth-inhibitory signals, evasion of programmed cell death, sustained angiogenesis and tissue invasion and metastasis. Since epigenetic changes (epimutations) are more
easily reversible (when compared with genetic mutations), this has inspired various research efforts aiming to identify dietary phytochemicals (nutri-epigenomics) which
can reverse epimutations and/or prevent cancer progression.

most of our complex characteristics and play a signicant role in


disease [52]. For example, miRNAs can change expression levels of
the epigenetic machinery (DNMT, HDAC, sirtuin (SIRT), polycomb
(Pc) proteins, etc.) by posttranscriptional gene regulation involving base pairing with 3 untranslated (UTR) regions in their target
mRNAs resulting in mRNA degradation or inhibition of translation
[16,53,5559]. Alternatively, long ncRNAs and tiRNAs can regulate
gene expression and/or DNA methylation by promoter association
[46,47,5052]. DNA-methylation can thus also be RNA-directed
[16,60].
Epigenetic defects in DNA methylation patterns at CpG sites
(epimutations), abnormalities in histone modications, chromatin
remodeling and noncoding RNAs (microRNA, long noncoding RNA)
or corrupt chromatin states of tumor suppressor genes or oncogenes recently emerged as major governing factors in tumor
progression and cancer drug sensitivity [14,1618,45,53,56,6163].
In addition, genetic mutations of epigenetic modifying enzymes
add another level of regulatory complexity [15,6468]. Recent
advances in genomic technologies have initiated large-scale
studies to map cancer-associated epigenetic variation, specifically variation in DNA methylation and chromatin states
[3,24,25,38,69,7072]. Given the prevalence of reversible epigenetic abnormalities in different cancers, epigenetic therapy holds
great promise for treatment. The future of specic and effective
epigenetic drug design will rely on our ability in understanding
epigenomic landscapes in normal and cancerous disease states
[7379].
3. Dietary chemoprevention of cancer-inammation
Cancer cells are distinguished by several distinct characteristics, such as self-sufciency in growth signal, resistance to growth
inhibition, limitless replicative potential, evasion of apoptosis,

sustained angiogenesis, and tissue invasion and metastasis [80]


(Fig. 3). Tumor cells acquire these properties due to cumulative
epigenetic changes of multiple genes and associated cell signaling pathways, most of which are linked to inammation. Immune
cells also inltrate in tumors, engage in an extensive and dynamic
crosstalk with cancer cells [42,81,82]. Inammation also affects
immune surveillance and responses to therapy [8386]. For that
reason, rationally designed drugs that target a single gene product
are unlikely to be of use in preventing or treating cancer. Moreover, targeted drugs can cause serious and even life-threatening
side effects or therapy resistance [80]. When a complex system
starts to dysfunction, it is generally best to x it early. Often,
cancers have a long latency period often 20 years or more. By
the time they are clinically detectable, the system has degenerated into a disorganized, chaotic mess at which point it may be
beyond repair [87]. Therefore, there is an urgent need for safe
and effective chemopreventive multifunctional drugs that act at
entire networks in the body, rather than single targets [88]. The
basic idea of cancer chemoprevention is to arrest or reverse the
progression of premalignant cells toward full malignancy using
physiological mechanisms that do not kill healthy cells [4,43,89,90].
The global demand for more affordable therapeutics and concerns about sides effects of commonly used drugs has renewed
interest in phytochemicals and traditional medicines which allow
chronic use [9193]. Studies on a wide spectrum of plant secondary
metabolites extractable as natural products from fruits, vegetables, teas, spices, and traditional medicinal herbs have identied
various bioactive polyphenols (genistein and soy polyphenols,
resveratrol, green tea polyphenols, curcumin, coffee polyphenols,
apple polyphenols), that regulate multiple cancer-inammation
pathways and epigenetic cofactors, are cost effective, exhibit low
toxicity, and are readily available [9496]. The recent advances
in genomics and metabolomics have enabled biologists to better

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

569

Fig. 4. Dietary modulation of transcription factor pathways which regulate chromatin oscillation dynamics between euchromatic and heterochromatic states at oncogenes
and tumor suppressor genes.

Fig. 5. Development of functional foods or dietary supplements as nutrition based epigenetic modulators of chromatin writers, readers and erasers in cancer chemoprevention. HAT, histone acetyltransferase; HDAC, histone deacetylase; DNMT, DNA methyltransferase; KMT, lysine methyltransferase; KDM, lysine demethylase; Me-CpG,
Methylcytosine; R, transcription repressor; A, Transcription activator; Ac, acety; Me, methyl.

570

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

investigate the potential use of immunomodulatory natural products for treatment or control of cancerous diseases. More recently,
evidence is emerging that specic combinations of polyphenols
maybe far more effective in protecting against cancer than isolated
compounds [92,97]. The cancer preventive or protective activities of the various immunomodulatory natural products lie in
their effects on cellular defenses including detoxifying and antioxidant enzyme systems, and the induction of anti-inammatory
and antitumor or antimetastasis responses, often by targeting
specic key transcription factors (i.e. like nuclear factor kappa B
(NFB), activator protein (AP-1), signal transducers and activators
of transcription (STAT3), nuclear factor erythroid 2-related factor (NRF2), peroxisome proliferator-activated receptor- (PPAR ),
estrogen receptor, liver X receptor (LXR), hypoxia inducible factor1 (HIF-1)), epigenetic cofactors as microRNAs which are involved
in tumor progression [59,98,99]. Typically, dysregulation of transcription factor activity is the result of numerous mechanisms,
such as changes in gene expression, proteinprotein interactions
and post-translational modications, leading to deregulation of
gene products that are involved in both inammation and carcinogenesis (Fig. 4). Remarkably, transient inammatory pathways
can trigger mitotic stable epigenetic switches from nontransformed to metastatic cancer cells via feedback signaling involving
NFB and STAT3 transcription factors, Lin28 and let-7 microRNAs
and the cytokine IL6 in the tumor microenvironment [100,101].
Similarly, emerging data demonstrate the direct inuence of certain anti-inammatory dietary factors (for example polyphenols,
isothiocyanates, epicatechins) and micronutrients (for example
folic acid, selenium) on heritable gene expression, activity of
the epigenetic machinery, DNA methylation, chromatin remodeling or microRNAs [11,102110] (Table 1). Because epigenetic
changes are reversible, developing drugs that control epigenetic
regulation now attracts substantial research investment, including
the development of functional foods or supplements as nutrition based epigenetic modulators for cancer chemoprevention
[10,12,59,92,111115] (Fig. 5). Sofar, it is not clear whether occasional consumption of polyphenols results in longterm epigenetic
regulation of gene expression and/or downstream chemopreventive effects.

4. Nutri-epigenomics: lifelong remodeling of our


epigenomes
Human epidemiological studies and appropriately designed
dietary interventions in animal models have provided considerable evidence to suggest that maternal nutritional imbalance and
metabolic disturbances, during critical time windows of development, may have a persistent effect on the health of offspring and
may even be transmitted to the next generation [102,116122].
This has led to the hypothesis of fetal programming and new
term developmental origin of health and disease (DOHaD): common disorders, such as cancer, obesity, cardiovascular disease
(CVD), diabetes, hypertension, asthma and even schizophrenia,
take root in early nutrition during gestation and continues during
lactation [90,102,123129]. Similarly, there is increasing evidence
in both plants and animals that nutritional intervention (caloric,
iron and protein restriction, polyphenol-, folate-, micronutrient-,
fat- or carbohydrate-rich diet), maternal diabetes, during pregnancy and lactation affects health in following generation(s)
[90,108,120,130133]. The various non-Mendelian features of
metabolic disease, cancer or chronic inammatory disorders, clinical differences between men and women or monozygotic twins and
uctuations in the course of the disease are consistent with epigenetic mechanisms in the inuence of fetal and/or lifelong nutrition
or stochastic events on adult phenotype [102,116121,134136].
Thus, lifetime shapes the multitude of epigenomes not only within,
but also across generations [102,117,120,125,137139]. Interest in
transgenerational epigenetic effects of food components has been
fueled by observations in Agouti mice fed with a soy polyphenol diet, which revealed epigenetic changes in DNA methylation
patterns in their offspring and protected against diabetes, obesity
and cancer across multiple generations [130,131,140]. Furthermore, maternal nutrient supplementation with soy polyphenols
was found to counteract xenobiotic-induced DNA hypomethylation
in early development [90,130,138,140142]. Other experiments
revealed the potential of polyphenols to inhibit the enzymatic
activity of human or bacterial DNMT in vitro in several studies
[143,144]. In silico DNMT docking studies revealed that polyphenols
may compete with cofactor SAM for binding to the catalytic pocket

Table 1
Polyphenols as epigenetic modulators in cancer chemoprevention.
Agent

Activity

Concentration

Refs.

Green tea/Cocoa
Catechins and
Polyphenols

Decrease DNMT activity in vitro


Decrease DNMT activity in vivo
Decrease DNA methylation in human
intervention study
Decrease H3K27 methylation
Decrease protein expression Bmi1, Suz12, Ezh2
Decrease DNMT activity in vitro
Decrease DNMT activity

0.250 M
>7 cups/day

[109,143,150,224232]

520 M
2050 M

[146]
[109,233]

2 M

[147]

10 nM-100 M,
6 days
50250 mg/kg

[122,140,149,234247]

10100 mg/day 12
weeks
0.50100 M

[248251]

1 nM100 M

[145,252254]

Coffee polyphenols
Flavonoids and
polyphenols
Apple polyphenols
Genistein, soy
polyphenols

Resveratrol

Curcumin

Decrease DNMT activity


Decrease DNMT expression/mRNA
Decrease DNMT activity in vitro
Increase DNA methylation in vivo in rodents
Increase/Decrease DNA methylation in human
intervention studies
Decrease DNMT expression/mRNA
Increase/Decrease histone acetylation
Decreased HDAC6 expression
Decrease DNA methylation in human
intervention study
Decrease DNA methylation in cell lines
Increase activity SIRT1
Increase Sir2 yeast
Decrease DNMT activity
Decrease p300, HDAC1, HDAC3
Decrease p300 HAT activity

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

of DNMT [145]. However, inhibition of DNMT maybe an indirect


consequence of methylation of the catechol moiety of polyphenols by catechol-O-methyltransferase (COMT) which competes for
the methyldonor pool available for DNA methylation [143,146].
Alternatively, polyphenols may change protein expression levels
of DNMT1, DNMT3B [147149]. In addition, treatment of cell culture models with tea polyphenols led to reduced genomic 5meC
levels and promoter hypomethylation of selected candidate genes
[146,150152]. Only very few epidemiological or human intervention studies with chemopreventive polyphenols have attempted
to investige effects on DNA methylation: despite the observation
of small changes in global DNA methylation related to diet, gender, age and sex, future longitudinal studies on bigger cohorts are
required to demonstrate its signicance and whether there is a
causal link between DNA methylation changes and disease progression [122,153156].
Recently, evidence emerged that also timing (preconception,
pregnancy, lactation, neonatal life, early life, pre-/post-menopause,
puberty) of various dietary exposures may be vitally important
in determining health benecial effects, as epigenetic plasticity
changes continually from conception to death [157,158]. If intervention with chemopreventive agents is started only after critical
events have taken place, a modulating effect would likely be
missed. For example epigenetic drugs prevented intestinal neoplasia in APCMin/+ mice by >95% when intervention was started
at the age of 1 week of age, whereas the compounds were ineffective when application started at 7 weeks of age [159]. Also
studies of human populations following famine have suggested
that pathologies in later life are dependent on the critical timing of nutritional insult during pregnancy [160162]. In principle;
epigenetic changes occurring during embryonic development will
have a much greater impact on the overall epigenetic status of
the organism because, as they can be transmitted over consecutive mitotic divisions, alterations occurring in single embryonic
stem cells will affect many more cells than those occurring in
adult stem and/or somatic cells during postnatal development
[119]. In addition to epigenetic imprinting during crucial periods of development, stochastic or genetically and environmentally
triggered epigenomic changes (epimutations) occur day after day
and accumulate over time, as maximal differences in DNA methylation proles are observed in aged monozygotic twins with a
history of non-shared environments [163,164]. Although it has long
been thought that the epigenomic prole is wiped clean in the
embryo shortly after fertilization, with the exception of imprinted
genes, methylation clearing is not complete after fertilization and
on a global DNA level is reduced to 10% [165,166]. Remarkably,
recent evidence suggests that DNA methylation is rather converted
into hydroxymethylation than erased [167170]. Alternatively, it
cannot be excluded that transgenerationally inherited nutritional
effects may also depend on polycomb proteins [120,137,171,172],
miRNA proles [16,173] or epigenetic capacitor properties of hsp
proteins present in the fertilized embryo [174176].

5. Epigenetic targets of bioactive dietary components for


cancer prevention and therapy
A next challenge will be to determine which adverse epigenomic
marks in cancer-inammation are reversible or can be prevented
by specic diets, natural phytochemicals or lifestyle changes
[102,108,117,157]. Numerous botanical species and plant parts
contain a diverse array of polyphenolic phytochemicals which exert
cancer chemopreventive effects in man by its anti-inammatory,
anti-oxidant, phytohormonal, homeostatic effects (hormesis) in
immune cells and/or cancer (stem)cells [89,113,177182]. Upon
re-exploration of its biological activities, various nutritional natural

571

compounds (including epigallocatechin gallate, resveratrol, genistein, curcumin, isothiocyanates, withanolides, . . .) were found to
interfere with enzymatic activity of DNMT, Class I, II, IV HDAC,
HAT and Class III HDAC sirtuins (SIRT) which modulate cancerinammation ([99,102111,127,183], and references included).
HDACs are zinc metalloproteins which rely on Zn2+ for their activity
and are divided into 4 classes based on their homology with yeast
HDACs. Class III HDACs, called sirtuins are zinc-independent but
nicotinamide adenine dicnucleotide (NAD+ )-dependent. Class I, II,
IV HDAC inhibitors characteristically contain a Zn2+ chelating group
consisting of a thiolate, thiol, hydroxamate, carboxylate, mercaptoamide, epoxide or ketone group. Natural HDAC inhibitors can be
divided in following groups based on their chemical characteristics:
carboxylates, organosuldes, isothiocyanates, hydroamates, cyclic
tetrapeptides and macrocyclic depsipeptides [105]. In contrast to
natural HDAC inhibitors, only few natural products (i.e. niacine,
vitamin B3, dihydrocoumarin) have been identied as inhibitors
of class III HDACs. Reciprocally, various natural avonoids have
been identied as activators of class III HDACs (SIRTs). Finally,
turmeric and green tea have been identied as sources of natural inhibitors of p300/CBP HAT. Finally DNMT inhibitors work
mainly through one of the following mechanisms, either covalent
trapping of DNMT through incorporation into DNA (i.e. nucleoside
analogs decitabine, 5-azacytidine), non-covalent blocking of DNMT
catalytic active site (i.e. EGCG, parthenolide), interruption of binding site of DNMT to DNA (i.e. procaine), degradation of DNMT (i.e.;
decitabine), or suppression of DNMT expression (i.e. miRNAs). Specic epigenetic effects of natural phytochemicals may be the result
of a superposition of combined concentration dependent actions of
the compound as a nuclear hormone receptor ligand and/or modulator of histone modifying enzymes and DNMTs [184188] which
may target chromatin dynamics of specic gene clusters. Although
effects of dietary factors and extracts have frequently been demonstrated in in vitro experiments at concentrations which can never
be achieved in vivo, epigenetics sheds a more realistic light on
dietary studies, as longlife exposure at physiological concentrations
can remodel the epigenome in a cumulative fashion by repetitive effects on the epigenetic machinery [189193]. Furthermore,
it should be evaluated which epigenetic changes are stable over
time. Interestingly, even transient exposure to a specic dietary
component can induce long-lasting epigenetic changes in the promoter of the NFB subunit p65, which acts as a master switch in
cancer-inammation [194]. Alternatively, compounds may chemically interfere with histone mark interacting effector domains
(such as chromo-, bromo- or tudor domains) [195197]. Though,
one should be careful with interpretation of in vitro compound
screenings or cofactor activity assays based on peptide-protein
interactions, which may not always represent true targets in vivo
[198,199].
From another perspective, chemopreventive phytochemicals
may indirectly modulate chromatin dynamics and epigenetic
effects upon interference with global cancer metabolism. As such,
epigenetic changes may follow biochemical metabolization of
dietary factors, which can deplete cellular pools of acetyl-CoA,
NAD+ and methyldonors, resulting in unbalanced DNA methylation
and/or protein acetylation or methylation [31,183,200]. For example, avanol-rich diets interfere with the methyldonor metabolism
and the available pool of S-adenosylmethionine, resulting in
changes in DNA methylation [201205] and histone methylation, which is also affected by alterations in SAM/SAH ratios
[206,207]. Furthermore, even without nutritional deciency of
methyl groups, impaired synthesis of SAM and perturbed DNA
methylation can happen when the need for the synthesis of
the detoxication enzyme glutathione transferase (GSH) synthesis increases [208]. Diets or nutritional compounds which affect
energy metabolism or mitochondrial respiration can have global

572

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

epigenetic effects upon changes in NAD+ availability and SIRT


activity [209]. Since SIRT activation has been linked to longevity
(increased lifespan and healthy aging) and mimics a caloric
restricted diet, SIRT activators such as resveratrol represent a major
class of caloric mimetic epigenetic modulator phytochemicals
which could reverse metabolic disease [200]. Along the same line,
avanol-rich diets interfere with the methyldonor metabolism and
the available pool of S-adenosylmethionine, resulting in changes in
DNA and histone methylation [202207]. As such, specic dietary
classes of functional food maybe designed as therapeutic epigenetic
modulators in cancer-inammation.

6. Conclusion and future perspectives


The phenotype of an individual is the result of complex
geneenvironment interactions in the current, past and ancestral
environment, leading to lifelong remodeling of our epigenomes.
In recent years, several studies have demonstrated that disruption of epigenetic mechanisms can alter immune function and
contribute to various cancers. Various replication-dependent and
-independent epigenetic mechanisms are involved in developmental programming, lifelong recording of environmental changes
and transmitting transgenerational effects. It is likely that understanding and manipulating the epigenome, a potentially reversible
source of biological variation, has great potential in chemoprevention or stabilization of cancer. Much attention is currently
focusing on modulating hyper/hypomethylation of key inammatory genes by dietary factors as an effective approach to cure or
protect against cancer-inammation [102110,127,183]. In this
respect, Let food be your epigenetic medicine could represent a
novel interpretation of what Hippocrates said already 25 centuries
ago. As such, it will be a challenge for future anti-inammatory
therapeutics and preventive cancer research to identify novel
epigenetic targets which allow selective modulation of the inammatory signaling network in the diseased tumor microenvironment
[89,96,210214]. Given several encouraging trials, prevention and
therapy of age- and lifestyle-related diseases by individualized
tailoring of optimal epigenetic diets or supplements are conceivable. However, these interventions will require intense efforts
to unravel the complexity of these epigenetic, genetic and environment interactions. Another goal is to evaluate their potential
reversibility with minimal side effects as diet components may
reprogram malignant cells as well as the host immune system
and HPA-axis depending on the bioavailability of the dietary
compounds [102,191,193,215217]. There is some concern that
epigenetic therapy with dietary inhibitors of DNMT, HDAC, histon(de)methylases in longterm treatment setups may suffer from
lack of specicity [184,195,198]. As such, the possible alternative is to combine nonselective epigenetic therapies with more
targeted approaches [218]. For example, combined treatment of
specic transcription factor inhibitors and/or hormone receptor
ligands with epigenetic drugs may trigger synergistic activities at
subsets of inammatory genes [218222]. An excellent example
of cooperation between a dietary vitamin A-derivative targeting
a nuclear receptor and the HDAC inhibitor butyrate has been
described in the treatment of acute promyelocytic leukemias [103].
Finally, microRNA and long ncRNA pathways also hold promise
to join soon the arsenal of epigenetic combination therapies, as
their target sequence specicity may bridge the gap between
genetic and epigenetic changes [16,5052,59]. As such, dietary
polyphenols may indirectly modulate the epigenome by miRNAs
which target specic epigenetic modier enzymes [54,223]. In
conclusion, cancer-inammaton studies are revealing a dazzling
complexity in the mechanisms leading to dynamic alterations of
the epigenome and the need of combination therapies targeting

different chromatin modiers, to reverse disease prone epigenetic


alterations for chemoprevention. Medical benets of dietary compounds as epigenetic modulators, especially with respect to their
chronic use as nutraceutical agents, will rely on our further understanding of their epigenetic effects during embryogenesis, early life,
aging, carcinogenesis as well as through different generations.

Acknowledgements
This research is partially nancial supported by FP7 grant FLAVIOLA (www.aviola.org), an NOI-grant (UA), a Research Grant from
the Multiple Myeloma Research Foundation (MMRF) and Interuniversity Attraction Poles (IAP) P6/18. We regret that all literature
could not be appropriately cited because of space constraints and
we apologize to those authors whose original work is not mentioned.

References
[1] Manolio TA, Collins FS, Cox NJ, Goldstein DB, Hindorff LA, Hunter DJ, et al. Finding the missing heritability of complex diseases. Nature 2009;461:74753.
[2] Maher B. Personal genomes: the case of the missing heritability. Nature
2008;456:1821.
[3] Rakyan VK, Down TA, Balding DJ, Beck S. Epigenome-wide association studies
for common human diseases. Nat Rev Genet 2011;12:52941.
[4] Anand P, Kunnumakkara AB, Sundaram C, Harikumar KB, Tharakan ST, Lai OS,
et al. Cancer is a preventable disease that requires major lifestyle changes.
Pharm Res 2008;25:2097116.
[5] Chaturvedi MM, Sung B, Yadav VR, Kannappan R, Aggarwal BB. NF-kappaB
addiction and its role in cancer: one size does not t all. Oncogene
2011;30:161530.
[6] Messinaand M, Hilakivi-Clarke L. Early intake appears to be the key to the
proposed protective effects of soy intake against breast cancer. Nutr Cancer
2009;61:7928.
[7] Shu XO, Zheng Y, Cai H, Gu K, Chen Z, Zheng W, et al. Soy food intake and
breast cancer survival. JAMA 2009;302:243743.
[8] Tennant DA, Duran RV, Gottlieb E. Targeting metabolic transformation for
cancer therapy. Nat Rev Cancer 2010;10:26777.
[9] Boffetta P, Couto E, Wichmann J, Ferrari P, Trichopoulos D, Bueno-deMesquita HB, et al. Fruit and vegetable intake and overall cancer risk in the
European Prospective Investigation into Cancer and Nutrition (EPIC). J Natl
Cancer Inst 2010;102:52937.
[10] Binghamand S, Riboli E. Diet and cancer the European prospective investigation into cancer and nutrition. Nat Rev Cancer 2004;4:20615.
[11] Huang J, Plass C, Gerhauser C. Cancer chemoprevention by targeting the
epigenome. Curr Drug Targets 2011;12:192556.
[12] vel Szic KS, Ndlovu MN, Haegeman G, Vanden Berghe W. Nature or nurture: let
food be your epigenetic medicine in chronic inammatory disorders. Biochem
Pharmacol 2010;80:181632.
[13] Leeand BM, Mahadevan LC. Stability of histone modications across mammalian genomes: implications for epigenetic marking. J Cell Biochem
2009;108:2234.
[14] Vanden Berghe W, Ndlovu MN, Hoya-Arias R, Dijsselbloem N, Gerlo
S, Haegeman G. Keeping up NF-kappaB appearances: epigenetic control
of immunity or inammation-triggered epigenetics. Biochem Pharmacol
2006;72:111431.
[15] Chi P, Allis CD, Wang GG. Covalent histone modications miswritten, misinterpreted and mis-erased in human cancers. Nat Rev Cancer 2010;10:45769.
[16] Guiland S, Esteller M. DNA methylomes, histone codes and miRNAs: tying it
all together. Int J Biochem Cell Biol 2009;41:8795.
[17] Davalosand V, Esteller M. MicroRNAs and cancer epigenetics: a macrorevolution. Curr Opin Oncol 2010;22:3545.
[18] Esteller M. Cancer epigenomics: DNA methylomes and histone-modication
maps. Nat Rev Genet 2007;8:28698.
[19] Bird A. DNA methylation patterns and epigenetic memory. Genes Dev
2002;16:621.
[20] Jonesand PA, Liang G. Rethinking how DNA methylation patterns are maintained. Nat Rev Genet 2009;10:80511.
[21] Lawand JA, Jacobsen SE. Establishing, maintaining and modifying DNA methylation patterns in plants and animals. Nat Rev Genet 2010;11:20420.
[22] Wang L, Chia NC, Lu X, Ruden DM. Hypothesis: environmental regulation of
5-hydroxymethylcytosine by oxidative stress. Epigenetics 2011;6:8536.
[23] Perillo B, Ombra MN, Bertoni A, Cuozzo C, Sacchetti S, Sasso A, et al. DNA
oxidation as triggered by H3K9me2 demethylation drives estrogen-induced
gene expression. Science 2008;319:2026.
[24] Ernstand J, Kellis M. Discovery and characterization of chromatin
states for systematic annotation of the human genome. Nat Biotechnol
2010;28:81725.

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576


[25] Ernst J, Kheradpour P, Mikkelsen TS, Shoresh N, Ward LD, Epstein CB, et al.
Mapping and analysis of chromatin state dynamics in nine human cell types.
Nature 2011;473:439.
[26] Luoand J, Kuo MH. Linking nutrient metabolism to epigenetics. Cell Sci Rev
2009;6:4954.
[27] Belletand MM, Sassone-Corsi P. Mammalian circadian clock and metabolism
the epigenetic link. J Cell Sci 2010;123:383748.
[28] Chang J, Zhang B, Heath H, Galjart N, Wang X, Milbrandt J. Nicotinamide adenine dinucleotide (NAD)-regulated DNA methylation alters CCCTC-binding
factor (CTCF)/cohesin binding and transcription at the BDNF locus. Proc Natl
Acad Sci U S A 2010;107:2183641.
[29] Wallace DC. Bioenergetics and the epigenome: interface between the environment and genes in common diseases. Dev Disabil Res Rev 2010;16:1149.
[30] Wallace DC. The epigenome and the mitochondrion: bioenergetics and the
environment [corrected]. Genes Dev 2010;24:15713.
[31] Ladurner AG. Chromatin places metabolism center stage. Cell
2009;138:1820.
[32] Teperino R, Schoonjans K, Auwerx J. Histone methyl transferases and
demethylases; can they link metabolism and transcription? Cell Metab
2010;12:3217.
[33] Bonuccelli G, Tsirigos A, Whitaker-Menezes D, Pavlides S, Pestell RG, Chiavarina B, et al. Ketones and lactate fuel tumor growth and metastasis: evidence
that epithelial cancer cells use oxidative mitochondrial metabolism. Cell Cycle
2010;9:350614.
[34] Martinez-Outschoorn UE, Prisco M, Ertel A, Tsirigos A, Lin Z, Pavlides
S, et al. Ketones and lactate increase cancer cell stemness, driving
recurrence, metastasis and poor clinical outcome in breast cancer: achieving personalized medicine via Metabolo-Genomics. Cell Cycle 2011;10:
127186.
[35] Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, Shih A, et al. Leukemic
IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell
2010;18:55367.
[36] Wellen KE, Hatzivassiliou G, Sachdeva UM, Bui TV, Cross JR, Thompson CB.
ATP-citrate lyase links cellular metabolism to histone acetylation. Science
2009;324:107680.
[37] Rathmelland JC, Newgard CB. Biochemistry. A glucose-to-gene link. Science
2009;324:10212.
[38] Brower V. Epigenetics: unravelling the cancer code. Nature 2011;471:S123.
[39] Mulero-Navarroand S, Esteller M. Chromatin remodeling factor CHD5 is
silenced by promoter CpG island hypermethylation in human cancer. Epigenetics 2008;3:2105.
[40] Herceg Z. Epigenetics and cancer: towards an evaluation of the impact of
environmental and dietary factors. Mutagenesis 2007;22:91103.
[41] Grivennikovand SI, Karin M. Inammation and oncogenesis: a vicious connection. Curr Opin Genet Dev 2010;20:6571.
[42] Aggarwal BB. Inammation, a silent killer in cancer is not so silent! Curr Opin
Pharmacol 2009;9:34750.
[43] Aggarwaland BB, Gehlot P. Inammation and cancer: how friendly is the
relationship for cancer patients? Curr Opin Pharmacol 2009;9:35169.
[44] Cyrand AR, Domann FE. The redox basis of epigenetic modications:
from mechanisms to functional consequences. Antioxid Redox Signal
2011;15:55189.
[45] Esteller M. Epigenetics in cancer. N Engl J Med 2008;358:114859.
[46] Taft RJ, Pang KC, Mercer TR, Dinger M, Mattick JS. Non-coding RNAs: regulators
of disease. J Pathol 2009;220:12639.
[47] Taft RJ, Glazov EA, Cloonan N, Simons C, Stephen S, Faulkner GJ, et al. Tiny
RNAs associated with transcription start sites in animals. Nat Genet 2009;41:
5728.
[48] Mattick JS, Taft RJ, Faulkner GJ. A global view of genomic information moving
beyond the gene and the master regulator. Trends Genet 2009.
[49] Mattick JS, Amaral PP, Dinger ME, Mercer TR, Mehler MF. RNA regulation of
epigenetic processes. BioEssays 2009;31:519.
[50] Tsai MC, Manor O, Wan Y, Mosammaparast N, Wang JK, Lan F, et al. Long noncoding RNA as modular scaffold of histone modication complexes. Science
2010.
[51] Gupta RA, Shah N, Wang KC, Kim J, Horlings HM, Wong DJ, et al. Long
non-coding RNA HOTAIR reprograms chromatin state to promote cancer
metastasis. Nature 2010;464:10716.
[52] De Santa F, Barozzi I, Mietton F, Ghisletti S, Polletti S, Tusi BK, et al. A large
fraction of extragenic RNA pol II transcription sites overlap enhancers. PLoS
Biology 2010;8:e1000384.
[53] Lujambio A, Calin GA, Villanueva A, Ropero S, Sanchez-Cespedes M, Blanco D,
et al. A microRNA DNA methylation signature for human cancer metastasis.
Proc Natl Acad Sci U S A 2008;105:1355661.
[54] Parasramka MA, Ho E, Williams DE, Dashwood RH. MicroRNAs, diet, and cancer: new mechanistic insights on the epigenetic actions of phytochemicals.
Mol Carcinog 2012;51:21330.
[55] Lujambioand A, Esteller M. How epigenetics can explain human metastasis:
a new role for microRNAs. Cell Cycle 2009;8:37782.
[56] Lujambioand A, Esteller M. CpG island hypermethylation of tumor suppressor
microRNAs in human cancer. Cell Cycle 2007;6:14559.
[57] Denis H, Ndlovu MN, Fuks F. Regulation of mammalian DNA methyltransferases: a route to new mechanisms. EMBO Rep 2011;12:64756.
[58] Ndlovu MN, Denis H, Fuks F. Exposing the DNA methylome iceberg. Trends
Biochem Sci 2011.

573

[59] Parasramka MA, Ho E, Williams DE, Dashwood RH. MicroRNAs, diet, and cancer: new mechanistic insights on the epigenetic actions of phytochemicals.
Mol Carcin 2011.
[60] Mahfouz MM. RNA-directed DNA methylation: mechanisms and functions.
Plant Signal Behav 2010;5.
[61] Backdahl L, Bushell A, Beck S. Inammatory signalling as mediator of epigenetic modulation in tissue-specic chronic inammation. Int J Biochem Cell
Biol 2009;41:17684.
[62] Hesson LB, Hitchins MP, Ward RL. Epimutations and cancer predisposition:
importance and mechanisms. Curr Opin Genet Dev 2010;20:2908.
[63] Laiand AY, Wade PA. Cancer biology and NuRD: a multifaceted chromatin
remodelling complex. Nat Rev Cancer 2011.
[64] Elsasser SJ, Allis CD, Lewis PW. Cancer. New epigenetic drivers of cancers.
Science 2011;331:11456.
[65] Varela I, Tarpey P, Raine K, Huang D, Ong CK, Stephens P, et al. Exome sequencing identies frequent mutation of the SWI/SNF complex gene PBRM1 in renal
carcinoma. Nature 2011;469:53942.
[66] Dalgliesh GL, Furge K, Greenman C, Chen L, Bignell G, Butler A, et al. Systematic sequencing of renal carcinoma reveals inactivation of histone modifying
genes. Nature 2010;463:3603.
[67] Ko M, Huang Y, Jankowska AM, Pape UJ, Tahiliani M, Bandukwala HS, et al.
Impaired hydroxylation of 5-methylcytosine in myeloid cancers with mutant
TET2. Nature 2010;468:83943.
[68] Delhommeau F, Dupont S, Della Valle V, James C, Trannoy S, Masse A,
et al. Mutation in TET2 in myeloid cancers. N Engl J Med 2009;360:
2289301.
[69] Berdascoand M, Esteller M. Aberrant epigenetic landscape in cancer: how
cellular identity goes awry. Dev Cell 2010;19:698711.
[70] Raney BJ, Cline MS, Rosenbloom KR, Dreszer TR, Learned K, Barber GP, et al.
ENCODE whole-genome data in the UCSC genome browser (2011 update).
Nucleic Acids Res 2011;39:D8715.
[71] Myers RM, Stamatoyannopoulos J, Snyder M, Dunham I, Hardison RC, Bernstein BE, et al. A users guide to the encyclopedia of DNA elements (ENCODE).
PLoS Biol 2011;9:e1001046.
[72] Birney E, Stamatoyannopoulos JA, Dutta A, Guigo R, Gingeras TR, Margulies
EH, et al. Identication and analysis of functional elements in 1% of the human
genome by the ENCODE pilot project. Nature 2007;447:799816.
[73] Hammand CA, Costa FF. The impact of epigenomics on future drug design and
new therapies. Drug Discov Today 2011;16:62635.
[74] Natoli G. Maintaining cell identity through global control of genomic organization. Immunity 2010;33:1224.
[75] Natoli G, Ghisletti S, Barozzi I. The genomic landscapes of inammation. Genes
Dev 2011;25:1016.
[76] Hakim O, Sung MH, Voss TC, Splinter E, John S, Sabo PJ, et al. Diverse gene
reprogramming events occur in the same spatial clusters of distal regulatory
elements. Genome Res 2011;21:697706.
[77] Tolhuis B, Blom M, Kerkhoven RM, Pagie L, Teunissen H, Nieuwland M, et al.
Interactions among polycomb domains are guided by chromosome architecture. PLoS Genet 2011;7:e1001343.
[78] van Steensel B. Chromatin: constructing the big picture. EMBO J
2011;30:188595.
[79] Mercurio C, Minucci S, Pelicci PG. Histone deacetylases and epigenetic therapies of hematological malignancies. Pharmacol Res 2010;62:1834.
[80] Hanahanand D, Weinberg RA. Hallmarks of cancer: the next generation. Cell
2011;144:64674.
[81] Mantovani A, Allavena P, Sica A, Balkwill F. Cancer-related inammation.
Nature 2008;454:43644.
[82] Grivennikov SI, Greten FR, Karin M. Immunity, inammation, and cancer. Cell
2010;140:88399.
[83] Iliopoulos D, Hirsch HA, Wang G, Struhl K. Inducible formation of breast cancer
stem cells and their dynamic equilibrium with non-stem cancer cells via IL6
secretion. Proc Natl Acad Sci U S A 2011;108:1397402.
[84] Sharma SV, Lee DY, Li B, Quinlan MP, Takahashi F, Maheswaran S, et al. A
chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell 2010;141:6980.
[85] Liu M, Sakamaki T, Casimiro MC, Willmarth NE, Quong AA, Ju X, et al.
The canonical NF-kappaB pathway governs mammary tumorigenesis in
transgenic mice and tumor stem cell expansion. Cancer Res 2010;70:
1046473.
[86] Rajasekhar VK, Studer L, Gerald W, Socci ND, Scher HI. Tumour-initiating
stem-like cells in human prostate cancer exhibit increased NF-kappaB signalling. Nat Commun 2011;2:162.
[87] Sporn MB. Perspective: the big C for chemoprevention. Nature
2011;471:S101.
[88] Deocaris CC, Widodo N, Wadhwa R, Kaul SC. Merger of ayurveda and tissue culture-based functional genomics: inspirations from systems biology. J
Transl Med 2008;6:14.
[89] Surh YJ. Cancer chemoprevention with dietary phytochemicals. Nat Rev Cancer 2003;3:76880.
[90] Jirtleand RL, Skinner MK. Environmental epigenomics and disease susceptibility. Nat Rev Genet 2007;8:25362.
[91] Singh S. From exotic spice to modern drug? Cell 2007;130:7658.
[92] Harvey A. Natural products in drug discovery. Drug Discov Today
2008;13:894901.
[93] Liand JWH, Vederas JC. Drug discovery and natural products: end of an era or
an endless frontier? Science 2009;325:1615.

574

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

[94] Lee KW, Bode AM, Dong Z. Molecular targets of phytochemicals for cancer
prevention. Nat Rev Cancer 2011;11:2118.
[95] Aggarwal BB, Prasad S, Reuter S, Kannappan R, Yadev VR, Park B, et al.
Identication of novel anti-inammatory agents from ayurvedic medicine
for prevention of chronic diseases: reverse pharmacology and bedside to
bench approach. Curr Drug Targets 2011.
[96] Deorukhkar A, Krishnan S, Sethi G, Aggarwal BB. Back to basics: how natural
products can provide the basis for new therapeutics. Expert Opin Invest Drugs
2007;16:175373.
[97] Kok TM, Breda SG, Manson MM. Mechanisms of combined action of different
chemopreventive dietary compounds. Eur J Nutr 2008;47:519.
[98] Meeran SM, Ahmed A, Tollefsbol TO. Epigenetic targets of bioactive
dietary components for cancer prevention and therapy. Clinical Epigenetics
2010;1:10116.
[99] Szarc vel Szic K, Ndlovu MN, Haegeman G, Vanden Berghe W. Nature or nurture: let food be your epigenetic medicine in chronic inammatory disorders.
Biochem Pharmacol 2010;80:181632.
[100] Iliopoulos D, Hirsch HA, Struhl K. An epigenetic switch involving NF-kappaB,
Lin28, Let-7 MicroRNA, and IL6 links inammation to cell transformation. Cell
2009;139:693706.
[101] Iliopoulos D, Jaeger SA, Hirsch HA, Bulyk ML, Struhl K. STAT3 activation of
miR-21 and miR-181b-1 via PTEN and CYLD are part of the epigenetic switch
linking inammation to cancer. Mol Cell 2010;39:493506.
[102] Burdgeand GC, Lillycrop KA. Nutrition, epigenetics, and developmental plasticity: implications for understanding human disease. Annu Rev Nutr 2010.
[103] Delageand B, Dashwood RH. Dietary manipulation of histone structure and
function. Annu Rev Nutr 2008;28:34766.
[104] Link A, Balaguer F, Goel A. Cancer chemoprevention by dietary polyphenols:
promising role for epigenetics. Biochem Pharmacol 2010.
[105] Folmer F, Orlikova B, Schnekenburger M, Dicato M, Diederich M. Naturally
occurring regulators of histone acetylation/deacetylation. Curr Nutr Food Sci
2010;6:7899.
[106] Hauserand AT, Jung M. Targeting epigenetic mechanisms: potential of natural
products in cancer chemoprevention. Planta Med 2008;74:1593601.
[107] Kontogiorgis C, Bompou E, Ntella M, Vanden Berghe W. Natural products
from mediterranean diet: from anti-inammatory agents to dietary epigenetic modulators. Anti-Inamm Anti-Allergy Agents Med Chem 2010;6:
10124.
[108] Kirk H, Cefalu WT, Ribnicky D, Liu Z, Eilertsen KJ. Botanicals as epigenetic modulators for mechanisms contributing to development of metabolic syndrome.
Metab Clin Exp 2008;57:S1623.
[109] Fang M, Chen D, Yang CS. Dietary polyphenols may affect DNA methylation. J
Nutr 2007;137:223S8S.
[110] Suzukiand T, Miyata N. Epigenetic control using natural products and synthetic molecules. Curr Med Chem 2006;13:93558.
[111] Arasaradnam RP, Commane DM, Bradburn D, Mathers JC. A review of dietary
factors and its inuence on DNA methylation in colorectal carcinogenesis.
Epigenetics 2008;3:1938.
[112] Hurtand EM, Farrar WL. Cancer stem cells: the seeds of metastasis? Mol Interv
2008;8:1402.
[113] Kawasaki BT, Hurt EM, Mistree T, Farrar WL. Targeting cancer stem cells with
phytochemicals. Mol Interv 2008;8:17484.
[114] Dashwoodand RH, Ho E. Dietary histone deacetylase inhibitors: from cells to
mice to man. Semin Cancer Biol 2007;17:3639.
[115] Dashwood RH, Myzak MC, Ho E. Dietary HDAC inhibitors: time to rethink
weak ligands in cancer chemoprevention? Carcinogenesis 2006;27:3449.
[116] Cooney CA. Germ cells carry the epigenetic benets of grandmothers diet.
Proc Natl Acad Sci U S A 2006;103:170712.
[117] Godfrey KM, Gluckman PD, Hanson MA. Developmental origins of metabolic
disease: life course and intergenerational perspectives. Trends Endocrinol
Metab 2010;21:199205.
[118] Weaver IC. Shaping adult phenotypes through early life environments. Birth
Defects Res C: Embryo Today 2009;87:31426.
[119] Aguilera O, Fernandez AF, Munoz A, Fraga MF. Epigenetics and environment:
a complex relationship. J Appl Physiol 2010;109:24351.
[120] Youngsonand NA, Whitelaw E. Transgenerational epigenetic effects. Annu Rev
Genomics Hum Genet 2008;9:23357.
[121] Gallou-Kabani C, Vige A, Gross MS, Junien C. Nutri-epigenomics: lifelong
remodelling of our epigenomes by nutritional and metabolic factors and
beyond. Clin Chem Lab Med 2007;45:3217.
[122] Howard TD, Ho SM, Zhang L, Chen J, Cui W, Slager R, et al. Epigenetic changes
with dietary soy in cynomolgus monkeys. PLoS ONE 2011;6:e26791.
[123] Gluckman PD, Hanson MA, Cooper C, Thornburg KL. Effect of in utero
and early-life conditions on adult health and disease. N Engl J Med
2008;359:6173.
[124] Anway MD, Cupp AS, Uzumcu M, Skinner MK. Epigenetic transgenerational
actions of endocrine disruptors and male fertility. Science 2005;308:14669.
[125] Anwayand MD, Skinner MK. Epigenetic transgenerational actions of
endocrine disruptors. Endocrinology 2006;147:S439.
[126] Barkerand DJ, Martyn CN. The maternal and fetal origins of cardiovascular
disease. J Epidemiol Community Health 1992;46:811.
[127] Jackson AA, Burdge GC, Lillycrop KA. Diet, nutrition and modulation of
genomic expression in fetal origins of adult disease. World Rev Nutr Diet
2010;101:5672.
[128] Chmurzynska A. Fetal programming: link between early nutrition, DNA
methylation, and complex diseases. Nutr Rev 2010;68:8798.

[129] Hochberg Z, Feil R, Constancia M, Fraga M, Junien C, Carel JC, et al. Child
health, developmental plasticity, and epigenetic programming. Endocr Rev
2011;32:159224.
[130] Dolinoyand DC, Jirtle RL. Environmental epigenomics in human health and
disease. Environ Mol Mutagen 2008;49:48.
[131] Waterland RA. Is epigenetics an important link between early life events and
adult disease? Horm Res 2009;71(Suppl. 1):136.
[132] Waterland RA, Travisano M, Tahiliani KG, Rached MT, Mirza S. Methyl donor
supplementation prevents transgenerational amplication of obesity. Int J
Obes (Lond) 2008;32:13739.
[133] Waterlandand RA, Jirtle RL. Early nutrition, epigenetic changes at transposons
and imprinted genes, and enhanced susceptibility to adult chronic diseases.
Nutrition 2004;20:638.
[134] Kaminsky ZA, Tang T, Wang SC, Ptak C, Oh GH, Wong AH, et al. DNA methylation proles in monozygotic and dizygotic twins. Nat Genet 2009;41:2405.
[135] Petronis A. Epigenetics and twins: three variations on the theme. Trends
Genet 2006;22:34750.
[136] Belland JT, Spector TD. A twin approach to unraveling epigenetics. Trends
Genet 2011;27:11625.
[137] Chong S, Youngson NA, Whitelaw E. Heritable germline epimutation is not the
same as transgenerational epigenetic inheritance. Nat Genet 2007;39:5745,
author reply 5756.
[138] Skinner MK, Manikkam M, Guerrero-Bosagna C. Epigenetic transgenerational
actions of endocrine disruptors. Reprod Toxicol 2011;31:33743.
[139] Hochberg Z, Feil R, Constancia M, Fraga M, Junien C, Carel JC, et al. Child
health, developmental plasticity, and epigenetic programming. Endocr Rev
2010;32:159224.
[140] Dolinoy DC, Weidman JR, Waterland RA, Jirtle RL. Maternal genistein alters
coat color and protects Avy mouse offspring from obesity by modifying the
fetal epigenome. Environ Health Perspect 2006;114:56772.
[141] Kujjo LL, Chang EA, Pereira RJ, Dhar S, Marrero-Rosado B, Sengupta S, et al.
Chemotherapy-induced late transgenerational effects in mice. PLoS ONE
2011;6:e17877.
[142] Dolinoy DC, Huang D, Jirtle RL. Maternal nutrient supplementation counteracts bisphenol A-induced DNA hypomethylation in early development. Proc
Natl Acad Sci U S A 2007;104:1305661.
[143] Lee WJ, Shim JY, Zhu BT. Mechanisms for the inhibition of DNA methyltransferases by tea catechins and bioavonoids. Mol Pharmacol 2005;68:101830.
[144] Gilbertand ER, Liu D. Flavonoids inuence epigenetic-modifying enzyme
activity: structurefunction relationships and the therapeutic potential for
cancer. Curr Med Chem 2010;17:175668.
[145] Liu Z, Xie Z, Jones W, Pavlovicz RE, Liu S, Yu J, et al. Curcumin is a potent DNA
hypomethylation agent. Bioorg Med Chem Lett 2009;19:7069.
[146] Leeand WJ, Zhu BT. Inhibition of DNA methylation by caffeic acid and
chlorogenic acid, two common catechol-containing coffee polyphenols. Carcinogenesis 2006;27:26977.
[147] Fini L, Selgrad M, Fogliano V, Graziani G, Romano M, Hotchkiss E, et al.
Annurca apple polyphenols have potent demethylating activity and can reactivate silenced tumor suppressor genes in colorectal cancer cells. J Nutr
2007;137:26228.
[148] Gerhauser C. Cancer chemopreventive potential of apples, apple juice, and
apple components. Planta Med 2008;74:160824.
[149] Li Y, Liu L, Andrews LG, Tollefsbol TO. Genistein depletes telomerase activity
through cross-talk between genetic and epigenetic mechanisms. Int J Cancer
2009;125:28696.
[150] Fang MZ, Wang Y, Ai N, Hou Z, Sun Y, Lu H, et al. Tea polyphenol
()-epigallocatechin-3-gallate inhibits DNA methyltransferase and reactivates methylation-silenced genes in cancer cell lines. Cancer Res 2003;63:
756370.
[151] Shu L, Khor TO, Lee JH, Boyanapalli SS, Huang Y, Wu TY, et al. Epigenetic CpG
demethylation of the promoter and reactivation of the expression of Neurog1
by curcumin in prostate LNCaP cells. AAPS J 2011;13:60614.
[152] Khor TO, Huang Y, Wu TY, Shu L, Lee J, Kong AN. Pharmacodynamics of curcumin as DNA hypomethylation agent in restoring the expression of Nrf2 via
promoter CpGs demethylation. Biochem Pharmacol 2011;82:10738.
[153] Zhang FF, Morabia A, Carroll J, Gonzalez K, Fulda K, Kaur M, et al. Dietary
patterns are associated with levels of global genomic DNA methylation in a
cancer-free population. J Nutr 2011;141:116571.
[154] Scoccianti C, Ricceri F, Ferrari P, Cuenin C, Sacerdote C, Polidoro S, et al.
Methylation patterns in sentinel genes in peripheral blood cells of heavy
smokers: Inuence of cruciferous vegetables in an intervention study. Epigenetics 2011;6.
[155] Khulan B, Cooper WN, Skinner BM, Bauer J, Owens S, Prentice AM, et al.
Periconceptional maternal micronutrient supplementation is associated with
widespread gender related changes in the epigenome: a study of a unique
resource in the Gambia. Hum Mol Genet 2012.
[156] Cooper WN, Khulan B, Owens S, Elks CE, Seidel V, Prentice AM, et al. DNA
methylation proling at imprinted loci after periconceptional micronutrient
supplementation in humans: results of a pilot randomized controlled trial.
FASEB J 2012.
[157] Burdge GC, Lillycrop KA, Jackson AA. Nutrition in early life, and risk of cancer
and metabolic disease: alternative endings in an epigenetic tale? Br J Nutr
2009;101:61930.
[158] Faulkand C, Dolinoy DC. Timing is everything: the when and how of environmentally induced changes in the epigenome of animals. Epigenetics
2011;6:7917.

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576


[159] Laird PW, Jackson-Grusby L, Fazeli A, Dickinson SL, Jung WE, Li E,
et al. Suppression of intestinal neoplasia by DNA hypomethylation. Cell
1995;81:197205.
[160] Painter RC, Osmond C, Gluckman P, Hanson M, Phillips DI, Roseboom TJ. Transgenerational effects of prenatal exposure to the Dutch famine on neonatal
adiposity and health in later life. BJOG 2008;115:12439.
[161] Lumeyand LH, Stein AD. Transgenerational effects of prenatal exposure to the
Dutch famine. BJOG 2009;116:868, author reply 868.
[162] Roseboom T, de Rooij S, Painter R. The Dutch famine and its long-term consequences for adult health. Early Hum Dev 2006;82:48591.
[163] Fraga MF, Ballestar E, Paz MF, Ropero S, Setien F, Ballestar ML, et al. Epigenetic
differences arise during the lifetime of monozygotic twins. Proc Natl Acad Sci
U S A 2005;102:106049.
[164] Christensen BC, Houseman EA, Marsit CJ, Zheng S, Wrensch MR, Wiemels JL,
et al. Aging and environmental exposures alter tissue-specic DNA methylation dependent upon CpG island context. PLoS Genet 2009;5:e1000602.
[165] Surani MA, Ancelin K, Hajkova P, Lange UC, Payer B, Western P, et al.
Mechanism of mouse germ cell specication: a genetic program regulating epigenetic reprogramming. Cold Spring Harb Symp Quant Biol 2004;69:
19.
[166] Hajkova P, Ancelin K, Waldmann T, Lacoste N, Lange UC, Cesari F, et al. Chromatin dynamics during epigenetic reprogramming in the mouse germ line.
Nature 2008;452:87781.
[167] Ficz G, Branco MR, Seisenberger S, Santos F, Krueger F, Hore TA, et al. Dynamic
regulation of 5-hydroxymethylcytosine in mouse ES cells and during differentiation. Nature 2011;473:398402.
[168] Wossidlo M, Nakamura T, Lepikhov K, Marques CJ, Zakhartchenko V, Boiani
M, et al. 5-Hydroxymethylcytosine in the mammalian zygote is linked with
epigenetic reprogramming. Nat Commun 2011;2:241.
[169] Wossidlo M, Arand J, Sebastiano V, Lepikhov K, Boiani M, Reinhardt R, et al.
Dynamic link of DNA demethylation, DNA strand breaks and repair in mouse
zygotes. EMBO J 2010;29:187788.
[170] Iqbal K, Jin SG, Pfeifer GP, Szabo PE. Reprogramming of the paternal genome
upon fertilization involves genome-wide oxidation of 5-methylcytosine. Proc
Natl Acad Sci U S A 2011;108:36427.
[171] Blewitt ME, Vickaryous NK, Paldi A, Koseki H, Whitelaw E. Dynamic reprogramming of DNA methylation at an epigenetically sensitive allele in mice.
PLoS Genet 2006;2:e49.
[172] Brackenand AP, Helin K. Polycomb group proteins: navigators of lineage pathways led astray in cancer. Nat Rev Cancer 2009;9:77384.
[173] Li Y, Kong D, Wang Z, Sarkar FH. Regulation of microRNAs by natural agents: an
emerging eld in chemoprevention and chemotherapy research. Pharmaceut
Res 2010;27:102741.
[174] Ruden DM, Xiao L, Garnkel MD, Lu X. Hsp90 and environmental impacts on
epigenetic states: a model for the trans-generational effects of diethylstibesterol on uterine development and cancer. Hum Mol Genet 2005;14(Spec no
1):R14955.
[175] Sollars V, Lu X, Xiao L, Wang X, Garnkel MD, Ruden DM. Evidence for an
epigenetic mechanism by which Hsp90 acts as a capacitor for morphological
evolution. Nat Genet 2003;33:704.
[176] Ruden DM, De Luca M, Garnkel MD, Bynum KL, Lu X. Drosophila nutrigenomics can provide clues to human genenutrient interactions. Annu Rev
Nutr 2005;25:499522.
[177] Dijsselbloem N, Vanden Berghe W, De Naeyer A, Haegeman G. Soy
isoavone phyto-pharmaceuticals in interleukin-6 affections: multi-purpose
nutraceuticals at the crossroad of hormone replacement, anti-cancer and
anti-inammatory therapy. Biochem Pharmacol 2004;68:117185.
[178] Blanpainand C, Fuchs E. Epidermal homeostasis: a balancing act of stem cells
in the skin. Nat Rev Mol Cell Biol 2009;10:20717.
[179] Crea F, Mathews LA, Farrar WL, Hurt EM. Targeting prostate cancer stem cells.
Anticancer Agents Med Chem 2009;9:110513.
[180] Shytle RD, Ehrhart J, Tan J, Vila J, Cole M, Sanberg CD, et al. Oxidative stress
of neural, hematopoietic, and stem cells: protection by natural compounds.
Rejuvenation Res 2007;10:1738.
[181] Bickford PC, Tan J, Shytle RD, Sanberg CD, El-Badri N, Sanberg PR. Nutraceuticals synergistically promote proliferation of human stem cells. Stem Cells
Dev 2006;15:11823.
[182] Zhou J, Zhang H, Gu P, Bai J, Margolick JB, Zhang Y. NF-kappaB pathway
inhibitors preferentially inhibit breast cancer stem-like cells. Breast Cancer
Res Treat 2008;111:41927.
[183] Vaqueroand A, Reinberg D. Calorie restriction and the exercise of chromatin.
Genes Dev 2009;23:184969.
[184] Mai A, Cheng D, Bedford MT, Valente S, Nebbioso A, Perrone A, et al.
Epigenetic multiple ligands: mixed histone/protein methyltransferase,
acetyltransferase, and class III deacetylase (sirtuin) inhibitors. J Med Chem
2008;51:227990.
[185] Kuniyasu H. The roles of dietary PPARgamma ligands for metastasis in colorectal cancer. PPAR research 2008;2008:529720.
[186] Denisonand MS, Nagy SR. Activation of the aryl hydrocarbon receptor by
structurally diverse exogenous and endogenous chemicals. Annu Rev Pharmacol Toxicol 2003;43:30934.
[187] Darbreand PD, Charles AK. Environmental oestrogens and breast cancer:
evidence for combined involvement of dietary, household and cosmetic
xenoestrogens. Anticancer Res 2010;30:81527.
[188] Newbold RR, Padilla-Banks E, Jefferson WN. Environmental estrogens and
obesity. Mol Cell Endocrinol 2009;304:849.

575

[189] Manachand C, Donovan JL. Pharmacokinetics and metabolism of dietary


avonoids in humans. Free Radical Res 2004;38:77185.
[190] Manach C, Mazur A, Scalbert A. Polyphenols and prevention of cardiovascular
diseases. Curr Opin Lipidol 2005;16:7784.
[191] Manach C, Williamson G, Morand C, Scalbert A, Remesy C. Bioavailability and
bioefcacy of polyphenols in humans. I. Review of 97 bioavailability studies.
Am J Clin Nutr 2005;81:230S42S.
[192] Manach C, Scalbert A, Morand C, Remesy C, Jimenez L. Polyphenols: food
sources and bioavailability. Am J Clin Nutr 2004;79:72747.
[193] Williamsonand G, Manach C. Bioavailability and bioefcacy of polyphenols in humans. II. Review of 93 intervention studies. Am J Clin Nutr
2005;81:243S55S.
[194] El-Osta A, Brasacchio D, Yao D, Pocai A, Jones PL, Roeder RG, et al. Transient
high glucose causes persistent epigenetic changes and altered gene expression during subsequent normoglycemia. J Exp Med 2008;205:240917.
[195] Zheng YG, Wu J, Chen Z, Goodman M. Chemical regulation of epigenetic modications: opportunities for new cancer therapy. Med Res Rev
2008;28:64587.
[196] Seet BT, Dikic I, Zhou MM, Pawson T. Reading protein modications with
interaction domains. Nat Rev Mol Cell Biol 2006;7:47383.
[197] Wigle TJ, Herold JM, Senisterra GA, Vedadi M, Kireev DB, Arrowsmith CH, et al.
Screening for inhibitors of low-afnity epigenetic peptide-protein interactions: an AlphaScreen-based assay for antagonists of methyl-lysine binding
proteins. J Biomol Screen 2010;15:6271.
[198] Altucciand L, Minucci S. Epigenetic therapies in haematological malignancies:
searching for true targets. Eur J Cancer 2009;45:113745.
[199] Pacholec M, Bleasdale JE, Chrunyk B, Cunningham D, Flynn D, Garofalo RS,
et al. SRT1720, SRT2183, SRT1460, and resveratrol are not direct activators of
SIRT1. J Biol Chem 2010;285:834051.
[200] Imaiand S, Guarente L. Ten years of NAD-dependent SIR2 family deacetylases:
implications for metabolic diseases. Trends Pharmacol Sci 2010;31:21220.
[201] Chen NC, Yang F, Capecci LM, Gu Z, Schafer AI, Durante W, et al. Regulation
of homocysteine metabolism and methylation in human and mouse tissues.
FASEB J 2010.
[202] Bistul G, Vandette E, Matsui S, Smiraglia DJ. Mild folate deciency induces
genetic and epigenetic instability and phenotype changes in prostate cancer
cells. BMCBiol 2010;8:6.
[203] Ghoshal K, Li X, Datta J, Bai S, Pogribny I, Pogribny M, et al. A folate- and
methyl-decient diet alters the expression of DNA methyltransferases and
methyl CpG binding proteins involved in epigenetic gene silencing in livers
of F344 rats. J Nutr 2006;136:15227.
[204] Kim JM, Hong K, Lee JH, Lee S, Chang N. Effect of folate deciency on
placental DNA methylation in hyperhomocysteinemic rats. J Nutr Biochem
2009;20:1726.
[205] Ulrich CM, Reed MC, Nijhout HF. Modeling folate, one-carbon metabolism,
and DNA methylation. Nutr Rev 2008;66(Suppl 1):S2730.
[206] Sharma P, Senthilkumar RD, Brahmachari V, Sundaramoorthy E, Mahajan
A, Sharma A, et al. Mining literature for a comprehensive pathway analysis: a case study for retrieval of homocysteine related genes for genetic and
epigenetic studies. Lipids Health Dis 2006;5:1.
[207] Pogribny IP, Tryndyak VP, Muskhelishvili L, Rusyn I, Ross SA. Methyl deciency, alterations in global histone modications, and carcinogenesis. J Nutr
2007;137:216S22S.
[208] Lee DH, Jacobs Jr DR, Porta M. Hypothesis: a unifying mechanism for nutrition and chemicals as lifelong modulators of DNA hypomethylation. Environ
Health Perspect 2009;117:1799802.
[209] Whittle JR, Powell MJ, Popov VM, Shirley LA, Wang C, Pestell RG. Sirtuins,
nuclear hormone receptor acetylation and transcriptional regulation. Trends
Endocrinol Metab 2007;18:35664.
[210] Paul AT, Gohil VM, Bhutani KK. Modulating TNF-alpha signaling with natural
products. Drug Discov Today 2006;11:72532.
[211] Rios JL, Recio MC, Escandell JM, Andujar I. Inhibition of transcription factors by
plant-derived compounds and their implications in inammation and cancer.
Curr Pharm Des 2009;15:121237.
[212] Khanna D, Sethi G, Ahn KS, Pandey MK, Kunnumakkara AB, Sung B, et al.
Natural products as a gold mine for arthritis treatment. Curr Opin Pharmacol
2007;7:34451.
[213] Bremnerand P, Heinrich M. Natural products as targeted modulators of the
nuclear factor-kappaB pathway. J Pharm Pharmacol 2002;54:45372.
[214] Karin M, Yamamoto Y, Wang QM. The IKK NF-kappa B system: a treasure trove
for drug development. Nat Rev Drug Discov 2004;3:1726.
[215] Dijsselbloem N, Goriely S, Albarani V, Gerlo S, Francoz S, Marine JC, et al.
A critical role for p53 in the control of NF-kappaB-dependent gene expression in TLR4-stimulated dendritic cells exposed to Genistein. J Immunol
2007;178:504857.
[216] Vanden Berghe W, Dijsselbloem N, Vermeulen L, Ndlovu N, Boone E, Haegeman G. Attenuation of mitogen- and stress-activated protein kinase-1-driven
nuclear factor-{kappa}B gene expression by soy isoavones does not require
estrogenic activity. Cancer Res 2006;66:485262.
[217] Ndlovu N, Van Lint C, Van Wesemael K, Callebert P, Chalbos D, Haegeman
G, et al. Hyperactivated NF-{kappa}B and AP-1 transcription factors promote highly accessible chromatin and constitutive transcription across the
interleukin-6 gene promoter in metastatic breast cancer cells. Mol Cell Biol
2009;29:5488504.
[218] Hervouet E, Vallette FM, Cartron PF. Dnmt3/transcription factor interactions
as crucial players in targeted DNA methylation. Epigenetics 2009;4:48799.

576

W. Vanden Berghe / Pharmacological Research 65 (2012) 565576

[219] Perissi V, Jepsen K, Glass CK, Rosenfeld MG. Deconstructing repression: evolving models of co-repressor action. Nat Rev Genet 2010;11:10923.
[220] Biddie SC, John S, Hager GL. Genome-wide mechanisms of nuclear receptor
action. Trends Endocrinol Metab 2010;21:39.
[221] Fiskus W, Wang Y, Sreekumar A, Buckley KM, Shi H, Jillella A, et al. Combined epigenetic therapy with the histone methyltransferase EZH2 inhibitor,
3-deazaneplanocin A and the histone deacetylase inhibitor panobinostat
against human AML cells. Blood 2009;114:273343.
[222] Di Croce L, Raker VA, Corsaro M, Fazi F, Fanelli M, Faretta M, et al. Methyltransferase recruitment and DNA hypermethylation of target promoters by
an oncogenic transcription factor. Science 2002;295:107982.
[223] Milenkovic D, Deval C, Gouranton E, Landrier JF, Scalbert A, Morand C,
et al. Modulation of miRNA expression by dietary polyphenols in apoE
decient mice: a new mechanism of the action of polyphenols. PLoS ONE
2012;7:e29837.
[224] Chuang JC, Yoo CB, Kwan JM, Li TW, Liang G, Yang AS, et al. Comparison of
biological effects of non-nucleoside DNA methylation inhibitors versus, 5aza-2 -deoxycytidine. Mol Cancer Ther 2005;4:151520.
[225] Stresemann C, Brueckner B, Musch T, Stopper H, Lyko F. Functional diversity
of DNA methyltransferase inhibitors in human cancer cell lines. Cancer Res
2006;66:2794800.
[226] Berletch JB, Liu C, Love WK, Andrews LG, Katiyar SK, Tollefsbol TO. Epigenetic
and genetic mechanisms contribute to telomerase inhibition by EGCG. J Cell
Biochem 2008;103:50919.
[227] Kato K, Long NK, Makita H, Toida M, Yamashita T, Hatakeyama D, et al. Effects
of green tea polyphenol on methylation status of RECK gene and cancer cell
invasion in oral squamous cell carcinoma cells. Br J Cancer 2008;99:64754.
[228] Gao Z, Xu Z, Hung MS, Lin YC, Wang T, Gong M, et al. Promoter demethylation
of WIF-1 by epigallocatechin-3-gallate in lung cancer cells. Anticancer Res
2009;29:202530.
[229] Pandey M, Shukla S, Gupta S. Promoter demethylation and chromatin remodeling by green tea polyphenols leads to re-expression of GSTP1 in human
prostate cancer cells. Int J Cancer 2010;126:252033.
[230] Volate SR, Muga SJ, Issa AY, Nitcheva D, Smith T, Wargovich MJ. Epigenetic modulation of the retinoid X receptor alpha by green tea in the
azoxymethane-Apc Min/+ mouse model of intestinal cancer. Mol Carcinog
2009;48:92033.
[231] Yuasa Y, Nagasaki H, Akiyama Y, Hashimoto Y, Takizawa T, Kojima K, et al. DNA
methylation status is inversely correlated with green tea intake and physical
activity in gastric cancer patients. Int J Cancer 2009;124:267782.
[232] Balasubramanian S, Adhikary G, Eckert RL. The Bmi-1 polycomb protein
antagonizes the ()-epigallocatechin-3-gallate-dependent suppression of
skin cancer cell survival. Carcinogenesis 2010;31:496503.
[233] Paluszczak J, Krajka-Kuzniak V, Baer-Dubowska W. The effect of dietary
polyphenols on the epigenetic regulation of gene expression in MCF7 breast
cancer cells. Toxicol Lett 2010;192:11925.
[234] King-Batoon A, Leszczynska JM, Klein CB. Modulation of gene methylation by genistein or lycopene in breast cancer cells. Environ Mol Mutagen
2008;49:3645.
[235] Spurling CC, Suhl JA, Boucher N, Nelson CE, Rosenberg DW, Giardina C. The
short chain fatty acid butyrate induces promoter demethylation and reactivation of RARbeta2 in colon cancer cells. Nutr Cancer 2008;60:692702.
[236] Fang MZ, Chen D, Sun Y, Jin Z, Christman JK, Yang CS. Reversal of hypermethylation and reactivation of p16INK4a, RARbeta, and MGMT genes by genistein
and other isoavones from soy. Clin Cancer Res 2005;11:703341.
[237] Majid S, Dar AA, Ahmad AE, Hirata H, Kawakami K, Shahryari V, et al. BTG3
tumor suppressor gene promoter demethylation, histone modication and
cell cycle arrest

[238] Vardi A, Bosviel R, Rabiau N, Adjakly M, Satih S, Dechelotte P, et al. Soy phytoestrogens modify DNA methylation of GSTP1, RASSF1A, EPH2 and BRCA1
promoter in prostate cancer cells. In Vivo 2010;24:393400.
[239] Morey Kinney SR, Zhang W, Pascual M, Greally JM, Gillard BM, Karasik E, et al.
Lack of evidence for green tea polyphenols as DNA methylation inhibitors in
murine prostate. Cancer Prev Res (Phila) 2009;2:106575.
[240] Day JK, Bauer AM, DesBordes C, Zhuang Y, Kim BE, Newton LG, et al. Genistein
alters methylation patterns in mice. J Nutr 2002;132:2419S23S.
[241] Tang WY, Newbold R, Mardilovich K, Jefferson W, Cheng RY, Medvedovic
M, et al. Persistent hypomethylation in the promoter of nucleosomal binding protein 1 (Nsbp1) correlates with overexpression of Nsbp1 in mouse
uteri neonatally exposed to diethylstilbestrol or genistein. Endocrinology
2008;149:592231.
[242] Guerrero-Bosagna CM, Sabat P, Valdovinos FS, Valladares LE, Clark SJ. Epigenetic and phenotypic changes result from a continuous pre and post natal
dietary exposure to phytoestrogens in an experimental population of mice.
BMC Physiol 2008;8:17.
[243] Qin W, Zhu W, Shi H, Hewett JE, Ruhlen RL, MacDonald RS, et al.
Soy isoavones have an antiestrogenic effect and alter mammary promoter hypermethylation in healthy premenopausal women. Nutr Cancer
2009;61:23844.
[244] Hong T, Nakagawa T, Pan W, Kim MY, Kraus WL, Ikehara T, et al. Isoavones
stimulate estrogen receptor-mediated core histone acetylation. Biochem Biophys Res Commun 2004;317:25964.
[245] Basak S, Pookot D, Noonan EJ, Dahiya R. Genistein down-regulates androgen
receptor by modulating HDAC6-Hsp90 chaperone function. Mol Cancer Ther
2008;7:3195202.
[246] Jawaid K, Crane SR, Nowers JL, Lacey M, Whitehead SA. Long-term genistein
treatment of MCF-7 cells decreases acetylated histone 3 expression and alters
growth responses to mitogens and histone deacetylase inhibitors. J Steroid
Biochem Mol Biol 2010;120:16471.
[247] Kikuno N, Shiina H, Urakami S, Kawamoto K, Hirata H, Tanaka Y, et al.
Genistein mediated histone acetylation and demethylation activates tumor
suppressor genes in prostate cancer cells. Int J Cancer 2008;123:55260.
[248] Zhu W, Qin W, Zhang K, Rottinghaus GE, Chen YC, Kliethermes B, et al.
Trans-resveratrol alters mammary promoter hypermethylation in women at
increased risk for breast cancer. Nutr Cancer 2012.
[249] Papoutsis AJ, Borg JL, Selmin OI, Romagnolo DF. BRCA-1 promoter hypermethylation and silencing induced by the aromatic hydrocarbon receptorligand
TCDD are prevented by resveratrol in MCF-7 cells. J Nutr Biochem
2011.
[250] Papoutsis AJ, Lamore SD, Wondrak GT, Selmin OI, Romagnolo DF. Resveratrol prevents epigenetic silencing of BRCA-1 by the aromatic hydrocarbon
receptor in human breast cancer cells. J Nutr 2010;140:160714.
[251] Howitz KT, Bitterman KJ, Cohen HY, Lamming DW, Lavu S, Wood JG, et al.
Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan. Nature 2003;425:1916.
[252] Chen Y, Shu W, Chen W, Wu Q, Liu H, Cui G. Curcumin, both histone
deacetylase and p300/CBP-specic inhibitor, represses the activity of nuclear
factor kappa B and Notch 1 in Raji cells. Basic Clin Pharmacol Toxicol
2007;101:42733.
[253] Fuand S, Kurzrock R. Development of curcumin as an epigenetic agent. Cancer
2010;116:46706.
[254] Balasubramanyam K, Varier RA, Altaf M, Swaminathan V, Siddappa NB, Ranga
U, et al. Curcumin, a novel p300/CREB-binding protein-specic inhibitor
of acetyltransferase, represses the acetylation of histone/nonhistone proteins and histone acetyltransferase-dependent chromatin transcription. J Biol
Chem 2004;279:5116371.

Вам также может понравиться