Open access peer-reviewed chapter

DNA Damage and Repair Mechanisms Triggered by Exposure to Bioflavonoids and Natural Compounds

Written By

Donna Goodenow, Kiran Lalwani and Christine Richardson

Submitted: 28 May 2020 Reviewed: 10 December 2020 Published: 13 January 2021

DOI: 10.5772/intechopen.95453

From the Edited Volume

DNA - Damages and Repair Mechanisms

Edited by Payam Behzadi

Chapter metrics overview

669 Chapter Downloads

View Full Metrics

Abstract

Eukaryotic cells use homologous recombination (HR), classical end-joining (C-NHEJ), and alternative end-joining (Alt-EJ) to repair DNA double-strand breaks (DSBs). Repair pathway choice is controlled by the activation and activity of pathways specific proteins in eukaryotes. Activity may be regulated by cell cycle stage, tissue type, and differentiation status. Bioflavonoids and other environmental agents such as pesticides have been shown to biochemically act as inhibitors of topoisomerase II (Top2). In cells, bioflavonoids directly lead to DNA double-strand breaks through both Top2-dependent and independent mechanisms, as well as induce DNA damage response (DDR) signaling, and promote alternative end-joining and chromosome alterations. This chapter will present differences in expression and activity of proteins in major DNA repair pathways, findings of Top2 inhibition by bioflavonoids and cellular response, discuss how these compounds trigger alternative end-joining, and conclude with implications for genome instability and human disease.

Keywords

  • environmental compounds
  • bioflavonoids
  • DNA double-strand breaks
  • topoisomerase II
  • DNA break repair
  • genome instability

1. Introduction

The faithful repair of deoxyribonucleic acid (DNA) lesions is central to the maintenance of genomic integrity [1]. DNA double-strand breaks (DSBs) occur during normal developmental processes including meiosis, mating-type switching, V(D)J recombination, antigen receptor gene rearrangement, and also through normal activity of topoisomerase II (Top2) [2, 3, 4, 5]. DSBs also result from exposure to exogenous sources such as ionizing radiation (IR), reactive oxygen species, and chemotherapeutic agents including inhibitors of Top2 [6, 7, 8, 9]. Aberrant repair of DSBs may be mutagenic and result in cell lethality or promote oncogenic transformation. Repair of DSBs in eukaryotes occurs by either homology-dependent or homology-independent (also known as end-joining or illegitimate) mechanisms [10, 11, 12, 13]. In yeast, homology-dependent repair predominates over end-joining [10, 14]. In mammalian cells, direct examination of repair products has demonstrated the predominant use of end-joining [13]. The majority of studies generate targeted DSBs by endonucleases or lasers, and introduce artificial repair substrates into the system [15]. However, exposure to natural compounds can lead to multiple DSBs in a variety of chromatin regions and contexts [16, 17, 18, 19, 20]. Understanding how cells respond to these compounds and repair damage caused by them has important implications for genome stability.

Bioflavonoids are natural compounds in soy, fruits, vegetables, tea, coffee, and wine, and contained in energy drinks and dietary supplements [21, 22, 23, 24]. Bioflavonoids are also in pesticides and flame retardants [25, 26, 27]. Bioflavonoids inhibit the enzyme topoisomerase II (Top2) to promote DSBs, and recent studies have elucidated the cellular mechanisms used to repair the DSBs induced by bioflavonoids [16, 28, 29]. This chapter will discuss cell type differences in expression and activity of proteins in major DNA repair pathways, summarize findings of cellular response to bioflavonoids and Top2 inhibition, discuss how these compounds trigger alternative end-joining, and conclude with implications for genome instability and human disease.

1.1 DNA double-strand break repair

There are three main repair pathways to deal with DNA double-strand breaks (DSBs) in eukaryotic cells. These include classic nonhomologous end-joining (C-NHEJ) (Figure 1A) that modifies and allows for ligation of ends, alternative end-joining (Alt-EJ) that generates short overhangs or exposes small regions of homology via resection to promote ligation of ends (Figure 1B), and homologous recombination (HR) that uses a homologous sister chromatid, chromosome, or other sequence as a template to direct repair synthesis (Figure 1C) [10, 30]. HR is the most accurate using a homologous template as a donor sequence. DSBs are recombination initiators in both meiotic and mitotic cells [31, 32, 33]. However, HR has the most protein involvement, is tightly regulated, largely limited to S phase, and kinetically slow. C-NHEJ is utilized throughout the cell cycle and is kinetically fast. Alt-EJ is less well characterized than the other two and considered a backup repair mechanism when HR or C-NHEJ cannot be used. For a DSB to be repaired by HR or either of the end-joining pathways, damage must first be sensed, then signal transduction pathways must be activated for the DNA damage response (DDR) to bring proteins necessary for repair to the site(s) of damage. Indirect signaling and direct repair protein levels along with histone modifications appear to direct DSB repair pathway selection [34, 35, 36, 37]. Despite decades of extensive study of DSB repair, scientists continue to identify and characterize new factors mechanistically involved in DSB end processing, repair itself, as well as pathway choice [38].

Figure 1.

The DNA double-strand breaks (DSB) are repaired by the three pathways; these are – A) non-homologous end joining (C-NHEJ) which modifies the ends and allow ligation of the broken ends to repair the DSB; B) alternate end-joining (alt-EJ of EJ) creates short DNA overhangs with small regions of homology and ligates the resected broken ends; and C) homologous recombination (HR) that uses a homologous sequence from sister chromatid or homologous chromosome or a homologous sequence within the genome.

1.2 End-joining pathway choice

Repair of DNA DSBs by C-NHEJ or Alt-EJ is characterized by ligation of two DSB ends in close proximity to each other (Figure 1A and B). Initial binding of the Ku70–80 heterodimer competes with poly(ADP-ribose) polymerase 1 (PARP1) for binding to the DSB. If Ku70–80 binds first there is minimal end processing and C-NHEJ is used [39, 40, 41, 42, 43, 44]. For C-NHEJ, DNA-dependent protein kinase catalytic subunit (DNA-PKcs) is recruited to the Ku complex. DNA-PKcs can determine if the ends are blunt, as from a nuclease cleavage or from RAG during V(D)J recombination, or if there are overhangs or protein/group adducts. If the break is clean, DNA-PKcs recruits XRCC4-XLF and LigaseIV, and these proteins work together to ligate the DNA ends [39, 42, 45]. However, if there is an overhang or proteins are attached to the break site, DNA-PKcs recruits the ARTEMIS complex for processing. ARTEMIS can release protein groups and with its nuclease activity to digest the DSB ends until they are blunt to facilitate ligation of the ends [46].

If PARP1 binds to the DSB before Ku70–80, it immediately adds branched poly(ADP-ribose) (PAR) groups to itself and histones in close proximity. The branched PAR recruit the Mre11-Rad50-Nbs1 (MRN) complex to process the ends and proceed by HR or Alt-EJ. Alt-EJ seems to act as a salvage repair mechanism for when HR and C-NHEJ are blocked. It is likely that Alt-EJ occurs when processing for HR has started following PARP1 binding to the break first, Ku70–80 is depleted, because the DSB ends have proteins bound to block template invasion, or the cell is in G1 phase of the cell cycle no homologous template is readily available for repair. Alt-EJ involves MRN and CtIP to resect the DSB ends in a 3′ to 5′ fashion, termed short range end resection, of 5–25 nucleotides to create short DNA overhangs with small regions of homology. Polymerase θ is utilized in Alt-EJ. After processing, XRCC1 and LigaseIII act in a complex to ligate the ends and remove the overhanging bases. Alt-EJ is more mutagenic than HR or C-NHEJ and associated with chromosomal rearrangements and translocations [44, 47, 48, 49].

1.3 Homologous recombination requires chromatin remodeling and DDR

To initiate HR (Figure 1C), PARP1 is recruited to the DSB first and immediately adds branched PAR groups to itself and histones in close proximity. The branched PAR recruit the MRN complex and inactive ATM kinase dimers with the acetyltransferase TIP60 attached. PARG quickly removes the PAR groups allowing the MRN complex to bind to the DSB. MRN allows ATM to bind at the DSB and activate through auto-phosphorylation and acetylation by TIP60, thereby allowing TIP60 to dissociate. Once active, ATM will phosphorylate a large number of target proteins including the MRN complex and CtIP that process DSB ends [12, 34, 42, 50].

Chromatin remodeling is extensive and required for HR-mediated DSB repair. Histone H2AX is phosphorylated by ATM as well as acetylated by TIP60. Phospho-H2AX (γ-H2AX) has some chromatin remodeling functions and acts as a signal to recruit additional proteins involved. γ-H2AX will spread away from the DSB to decorate chromatin up to 2 Mb away. MDC1, which assists with chromatin remodeling, becomes phosphorylated by ATM and recruits RNF6 dimers that have ubiquitination functions. HERC2 associates with phosphorylated RNF6 and appears to recruit PIAS4 which has SUMOlyation capabilities. RNF6 becomes SUMOlyated and mono-ubiquitinates histones in the area, which recruits RNF168, another ubiquitin ligase, that is SUMOlyated and poly-ubiquitinates nearby histones. The poly-ubiquitin trees tether BRCA1-A complexes by RAP80 mediators. These complexes cause histone modifications that bring in 53BP1, which has more histone remodeling functions and can inhibit MRN and CtIP-mediated end resection [34, 39, 43, 50].

Phosphorylation of target proteins by ATM also triggers DDR. Chk2 has protein kinase activity allowing it to phosphorylate a number of effector proteins in the cell cycle checkpoint including p53 which can be modified by either Chk2 or ATM (or ATR or Chk1). ARF protein (p14) seems to stabilize TIP60 interactions with ATM for better activation and is associated with maintaining genome stability [34].

While the histone remodeling is occurring and other proteins are being recruited, MRN and CtIP resect the DSB ends short range end resection, then Exo1 or Dna2 nucleases act in long range end bidirectional resection in a 5′ to 3′ direction away from the DSB. Exo1 has dsDNA nuclease function, while Dna2 must act with a helicase like BLM or WRN to unwind DNA for its ssDNA nuclease abilities [34, 37, 43, 51]. While long range end resection is occurring, RPA binds to the 3’ ssDNA overhang to protect from nucleases. After this resection, one type of HR can occur called single strand annealing (SSA), where the two pieces of RPA coated DNA associate with one another with the help of Rad52 and if regions of homology are found they anneal to one another. Non-homologous flaps are cleaved off by enzymes like XPF-ERCC1 and ligated by LigaseIII. This type of HR can cause large deletions [12, 43, 50, 52].

Canonical HR, as well as break-induced replication (BIR) and synthesis-dependent strand annealing (SDSA) use BRCA 1 and 2 with Rad51 for homology searches that cause strand-invasion, D-loop formation and resolution/dissolution. RPA must be dissociated from the ssDNA for Rad51 binding, mediated by DSS1 and BRCA2 which displace RPA and stabilize ATP on Rad51 increasing its binding affinity for the ssDNA. Once Rad51 is loaded on the DNA and the nucleofilament has formed, it can invade neighboring DNA to search for homology with BRCA1 [34, 43, 50]. Homology less than 7 nt in length is a weak interaction and Rad51 not sufficient to initiate HR, but 7 nt or longer allows the strand to interact more strongly [50]. If significant homology is present, the ATP on Rad51 is hydrolyzed causing the dsDNA to dissociate and the nucleofilament anneals with the template strand. RPA stabilizes this D-loop formation by binding to the displaced strand. DNA Polymerase δ or ε uses the invading strand as a primer to initiate synthesis [12, 39, 50, 53]. Resolution can happen with crossover or non-crossover products and different sets of resolvases mediate this process. For one-sided ends that utilize BIR, DNA Pol δ is used and synthesis continues until the end of the chromosome causing gene conversion that can be highly mutagenic [54].

Advertisement

2. Topoisomerase II, inhibitors and poisons

Topisomerase II (Top2) is a regulatory enzyme that relaxes supercoiled DNA for transcription (Top2β) and replication (Top2α). As shown in Figure 2, Top2 acts in a multistep cleavage and religation reaction: (1) Top2 binds to two dsDNA molecules at Top2 recognition sequences; (2) a transient DSB is generated in the first DNA helix (G-segment) creating a cleavage complex; (3) ATP hydrolysis drives a conformational change allowing the second dsDNA helix to pass through the DSB; (4) Top2 mediates religation of the DSB and the T DNA segment is released; (5) the G DNA segment is released and the enzyme returns to its original conformation (Figure 2). A catalytic Top2 inhibitor such as dexrazoxane acts to prevent DNA from binding to Top2 at step 1 preventing any part of the catalytic cycle [55, 56, 57, 58, 59].

Figure 2.

Top2 acts in a multistep cleavage and religation reaction. 1) Top2 binds the G and T dsDNA molecules at Top2 recognition sequences. 2) ATP binding catalyzes the DNA DSB in the G segment, which allows the T segment to pass through the break. (3) ATP hydrolysis drives a conformational change allowing the second dsDNA helix to pass through the DSB; (4) Top2 mediates religation of the DSB and the T DNA segment is released; (5) the G DNA segment is released and the enzyme returns to its original conformation.

Advertisement

3. Bioflavonoids and other natural compounds as Top2 inhibitors

A class of chemical compounds called bioflavonoids are contained in soy, fruits, vegetables, tea, coffee, wine, energy drinks, and dietary supplements [21, 22, 23, 24, 25, 26, 27]. Bioflavonoids are characterized by multiple phenolic rings that are central to their ability to inhibit the enzyme Top2 in a similar manner to the chemotherapeutic drug etoposide [16, 28, 29]. Some pesticides and flame retardants also contain multiple phenolic rings and have been identified as Top2 inhibitors. Bioflavonoids are separated into 12 different sub-classes based upon their structure; however only six are contained in dietary sources: flavanols, flavonols, flavones, isoflavones, flavanones, and anthocyanidins (Figure 3) [60, 61].

Figure 3.

Basic chemical structures of dietary bioflavonoids. The middle circled backbone represents the general bioflavonoid poly-phenol ring structure. The six structures surrounding show the general structural differences between the sub-groups.

3.1 Isoflavones

Isoflavones are polyphenolic secondary plant metabolites produced through the flavonoid-producing phenyl-propanoid synthesis pathway (Figure 4). In order for isoflavone production, the plant must express the isoflavone synthase enzyme which converts flavanone precursors into isoflavones. This isoflavone synthase is only expressed in legumes and a few other select species. Plants with the highest concentrations of isoflavones are soy, red clover, and kudzu. The amount of isoflavone depends upon the conditions the plants were grown, and the final concentration of isoflavones in food products (including dietary supplements) depends upon which portion of the plant is used and the processing methods. Genistein, daidzein, glycitein, formononetin, biochanin A and irilone are the main isoflavones isolated from plants [60, 62, 63]. Genistein and daidzein are of particular interest due to their high concentration in soy products [60]. Genistein is an estrogen derivative available at health food stores as dietary and menopausal supplements, and a soy phytoestrogen present in foods, particularly soybeans, and infant soy formulas [23, 64, 65].

Figure 4.

Structure of commonly found bioflavonoids flavanols: Genistein and Daidzein, flavonols: Kaemferol quercetin and Myricetin, and flavones: Luteolin.

Interest in isoflavones has spiked in the past 20 years. This is due to the attribution of consumption of isoflavone-containing products with lower occurrences of coronary heart disease, breast and prostate cancer. This hypothesis derived from observations that citizens of Asian countries have lower incidence of these diseases compared to citizens of Western countries, and that citizens in Asian countries typically ingest 8-50 mg/day of isoflavones compared to citizens in Western countries who ingest only 0.1–3.3 mg/day [66, 67] .

Due to this potential health relevance, studies examined the impact of high intake of isoflavones, but the results have been inconclusive [62]. In animal models, increased genistein intake resulted in increased rates of pituitary and mammary gland tumors and stimulated MCF-7 tumor growth. Additionally, while increased genistein intake in post-menopausal women in Asian countries decreased breast cancer risk, this decreased risk was not sustained in post-menopausal women in Western countries, including both native inhabitants and Asian immigrants. Some studies, particularly of British women, showed that increased serum genistein levels in women with early stage breast cancer had increased transcription of cell cycle progression and cell proliferation genes [62].

3.2 Flavones

Flavones are the end product of a complex multi-step synthetic pathway that occurs within a wide variety of plants (Figure 4). This pathway begins with phenylalanine that is converted through the generalized phenylpropanoid pathway that synthesizes most flavonoids. Subsequently, p-coumaroyl-CoA must be synthesized into chalcone with chalcone synthase. Chalcone can be isomerized into a flavanone by chalcone isomerase. Finally, flavone synthase class I or II enzymes catalyze the synthesis of a flavone from flavanones. Flavones, similar to flavonols, can protect the plant from UV-B radiation. Flavones have the additional ability to provide protection against biological attacks from pathogenic microbes by acting as signaling molecules to activate differential gene transcription to prevent the growth of microorganisms after invasion. Additionally, flavones can be expressed to deter insects and nematodes from eating the plant or to interfere with the growth and reproduction of other plants [60].

Flavones are found across a variety of plant species, and expression of flavones appears to be widespread within the plant, from the roots to the leaves. However, though flavones are found throughout the plant kingdom, they are found much less commonly in fruits and vegetables as compared to flavonols. Apigenin and luteolin are the main flavonols contained in food sources including celery, parsley, thyme, red peppers, and fruit skins [61, 68]. In humans, flavones, much like isoflavones and flavonols, seem to have antioxidant and anti-tumor capabilities and to affect signal transduction pathways [69].

3.3 Flavonols

Flavonols are primarily in fruits, vegetables, red wine, and tea and they compose the largest portion of humans’ bioflavonoid intake given their distribution across a wide number of plant species (Figure 4) [61]. Within plants it has been shown that flavonols have the ability to protect the plant against UV-B damage, and they protect the plants against oxidative damage with their antioxidant capability [70, 71]. Scientists and physicians want to determine ways to utilize the antioxidant capability of flavonols in human populations as a protectant against cardiovascular and neurological disease and against exercise induced oxidation in smokers and athletes [72, 73].

The most common flavonols in foods are quercetin, kaempferol, myricetin, and fiestin, with a majority of published literature focusing upon the first three. Similar to isoflavones the concentration of flavonol in the food product depends upon the plant, the growth conditions, and the part of the plant used. Flavonols are found in highest concentrations in the leaves, flowers, and fruits, which are exposed to sunlight; the exception to this being onions which grow below ground [70, 71]. The human dietary source of flavonols is dependent on culture and region. Humans residing in Asian countries typically ingest flavonols through green tea, while the Netherlands, United States and Denmark inhabitants mainly ingest them from onions, apples, and tea. Citizens of Mediterranean areas ingest flavonols from green vegetables. Within Italy, red wine is the main source of flavonols, though inhabitants of Northern villages also have a high intake from salads, soups, fruits. The prevalence of flavonols in the human diet has produced a large interest in understanding their multiple cellular effects and potential impact on human health [70].

3.4 Additional compounds as Top2 inhibitors

Additional natural compounds other than bioflavonoids may also act as inhibitors of Top2. Bakuchicin from the furanocoumarin family is present in fruits and legumes [74]. In research conducted to study DNA-polymerase inhibition activity of Psoralea corylifolia L. (Leguminosae), bakuchincin was found to be a weak Top2 inhibitor [75]. Additional reported naturally occurring Top1 and Top2 inhibitors are benzophenone compounds such as xanthochymol and Garcinol at effective concentrations comparable to those of etoposide (∼25 − 100 μM) [76, 77]. A comparative study between the naturally occurring constituent of black seed thymoquinone used as a spice in eastern cooking and a known Top2 inhibitor 1,4-benzoquinone showed structural and functional similarity between the two compounds and the ability to induce DNA cleavage [78].

Triterpenoids are present in plants, widely distributed within the root, stem, leaves, bark. They are components in the waxy covering of fruits and herbs such as jujube, lavender, and thyme [79]. Triterpenoids have two major components, C5 units and isopentyl diphosphate [80], and are generally present as saponins that act as defense chemicals for protection against microbes. Triterpenoids betulin lupane and oleanane from the bark of Phyllanthus flexuosus, derivatives of betulinc acid, and oxygenated derivatives of oleanane called celastroloids were reported to act as human Top2 inhibitors to varying degrees [81, 82, 83, 84]. In addition, betulinic acid which is an oxidative derivative of betullin inhibits cell proliferation by inhibiting topoisomerase-DNA binding and suppressing NF-κB activation [83].

Halogenated compounds in household and baby products include polychlorinated biphenyls (PCBs), detectable in indoor carpets, and polybrominated diphenyl ethers (PBDEs), used as flame retardants, increase DNA cleavage by TopIIα in vitro and in cultured human cells [85]. Recent CRISPR-Cas9 screening against a large panel of genotoxic agents identified the synthetic small molecule pyridostatin as a Top2 inhibitor. Pyridostatin is a G-quadruplex stabilizer and this stabilization mechanism may lead to Top2 trapping on DNA [38].

Advertisement

4. Flavonols, flavonols, flavones, isoflavones, flavanones, and anthocyanidins act as Top2 poisons and trigger illegitimate DNA repair mechanisms

A catalytic Top2 inhibitor such as dexrazoxane acts to prevent DNA from binding to Top2 thus preventing any part of the catalytic cycle to occur [55, 56, 57, 58, 59]. By contrast, some chemicals including bioflavonoids act as Top2 “poisons” (Figure 5) [28, 86]. A Top2 poison acts on Top2 after DNA binding and prevents the normal function of Top2 (step 2 of catalytic cycle, see Figure 2). Top2 poisons can be further classified as covalent or traditional poisons. The potential as a covalent or traditional poison is dependent on biochemical structure. These groups are not mutually exclusive and individual bioflavonoids can act through one or both mechanisms [29, 86].

Figure 5.

Bioflavonoid classification as a covalent or traditional topoisomerase II poison. The blue boxed regions indicate required biochemical features for a traditional Top2 poison. The red 3′, 4′ and 5’-OH groups on the B ring are necessary for covalent Top2 poisons. The 4’-OH group is required for covalent binding, while the 3′, 5′ increase the binding affinity, therefore a bioflavonoid with all 3-OH groups would be a strong covalent poison.

4.1 Bioflavonoids as covalent Top2 poisons

Flavanols, flavonols, flavones, flavanones, and anthocyanidins (but not isoflavones) have the potential to act as strong covalent Top2 poisons [86]. A covalent Top2 poison works in a redox-dependent manner, binding to a distal site on the Top2 enzyme and increasing its ability to cause a DSB in step 2 of the catalytic multi-step reaction through conformational changes to the enzyme. The key structural component for a covalent poison is having 3 –OH groups on the B ring of the bioflavonoid structure. However, it is likely bioflavonoids with 2 –OH groups on the B ring act as a weak covalent poison and the ability to act as a covalent poison increases with more –OH groups (Figure 5) [49, 50]. A 4’-OH group on the B ring is necessary for binding, and 3′ and 5’-OH groups improve covalent binding strength. Thus, a strong covalent poison contains 3 –OH groups on the B ring of the bioflavonoid structure. For example, among the flavonols, structure predicts that myricetin has high activity, quercetin has intermediate activity, and kaempferol has weak activity, if at all, as a covalent Top2 poison (Figure 4). Cell free studies support this and show that myricetin as well as epigallocatechin-gallate (ECGC) act as strong coalvent poisons, quercetin acts as a weak traditional poison, but kaempferol does not have this activity [29, 87].

4.2 Bioflavonoids as traditional Top2 poisons

Flavones, flavonols, isoflavones, and flavanones (but not flavanols or anthocyanidins) have the all act as traditional (or interfacial) Top2 poisons. The key structural components for a traditional Top2 poison are a 5’-OH group in the A ring, a 4’-OH group in the B ring, and a 4′ = O in the C ring (Figure 5). A traditional (or interfacial) Top2 poison stalls the enzyme by binding to the active site of the enzyme preventing religation, thereby resulting in the formation of a stabilized cleavage complex (SCC) [88, 89]. Flavonols are strong traditional poisons and both cell free and cell culture systems support this. Similarly, experiments in cell culture systems examining the kinetics of DSB repair following exposure to acute doses of bioflavonoids support the model that flavonols, flavones, and isoflavones including kaempferol, quercetin, myricetin, genistein, and luteolin and each act as a traditional Top2 poison. However, combinatorial activity of genistein, quercetin and luteolin together suggests they may have weak covalent poisoning capabilities when they have to compete for the traditional poisoning binding site [29].

4.3 Bioflavonoids trigger illegitimate DNA repair mechanisms

Bioflavonoids with either covalent and traditional Top2 poisoning activity induce the DSB-mediated DDR as evidenced by induction of γ-H2AX foci, ATM phosphorylation, and p53 signaling [90, 91, 92]. However, a more direct role or influence of these compounds on the repair of damage is not as clear [93]. Acute doses induce DNA damage and DDR as detected by γ-H2AX foci and phosphorylation of ATM in stem cells and CD34+ hematopoietic progenitor cells [94, 95]. Genistein and quercetin inhibit Top2 to induce DNA DSBs, and also appear to influence DSB repair pathway choice. Protein level analysis for HR, C-NHEJ, and Alt-EJ specific proteins suggests that genistein and quercetin suppress HR by reducing BRCA2 and Rad51 expression, as well as suppress C-NHEJ by suppressing levels of DNA-PKcs, Ku80, XLF and XRCC4 and trigger Alt-EJ by increasing levels of CtIP and Polymerase θ [96, 97]. DNA reporter assays suggest that quercetin interferes with DNA repair mechanisms such as HR and C-NHEJ by inhibition of PI3K/Akt signaling. In support of these studies, exposure to multiple bioflavonoids promotes the generation of chromosomal translocations in a dose-dependent manner [29, 87].

Bioflavonoids that have traditional Top2 poisoning activity lead to trapped SCCs on the DNA. Removal of SCCs is performed by the small ubiquitin-related modifier ligase ZNF45/tyrosyl-DNA phosphodiesterase 2 (ZATT/TDP2) complex. Removal of the SCC is required for DSB repair by C-NHEJ. If ZATT/TDP2 does not remove the SCC, the MRN complex or CtIP with nuclease activity may resect the DNA ends with the SCC attached to allow for DSB repair by HR or Alt-EJ [55, 94, 98, 99, 100, 101]. Inhibition or mutation of multiple DNA repair proteins potentiates cytotoxicity of Top2 inhibitors, and MRE11 plays a direct mechanistic role in removal of Top2-DNA complexes in yeast and mammals [102, 103].

Advertisement

5. Pleiotropic effects of bioflavonoids

Due to their antioxidant capacity, bioflavonoids are included in dietary supplements for their presumed health benefits in protecting against inflammation, cardiovascular diseases, and cancer [87]. These beneficial health properties are due to the number of pleiotropic effects bioflavonoids have on cells by impacting signal transduction pathways, DSB repair and the cellular epigenetic landscape, which can lead to protein level changes, cell cycle stalling, and apoptosis [16, 69].

5.1 Bioflavonoids and signal transduction pathways

Bioflavonoids have antioxidant and anti-inflammatory properties. Their antioxidant properties are due to their ability to reduce reactive oxygen species of the multiple –OH groups in their chemical structure. Their anti-inflammatory properties are due to their interference with signal transduction pathways and down-regulation in the production of pro-inflammatory cytokines. Bioflavonoids decrease inflammation and immune cell recruitment through interference with the ERK/MAP kinase and NF-κB signal transduction pathways which can be beneficial to human health. NF-κB is a transcription factor that upon activation is transported into the nucleus and binds to the promoter region for a number of cytokines and apoptotic genes; therefore reduced pathway activation leads to lower pro-inflammatory cytokine production and increased cell survival [104]. Extracts from the plant Ginkgo biloba, rich in bioflavonoids, act as an herbal antioxidant, augment the transcription of TNF-α causing reduced activation of the NF-κB pathway. Apigenin has shown similar down regulatory effects on cytokine production likely through the modulation of NF-κB activation [105]. Quercetin and fisetin inhibit pro-inflammatory cytokine production through the suppression of NF-κB activation by decreased phosphorylation of extracellular signal-regulated (ERK) kinase and p38 mitogen-activated protein (MAP) kinase that are activators of NF-κB [106, 107, 108]. Myricetin has been shown to affect the phosphatidylinositol 3-kinase (PI3-K) pathway inducing apoptosis in pancreatic cells [109].

5.2 Bioflavonoids and epigenetic modifications

Studies in cancer cell lines demonstrate epigenetic modifications caused by bioflavonoids. Genistein, quercetin, curcumin, EGCG, hesperidin, and naringin are inhibitors of DNA methyltransferases leading to hypomethylation of DNA. In addition, many of these bioflavonoids have also been shown to act on histone acetyltransferases and histone deacetyltransferases causing cell wide alterations in histone epigenetic modification patterns [109].

Long-term epigenetic effects of bioflavonoids compounds were addressed in several mouse model studies. Exposure to genistein through maternal diet during pregnancy can have long-lasting effects on the progeny. In agouti mouse pups exposed to genistein from conception until birth, epigenetic changes were observed as altered coat color, as well as significant downregulation of genes involved in hematopoiesis of bone marrow cells, increased erythropoiesis, and a permanent signature hypermethylation of repetitive elements in hematopoietic lineages [110]. Likewise, in mice exposed to quercetin from conception until birth resulted in upregulated iron-associated cytokine expression, significantly increased iron storage in the liver, and hypermethylation of repetitive elements. Epigenetic modifications lead to long term gene expression changes of cytokines associated with inflammation in the liver of the mice in adulthood [111, 112].

Advertisement

6. Implications for human health

6.1 Potential anti-cancer applications

While bioflavonoids can be beneficial through intake at low or moderate doses, high doses and acute exposure of bioflavonoids may more drastically inhibit Top2 and impact genome integrity and cell survival, thus changing their overall impact on cells and human health. In vitro studies support the idea that bioflavonoids genistein and quercetin may act as chemo-preventive or anti-cancer agents by altering major processes within cancer cells such as apoptosis, cell cycle, angiogenesis and metastasis [113, 114]. Genistein has synergistic behavior with well-known anticancer drugs adriamycin, docetaxel, and tamoxifen, suggesting a potential role in combination cancer therapy [78]. Quercetin in combination with doxorubicin was found to be more effective in inducing apoptosis within the SKOV-3 cells [114]. A combinatory treatment with quercetin and curcumin synergistically induce anti-cancer activity in triple-negative breast cancer cells by modulating tumor suppressor genes in particular enhancing BRCA1 expression [115].

Several bioflavonoids have been investigated as alternate cancer therapeutics that are less genotoxic than traditional chemotherapeutics but equally effective. High concentrations of myricetin causes Top2-mediated DNA damage and apoptosis in K652 cells [116]. Fisetin interrupts the MAPK-dependent NF-κB signaling pathway in cervical cancer cells, inhibiting migration and invasion [114]. Several in vitro and in vivo studies indicate that luteolin can suppress metastasis of breast cancer by reversing epithelial-mesenchymal transition, or by acting as an antiangiogenic therapeutic inhibiting VEGF production and suppressing invasion [117, 118].

While these observations strengthen the notion that flavonoids could be useful anti-cancer agents, to date minimal clinical studies have demonstrated that these bioflavonoids retain anti-cancer properties in humans in vivo. A Phase I study/pharmacokinetic trial of quercetin in cancer patients intravenously injected quercetin in 11 patients with cancer at varying doses of 60–2000 mg/m2 and identified 945 mg/m2 as a safe and effective dose [119].

6.2 Potential inducers of infant leukemia

Aberrant repair of DNA DSBs caused by either endogenous or exogenous agents has the potential to result in DNA sequence mutations or genome rearrangements such as chromosomal translocations which can lead to disease. Negative consequences of high bioflavonoid intake can be observed most prominently in pregnant women. Epidemiological data from countries whose citizens have higher bioflavonoid intake (particularly soy products) had a 2–3 times higher incidence of infant leukemia, characterized by chromosomal translocation, suggesting maternal intake of high amounts of bioflavonoids could lead to this particular genome rearrangement and infant leukemia [120].

Infant leukemia typically occurs due to translocation events involving the mixed lineage leukemia (MLL) gene. Most of the MLL rearrangements observed in patients with infant leukemia and therapy-related leukemia (tAML) cluster together in a well-defined region of the MLL locus [121]. tAML is associated with treatment with Top2 poisons etoposide or doxorubicin [86, 88, 121] which has led to the hypothesis and working model that ingestion of natural Top2 poisons including bioflavonoids can lead to these translocation events and tumorigenesis [121, 122]. In support of this, boflavonoids have been shown to inhibit Top2 and induce MLL cleavage and translocations in hematopoietic stem cell-enriched populations [87, 121].

Foods contain multiple different bioflavonoids, and bioflavonoids are bio-accumulative which likely increases plasma concentrations [123]. Study of the potential for environmental or dietary compounds to induce infant leukemias is more relevant since they cross the placental barrier as shown with the synthetic bioflavonoid EMD-49209 [124], genistein [111, 125], quercetin [111], herbal medicines, dipyrone, and pesticides including the mosquitocidal Baygon [126, 127]. Genotoxic effects of quercetin on the human hemopoietic stem and progenitor cells (HSPCs) were shown using a genetically engineered placental barrier model from a specialized human cell line. This study showed that approximately 10% of quercetin from the maternal side is capable of crossing the placental barrier and accumulating in the fetus. Exposure in utero is likely more damaging due to differences in metabolic and excretion rates of mother and fetus [128] as well as rapidly developing and proliferating fetal cells that are more sensitive to Top2 inhibiting agents [129].

Advertisement

7. Conclusion

Bioflavonoids are prevalent in the human diet from natural sources such as fruits and vegetables, but are also found at supranatural concentrations in dietary supplements and energy drinks. These chemical compounds have numerous cellular effects including interfering with signal transduction pathways, modifying the DNA damage response and epigenetic markers, and poisoning of Top2 causing DNA DSBs and leading to aberrant repair. Given the number of cellular pathways bioflavonoids affect, and the DNA damage caused by bioflavonoid exposure, it is possible that bioflavonoids could be used as natural analogs of traditional chemotherapeutic agents. However, more research is needed to understand how these bioflavonoids cause DNA damage through Top2-dependent or -independent pathways to understand potential off-target negative effects. In addition, further research will be needed to understand the dose-dependent activities of bioflavonoids and at what doses they may be chemo-protective versus what threshold doses they may induce DNA damage that is mutagenic, and finally at what high acute doses they may induce DNA damage and apoptosis to act as effective alternative to traditional chemotherapeutic agents.

Advertisement

Acknowledgments

CR is funded in part by NIH/NIGMS and a Faculty Research Grant (UNC Charlotte). KL was funded in part by Proposal Development Summer Fellowship (UNC Charlotte).

Advertisement

Conflict of interest

The authors indicate no conflict of interest.

References

  1. 1. van Gent DC, Hoeijmakers JH, Kanaar R. Chromosomal stability and the DNA double-stranded break connection. Nat Rev Genet. 2001; 2(3): 196-206.
  2. 2. Pandita TK and Richardson C. Chromatin remodeling finds its place in the DNA double-strand break response. Nucleic Acids Res. 2009; 37(5):1363-77.
  3. 3. Lieber MR. The mechanism of V(D)J recombination: a balance of diversity, specificity, and stability. Cell, 1992; 70(6):873-6.
  4. 4. Lieber MR, Yu K, Raghavan SC Roles of nonhomologous DNA end joining, V(D)J recombination, and class switch recombination in chromosomal translocations. DNA Repair (Amst), 2006; 5(9-10):1234-45.
  5. 5. Keeney S, Giroux CN, Kleckner N. Meiosis-specific DNA double-strand breaks are catalyzed by Spo11, a member of a widely conserved protein family. Cell, 1997; 88(3):375-84.
  6. 6. Waris G and Ahsan H. Reactive oxygen species: role in the development of cancer and various chronic conditions. J Carcinog, 2006; 5:14.
  7. 7. Osheroff N, Corbett A, and Robinson M. Mechanism of action of topoisomerase II-targeted antineoplastic drugs. Adv Pharmacol., 1994; 29B: 105-126.
  8. 8. Sung PA, Libura J, and Richardson C. Etoposide and illegitimate DNA double-strand break repair in the generation of MLL translocations: New insights and new questions. DNA Repair (Amst), 2006; 5(9-10):1109-18.
  9. 9. Ward J. DNA damage produced by ionizing radiation in mammalian cells: identities, mechanisms of formation, and reparability. Prog Nucleic Acid Res Mol Biol, 1988; 35:95-125.
  10. 10. Lalwani K, Goodenow D, Richardson C. Eukaryotic Recombination: Initiation by double-strand breaks. Encyclopedia of Life Sciences. 2020; 1:69-76. DOI:10.1002/9780470015902.a0029148
  11. 11. Sancar A, Lindsey-Boltz LA, Ünsal-Kaçmaz K, Linn S. Molecular Mechanisms of Mammalian DNA Repair and the DNA Damage Checkpoints. Annual Review of Biochemistry, 2004; 73(1):39-85.
  12. 12. Jasin M and Rothstein R. Repair of strand breaks by homologous recombination. Cold Spring Harbor Perspectives in Biology, 2013; 5(11),1-18. https://doi.org/10.1101/cshperspect.a012740
  13. 13. Pannunzio NR, Watanabe G, Lieber MR. Nonhomologous DNA end-joining for repair of DNA double-strand breaks. JBC. 2018; 293(27):10512-10523.
  14. 14. Haber JE. Chromosome breakage and repair. Genetics, 2006; 173(3):1181-1185.
  15. 15. Richardson C, Elliott B, and Jasin M. Chromosomal double-strand breaks introduced in mammalian cells by expression of I-Sce I endonuclease. Methods Mol Biol, 1999; 113:453-63.
  16. 16. Bariar B, Vestal CG, Richardson, C. Long-term impact of chromatin remodeling and DNA damage in stem cells induced by environmental toxins and dietary agents. J Environ Pathology, Toxicology, and Oncology, 2013; 32(4):305-25. PMID:24579784
  17. 17. Baccarelli A, Bollati V. Epigenetics and environmental chemicals. Curr Opin Pediatr. 2009; 21(2):243-51.
  18. 18. Lai W, Li H, Liu S, Tao Y. Connecting chromatin modifying factors to DNA damage response. Int J Mol Sci. 2013; 14(2):2355-69.
  19. 19. Xu Y, Price BD. Chromatin dynamics and the repair of DNA double strand breaks. Cell Cycle. 2011; 10(2):261-7.
  20. 20. Pandita TK, Richardson C. Chromatin remodeling finds its place in the DNA double-strand break response. Nucleic Acids Res. 2009; 37(5): 1363-77.
  21. 21. Vanhees K, de Bock L, Godschalk RW, van Schooten FJ, van Waalwijk van Doorn-Khosrovani SB. Prenatal exposure to flavonoids: implication for cancer risk. Toxicol Sci. 2011;120(1): 59-67.
  22. 22. Ross JA, Kasum CM. Dietary flavonoids: bioavailability, metabolic effects, and safety. Annu Rev Nutr 2002; 22:19-34.
  23. 23. Williamson-Hughes PS, Flickinger BD, Messina MJ, Empie MW. Isoflavone supplements containing predominantly genistein reduce hot flash symptoms: a critical review of published studies. Menopause. 2006; 13(5):831-839.
  24. 24. Setchell KD, Zimmer-Nechemias L, Cai J, Heubi JE. Exposure of infants to phyto-oestrogens from soy-based infant formula. Lancet, 1997; 350(9070):23-27.
  25. 25. Alexander FE, Patheal SL, Biondi A, Brandalise S, Cabrera ME, Chan LC, Chen Z, Cimino G, Cordoba JC, Gu LJ, Hussein H, Ishii E, Kamel AM, Labra S, Magalhaes IQ , Mizutani S, Petridou E, de Oliveira MP, Yuen P, Wiemels JL, Greaves MF. Transplacental chemical exposure and risk of infant leukemia with MLL gene fusion. Cancer Res, 2001; 61(6):2542-2546.
  26. 26. Pombo-de-Oliveira MS, Koifman S, Brazilian Collaborative Study Group of Infant Acute L. Infant acute leukemia and maternal exposures during pregnancy. Cancer Epidemiol Biomarkers Prev, 2006; 15(12):2336-2341.
  27. 27. Ward MH, Colt JS, Metayer C, Gunier RB, Lubin J, Crouse V, Nishioka MG, Reynolds P, Buffler PA. Residential exposure to polychlorinated biphenyls and organochlorine pesticides and risk of childhood leukemia. Environ Health Perspect, 2009; 117(6):1007-1013.
  28. 28. Bandele OJ and Osheroff N. Bioflavonoids as poisons of human topoisomerase IIα and IIβ. Biochemistry, 2007; 46(20):6097-6108. https://doi.org/10.1021/bi7000664
  29. 29. Goodenow D, Emmanuel F, Berman C, Sahyouni M, and Richardson, C. Bioflavonoids cause DNA double-strand breaks and chromosomal translocations through topoisomerase II-dependent and -independent mechanisms. Mutation Research - Genetic Toxicology and Environmental Mutagenesis, 2020; 849:503144. https://doi.org/10.1016/j.mrgentox.2020.503144
  30. 30. Sirbu BM, and Cortez D. DNA damage response: three levels of DNA repair regulation. Cold Spring Harbor Perspectives in Biology, 2013; 5(8):a012724–a012724. https://doi.org/10.1101/cshperspect.a012724
  31. 31. Richardson C, Moynahan ME, and Jasin M. Double-strand break repair by interchromosomal recombination: suppression of chromosomal translocations. Genes Dev, 1998; 12(24):3831-3842.
  32. 32. Richardson C and Jasin, M. Frequent chromosomal translocations induced by DNA double-strand breaks. Nature, 2000; 405:697-700.
  33. 33. White R, Sung P, Vestal CG, Benedetto G, Cornelio N, and Richardson C. Double-strand break repair by interchromosomal recombination: an in vivo repair mechanism utilized by multiple somatic tissues in mammals. PlosONE, 2013; 8(12): 1-16. e84379. PMID: 24349572 PMCID: PMC3862804
  34. 34. Ciccia A and Elledge SJ. The DNA Damage Response: Making It Safe to Play with Knives. Molecular Cell, 2010; 40(2): 179-204. https://doi.org/10.1016/j.molcel.2010.09.019
  35. 35. Harper JW and Elledge SJ. The DNA Damage Response: Ten Years After. Molecular Cell, 2007; 28(5):739-745. https://doi.org/10.1016/j.molcel.2007.11.015
  36. 36. Rouse J and Jackson SP. Interfaces between the detection, signaling, and repair of DNA damage. Science, 2002; 297(5581):547-551. https://doi.org/10.1126/science.1074740
  37. 37. Zhou BB and Elledge SJ. The DNA damage response: putting checkpoints in perspective. Nature, 2000; 408(6811):433-439. https://doi.org/10.1038/35044005
  38. 38. Olivieri M, Cho T, Álvarez-Quilón A, Li K, Schellenberg M J, Zimmermann M, Hustedt N, Rossi SE, Adam S, Melo H, Heijink AM, Sastre-Moreno G, Moatti N, Szilard RK, McEwan A, Ling AK, Serrano-Benitez A, Ubhi T, Feng S, Durocher D. A Genetic Map of the Response to DNA Damage in Human Cells. Cell, 2020; 182(2):481-496.e21. https://doi.org/https://doi.org/10.1016/j.cell.2020.05.040
  39. 39. Byrne M, Wray, J, Reinert B, Wu Y, Nickoloff, J, Lee SH, Hromas R, Williamson E. Mechanisms of oncogenic chromosomal translocations. Annals of the New York Academy of Sciences, 2014; 1310(1):89-97. https://doi.org/10.1111/nyas.12370
  40. 40. Ferguson DO, Sekiguchi JM, Chang S, Frank, KM, Gao Y, Depinho RA, Alt FW. The nonhomologous end-joining pathway of DNA repair is required for genomic stability and the suppression of translocations. PNAS, 2000; 97:6630-6633.
  41. 41. Gómez-Herreros F, Romero-Granados R, Zeng Z, Álvarez-Quilón A, Quintero C, Ju L, Umans L, Vermeire L, Huylebroeck D, Caldecott KW, Cortés-Ledesma F. TDP2-Dependent Non-Homologous End-Joining Protects against Topoisomerase II-Induced DNA Breaks and Genome Instability in Cells and In Vivo. PLoS Genetics, 2013; 9(3). https://doi.org/10.1371/journal.pgen.1003226
  42. 42. Heijink AM, Krajewska M, Van Vugt MATM. The DNA damage response during mitosis. Mutation Research - Fundamental and Molecular Mechanisms of Mutagenesis, 2013; 750(1-2):45-55. https://doi.org/10.1016/j.mrfmmm.2013.07.003
  43. 43. Ranjha L, Howard SM, and Cejka P. Main steps in DNA double-strand break repair: an introduction to homologous recombination and related processes. Chromosoma, 2018; 127(2):187-214. https://doi.org/10.1007/s00412-017-0658-1
  44. 44. Wray J, Williamson EA, Singh SB, Wu Y, Cogle CR, Weinstock DM, Zhang Y, Lee SH, Zhou D, Shao L, Hauer-Jensen M, Pathak R, Klimek V, Nickoloff JA, Hromas R. PARP1 is required for chromosomal translocations. Blood, 2013; 121(21) 4359-4365. https://doi.org/10.1182/blood-2012-10-460527
  45. 45. Nussenzweig A and Nussenzweig MC. Origin of chromosomal translocations in lymphoid cancer. Cell, 2010; 141(1):27-38. https://doi.org/10.1016/j.cell.2010.03.016
  46. 46. Yannone SM, Khan IS, Zhou RZ, Zhou T, Valerie K, Povirk LF. Coordinate 5′ and 3′ endonucleolytic trimming of terminaly blocked blunt DNA double-strand break ends by Artemis nuclease and DNA dependent protein kinase. Nucleic Acids Res. 2008; 36(10):3354-3365.
  47. 47. Chang HHY, Pannunzio NR, Adachi N, and Lieber MR. Non-homologous DNA end joining and alternative pathways to double-strand break repair. Nature Reviews Molecular Cell Biology, 2017;18(8). https://doi.org/10.1038/nrm.2017.48
  48. 48. Libura J, Ward M, Solecka J, and Richardson C. Etoposide-initiated MLL rearrangements detected at high frequency in human primitive hematopoietic stem cells with in vitro and in vivo long-term repopulating potential. Eur J Haematology, 2008; 81(3):185-195. https://doi.org/10.1111/j.1600-0609.2008.01103.x
  49. 49. Zhang Y and Jasin M. An essential role for CtIP in chromosomal translocation formation through an alternative end-joining pathway. Nature Structural & Molecular Biology, 2011; 18(1):80-84. https://doi.org/10.1038/nsmb.1940
  50. 50. Wright WD, Shah SS, Heyer WD. Homologous recombination and the repair of DNA double-strand breaks. JBC, 2018; 293(27):10524-10535. https://doi.org/10.1074/jbc.TM118.000372
  51. 51. Rouse J and Jackson SP. Interfaces between the detection, signaling, and repair of DNA damage. Science, 2002; 297(5581):547-551. https://doi.org/10.1126/science.1074740
  52. 52. Pommier Y, Su, Y, Huang, SYN, Nitiss, JL. Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nature Reviews Molecular Cell Biology, 2016; 17(11):703-721. https://doi.org/10.1038/nrm.2016.111
  53. 53. Nambiar M, and Raghavan SC. How does DNA break during chromosomal translocations? Nucleic Acids Research, 2011; 39(14):5813-5825. https://doi.org/10.1093/nar/gkr223
  54. 54. Donnianni RA, Zhou Z-X, Lujan SA, Al-Zain A, Garcia V, Glancy E, Burkholder AB, Kunkel TA, Symington LS. DNA Polymerase Delta synthesizes both strands during Break-induced Replications. Mol Cell. 2019; 76(3):371-381.
  55. 55. Berger JM. Type II DNA topoisomerases. Current Opinion in Structural Biology, 1998; 8(1):26-32. https://doi.org/10.1021/acs.jmedchem.6b00966
  56. 56. Deweese JE and Osheroff N. The DNA cleavage reaction of topoisomerase II: Wolf in sheep’s clothing. Nucleic Acids Research, 2009; 37(3):738-748. https://doi.org/10.1093/nar/gkn937
  57. 57. Nitiss JL. DNA topoisomerase II and its growing repertoire of biological functions. Nature Reviews Cancer, 2009; 9(5): 327-337. https://doi.org/10.1038/nrc2608
  58. 58. Nitiss JL and Nitiss KC. Tdp2: A Means to Fixing the Ends. PLoS Genetics, 2013; 9(3): 2-4. https://doi.org/10.1371/journal.pgen.1003370
  59. 59. Schoeffler A J and Berger JM. DNA topoisomerases: Harnessing and constraining energy to govern chromosome topology. Quarterly Reviews of Biophysics, 2008; 41(1):41-101. https://doi.org/10.1017/S003358350800468X
  60. 60. Jiang N, Doseff AI, Grotewold E. Flavones: From biosynthesis to health benefits. Plants, 2016; 5(2):1-1256. https://doi.org/10.3390/plants5020027
  61. 61. Manach C, Scalbert A, Morand C, Rémésy C, and Jiméne, L. Polyphenols: Food sources and bioavailability. American Journal of Clinical Nutrition, 2004; 79(5):727-747. https://doi.org/10.1093/ajcn/79.5.727
  62. 62. Esch HL, Kleider C, Scheffler A, and Lehmann L. Isoflavones: Toxicological Aspects and Efficacy. In Nutraceuticals: Efficacy, Safety and Toxicity. 2016. https://doi.org/10.1016/B978-0-12-802147-7.00034-6
  63. 63. Spagnuolo C, Russo GL, Orhan IE, Habtemariam S, Daglia M, Sureda A, Nabavi SF, Devi KP, Loizzo MR, Tundis R, and Nabavi SM. Genistein and cancer: current status, challenges, and future directions. Advances in Nutrition, 2015; 6(4):408-419. https://doi.org/10.3945/an.114.008052
  64. 64. Setchell KD, Zimmer-Nechemias L, Cai J, Heubi JE. Exposure of infants to phyto-oestrogens from soy-based infant formula. Lancet, 1997; 350(9070): 23-7.
  65. 65. Whitten PL, Patisaul HB. Cross-species and interassay comparisons of phytoestrogen action. Environ Health Perspect, 2001; 109 (Suppl 1):5-20.
  66. 66. Horn-Ross PL, John EM, Lee M, Stewart SL, Koo J, Sakoda LC, Shiau AC, Goldstein J, Davis P, Perez-Stable EJ. Phytoestrogen consumption and breast cancer risk in a multiethnic population: the Bay Area Breast Cancer Study. Am J Epidemiol, 2001; 154(5):434-441.
  67. 67. Mulligan AA, Welch AA, McTaggart AA, Bhaniani A, Bingham SA. Intakes and sources of soya foods and isoflavones in a UK population cohort study (EPIC-Norfolk). Eur J Clin Nutr, 2007; 61(2):248-254.
  68. 68. Kumar S and Pandey AK. Chemistry and biological activities of flavonoids: an overview. The Scientific World Journal, 2013; 162750. https://doi.org/10.1155/2013/162750
  69. 69. Kumar S and Pandey AK. Chemistry and biological activities of flaovnoids: An overview. Scientific World Journal, 2013. doi: 10.1155/2013/162750
  70. 70. Aherne SA and O’Brien NM. Dietary Flavonols: Chemistry, Food Content, and Metabolism in Chemistry and Structure of the Flavonoids. Nutrition, 2002; 18(1): 75-81.
  71. 71. Pollastri S and Tattini M. Flavonols: Old compounds for old roles. Annals of Botany, 2011; 108(7):1225-1233. https://doi.org/10.1093/aob/mcr234
  72. 72. D’Andrea G. Quercetin: A flavonol with multifaceted therapeutic applications? Fitoterapia, 2015; 106:256-271. https://doi.org/10.1016/j.fitote.2015.09.018
  73. 73. Devi KP, Malar DS, Nabavi SF, Sureda A, Xiao J, Nabavi SM, Daglia M. Kaempferol and inflammation: From chemistry to medicine. Pharmacological Research, 2015; 99:1-10. https://doi.org/10.1016/j.phrs.2015.05.002
  74. 74. Cadierno V. Metal-Catalyzed Routes for the Synthesis of Furocoumarins and Coumestans, 2015;77-100. Elsevier. https://doi.org/https://doi.org/10.1016/B978-0-12-800070-0.00004-9
  75. 75. Sun NJ, Woo SH, Cassady JM, and Snapka RM. DNA polymerase and topoisomerase II inhibitors from Psoralea corylifolia. Journal of Natural Products, 1998; 61(3):362-366. https://doi.org/10.1021/np970488q
  76. 76. Chen LG, Yang LL, Wang CC. Anti-inflammatory activity of mangostins from Garcinia mangostana. Food and Chemical Toxicology, 2008; 46(2):688-693. https://doi.org/10.1016/j.fct.2007.09.096
  77. 77. Di Micco S, Masullo M, Bandak AF, Berger JM, Riccio R, Piacente S, Bifulco G. Garcinol and Related Polyisoprenylated Benzophenones as Topoisomerase II Inhibitors: Biochemical and Molecular Modeling Studies. Journal of Natural Products, 2019; 82(10):2768-2779. https://doi.org/10.1021/acs.jnatprod.9b00382
  78. 78. Ashley RE. Natural-products-as-topoisomerase-II-poisons-Effects-of-thymoquinone-on-DNA-cleavage-mediated-by-human-topoisomerase-IIα, 2014. Chemical-Research-in-Toxicology.pdf.
  79. 79. Hordyjewska A, Ostapiuk A, Horecka A, Kurzepa J. Betulin and betulinic acid: triterpenoids derivatives with a powerful biological potential. Phytochemistry Reviews, 2019; 18(3):929-951. https://doi.org/10.1007/s11101-019-09623-1
  80. 80. Sawai S and Saito K. Triterpenoid biosynthesis and engineering in plants. Frontiers in Plant Science, 2011; 2:25. https://doi.org/10.3389/fpls.2011.00025
  81. 81. Wada SI, Iida A, Tanaka R. Screening of triterpenoids isolated from Phyllanthus flexuosus for DNA topoisomerase inhibitory activity. Journal of Natural Products, 2001; 64(12):1545-1547. https://doi.org/10.1021/np010176u
  82. 82. Wada SI and Tanaka R. Betulinic acid and its derivatives, potent DNA topoisomerase II inhibitors, from the bark of Bischofia javanica. Chemistry and Biodiversity, 2005; 2(5):689-694. https://doi.org/10.1002/cbdv.200590045
  83. 83. Abdel Bar FM, Khanfar MA, Elnagar AY, Liu H, Zaghloul AM, Badria FA, Sylvester PW, Ahmad KF, Raisch KP, El Sayed KA. Rational design and semisynthesis of betulinic acid analogues as potent topoisomerase inhibitors. Journal of Natural Products, 2009; 72(9):1643-1650. https://doi.org/10.1021/np900312u
  84. 84. Furbacher TR and Gunatilaka AAL. Catalytic inhibition of topoisomerase IIα by demethylzeylasterone, a 6-oxophenolic triterpenoid from Kokoona zeylanica. Journal of Natural Products, 2001; 64(10):1294-1296. https://doi.org/10.1021/np010123c
  85. 85. Bender RP, Lehmler HJ, Robertson LW, Ludewig G, Osheroff N. Polychlorinated biphenyl quinone metabolites poison human topoisomerase IIalpha: altering enzyme function by blocking the N-terminal protein gate. Biochemistry 2006; 45(33):10140-10152.
  86. 86. Bandele OJ, Clawson SJ, Osheroff N. Dietary polyphenols as topoisomerase II poisons: B ring and C ring substituents determine the mechanism of enzyme-mediated DNA cleavage enhancement. Chemical Research in Toxicology, 2008; 21(6):1253-1260. https://doi.org/10.1021/tx8000785
  87. 87. Bariar B, Vestal CG, Deem B, Goodenow D, Engledove RW, Sahyouni M, Richardson C. Bioflavonoids promote stable translocations between MLL - AF9 breakpoint cluster regions independent of normal chromosomal context: Model system to screen environmental risks. Env Mol Mutagenesis, 2019; 60(2):154-167. https://doi.org/10.0.3.234/em.22245
  88. 88. Pendleton M, Lindsey RH, Felix CA, Grimwade D, Osheroff N. Topoisomerase II and leukemia. Annals of the New York Academy of Sciences, 2014; 1310(1):98-110. https://doi.org/10.1111/nyas.12358
  89. 89. Smart DJ, Halicka HD, Schmuck G, Traganos F, Darzynkiewicz Z, Williams GM. Assessment of DNA double-strand breaks and γH2AX induced by the topoisomerase II poisons etoposide and mitoxantrone. Mutation Research - Fundamental and Molecular Mechanisms of Mutagenesis, 2008; 641(1-2):43-47. https://doi.org/10.1016/j.mrfmmm.2008.03.005
  90. 90. O’Prey J, Brown J, Fleming J, Harrison PR. Effects of dietary flavonoids on major signal transduction pathways in human epithelial cells. Biochem Pharmacol, 2003; 66(11):2075-2088.
  91. 91. Shimada M, Kato A, Habu T, Komatsu K. Genistein, isoflavonoids in soybeans, prevents the formation of excess radiation-induced centrosomes via p21 up-regulation. Mutat Res; 2011; 716(1-2):27-32.
  92. 92. Clewell RA, Sun B, Adeleye Y, Carmichael P, Efremenko A, McMullen PD, Pendse S, Trask OJ, White A, Andersen ME. Profiling dose-dependent activation of p53-mediated signaling pathways by chemicals with distinct mechanisms of DNA damage. Toxicol Sci, 2014; 142(1):56-73.
  93. 93. Charles C, Nachtergael A, Ouedraogo M, Belayew A, Duez P. Effects of chemopreventive natural products on non-homologous end-joining DNA double-strand break repair. Mutat Res Genet Toxicol Environ Mutagen, 2014; 768:33-41.
  94. 94. Schellenberg MJ, Lieberman JA, Herrero-Ruiz A, Butler LR, Williams JG, Muñoz-Cabello AM, Mueller GA, London RE, Cortés-Ledesma F, Williams RS. ZATT (ZNF451)–mediated resolution of topoisomerase 2 DNA-protein cross-links. Science, 2017; 357(6358):1412-1416. https://doi.org/10.1126/science.aam6468
  95. 95. Rübe CE, Fricke A, Widmann TA, Fürst T, Madry H, Pfreundschuh M, Rübe C. Accumulation of DNA damage in hematopoietic stem and progenitor cells during human aging. PloS One, 2011; 6(3): e17487–e17487. https://doi.org/10.1371/journal.pone.0017487
  96. 96. Ghosh A and Richardso, C. Role of XRCC4 downregulation in bioflavonoid-induced chromosomal translocations. Cancer Research, 79(13 Supplement), 2019; 2571. https://doi.org/10.1158/1538-7445.AM2019-2571
  97. 97. Biechonski S, Gourevich D, Rall M, Aqaqe N, Yassin M, Zipin-Roitman A, Trakhtenbrot L, Olender L, Raz Y, Jaffa AJ, Grisaru D, Wiesmuller L, Elad D, Milyavsky, M. Quercetin alters the DNA damage response in human hematopoietic stem and progenitor cells via TopoII- and PI3K-dependent mechanisms synergizing in leukemogenic rearrangements. Intl J of Cancer, 2017; 140(4): 864-876. https://doi.org/10.1002/ijc.30497
  98. 98. Aparicio T, Baer R, Gottesman M, Gautier J. MRN, CtIP, and BRCA1 mediate repair of topoisomerase II-DNA adducts. Journal of Cell Biology, 2016; 212(4):399-408. https://doi.org/10.1083/jcb.201504005
  99. 99. Atkin ND, Raimer HM, Wang YH. Broken by the Cut: A Journey into the Role of Topoisomerase II in DNA Fragility. Genes, 2019; 10(10):791. https://doi.org/10.3390/genes10100791
  100. 100. Ledesma FC, El Khamisy SF, Zuma MC, Osborn K, Caldecott KW. A human 5′-tyrosyl DNA phosphodiesterase that repairs topoisomerase-mediated DNA damage. Nature, 2009; 461(7264):674-678. https://doi.org/10.1038/nature08444
  101. 101. Nitiss KC, Nitiss JL, Hanakahi LA. DNA Damage by an essential enzyme: A delicate balance act on the tightrope. DNA Repair, 2019; 82:102639. https://doi.org/10.1016/j.dnarep.2019.102639
  102. 102. Hartsuiker E, Neale MJ, Carr AM. Distinct requirements for the Rad32(Mre11) nuclease and Ctp1(CtIP) in the removal of covalently bound topoisomerase I and II from DNA. Mol Cell 2009; 33(1):117-123.
  103. 103. Lee KC, Padget K, Curtis H, Cowell IG, Moiani D, Sondka Z, Morris NJ, Jackson GH, Cockell SJ, Tainer JA, Austin CA. MRE11 facilitates the removal of human topoisomerase II complexes from genomic DNA. Biol Open, 2012; 1(9):863-873.
  104. 104. Gohil K and Packer L. Bioflavonoid-rich botanical extracts show antioxidant and gene regulatory activity. Annals of the New York Academy of Sciences, 2002; 957: 70-77. https://doi.org/10.1111/j.1749-6632.2002.tb02906.x
  105. 105. Gerritsen ME, Carley WW, Ranges GE, Shen CP, Phan SA, Ligon GF, Perry CA. Flavonoids inhibit cytokine-induced endothelial cell adhesion protein gene expression. Am J Pathol, 1995; 147(2): 278-292.
  106. 106. Cho S, Park S, Kwon M, Jeong TS, Bok S, Choi W, Jeong W, Ryu S, Do S, Lee C, Song J, Jeong K. Quercetin suppresses proinflammatory cytokines production through MAP kinases and NF- κ B pathway in lipopolysaccharide-stimulated macrophage. Molecular and Cellular Biochemistry, 2003; 243 153-160.
  107. 107. Chou RH, Hsieh SC, Yu YL, Huang MH, Huang YC, Hsieh YH. Fisetin Inhibits Migration and Invasion of Human Cervical Cancer Cells by Down-Regulating Urokinase Plasminogen Activator Expression through Suppressing the p38 MAPK-Dependent NF-κB Signaling Pathway. PLoS ONE, 2013; 8(8): 1-12. https://doi.org/10.1371/journal.pone.0071983
  108. 108. Ishikawa Y, Sugiyama H, Stylianou E, Kitamura M. Bioflavonoid quercetin inhibits interleukin-1-induced transcriptional expression of monocyte chemoattractant protein-1 in glomerular cells via suppression of nuclear factor-kappaB. J Amer Soc Nephrology, 1999; 10(11):2290-2296.
  109. 109. Phillips PA, Sangwan V, Borja-Cacho D, Dudeja V, Vickers SM, Saluja AK. Myricetin induces pancreatic cancer cell death via the induction of apoptosis and inhibition of the phosphatidylinositol 3-kinase (PI3K) signaling pathway. Cancer Letters, 2011; 308(2):181-188. https://doi.org/10.1016/j.canlet.2011.05.002
  110. 110. Russo GL, Vastolo V, Ciccarelli M, Albano L, Macchia E, Ungaro P, Luigi G, Vastolo V, Ciccarelli M, Albano L. Dietary polyphenols and chromatin remodeling. Critical Reviews in Food Science and Nutrition, 2017; 57(12):2589-2599. https://doi.org/10.1080/10408398.2015.1062353
  111. 111. Vanhees K, Coort S, Ruijters EJB, Godschalk RWL, Schooten FJ, Van Waalwijk van Doorn-Khosrovani SB. Epigenetics: prenatal exposure to genistein leaves a permanent signature on the hematopoietic lineage. The FASEB Journal, 2011; 25(2):797-807. https://doi.org/10.1096/fj.10-172155
  112. 112. Vanhees K, Godschalk RW, Sanders A, Van Waalwijk van Doorn-Khosrovani, SB, Van Schooten FJ. Maternal quercetin intake during pregnancy results in an adapted iron homeostasis at adulthood. Toxicology, 2011; 290(2-3):350-358. https://doi.org/10.1016/j.tox.2011.10.017
  113. 113. Nabavi SF, Devi KP, Loizzo MR, Tundis R, Nabavi SM. Genistein and Cancer: Current Status and Challenges. Advances in Nutrition, 2015; 6:408-19. https://doi.org/10.3945/an.114.008052.408
  114. 114. Vafadar A, Shabaninejad Z, Movahedpour A, Fallahi F, Taghavipour M, Ghasemi Y, Akbari M, Shafiee A, Hajighadimi S, Moradizarmehri S, Razi E, Savardashtaki A, Mirzaei H. Quercetin and cancer: New insights into its therapeutic effects on ovarian cancer cells. Cell and Bioscience, 2020; 10(1):1-17. https://doi.org/10.1186/s13578-020-00397-0
  115. 115. Kundur S, Prayag A, Selvakumar P, Nguyen H, McKee L, Cruz C, Srinivasan A, Shoyele S, Lakshmikuttyamma A. Synergistic anticancer action of quercetin and curcumin against triple-negative breast cancer cell lines. Journal of Cellular Physiology, 2019;234(7):11103-11118. https://doi.org/10.1002/jcp.27761
  116. 116. López-Lázaro M, Willmore E, Austin CA. The dietary flavonoids myricetin and fisetin act as dual inhibitors of DNA topoisomerases I and II in cells. Mutation Research - Genetic Toxicology and Environmental Mutagenesis, 2010;696(1):41-47. https://doi.org/10.1016/j.mrgentox.2009.12.010
  117. 117. Cook MT. Mechanism of metastasis suppression by luteolin in breast cancer. Breast Cancer (Dove Medical Press), 2018; 10:89-100. https://doi.org/10.2147/BCTT.S144202
  118. 118. Li H, Lin D, Kuang G, Wan J, Zhang X, Li H, Xia G. Luteolin suppresses the metastasis of triple-negative breast cancer by reversing epithelial-to-mesenchymal transition via downregulation of β-catenin expression. Oncology Reports, 2017; 37(2):895-902. https://doi.org/10.3892/or.2016.5311
  119. 119. Ferry DR, Smith A, Malkhandi J, Fyfe D W, deTakats PG, Anderson D, Baker J, Kerr DJ. Phase I clinical trial of the flavonoid quercetin: pharmacokinetics and evidence for in vivo tyrosine kinase inhibition. Clinical Cancer Research, 1996; 2(4):659-668.
  120. 120. Pombo-de-Oliveira MS, Koifman S, Brazilian Collaborative Study Group of Infant Acute L. 2006. Infant acute leukemia and maternal exposures during pregnancy. Cancer Epidemiol Biomarkers Prev 2006; 15(12):2336-2341.
  121. 121. Strick R, Strissel PL, Borgers S, Smith SL, Rowley JD. Dietary bioflavonoids induce cleavage in the MLL gene and may contribute to infant leukemia. PNAS, 2000; 97(9):4790-4795. https://doi.org/10.1073/pnas.070061297
  122. 122. Spector LG, Xie Y, Robison LL, Heerema NA, Hilden JM, Lange B, Felix CA, Davies SM, Slavin J, Potter JD, Blair CK, Reaman GH, Ross JA. Maternal Diet and Infant Leukemia: The DNA Topoisomerase II Inhibitor Hypothesis: A Report from the Children’s Oncology Group. Cancer Epidemiology Biomarkers & Prevention, 2005; 14(3):651-655. https://doi.org/10.1158/1055-9965.EPI-04-0602
  123. 123. Ross JA, Kasum CM. Dietary flavonoids: bioavailability, metabolic effects, and safety. Annu Rev Nutr 2002; 22:19-34.
  124. 124. Schroder-van der Elst JP, van der Heide D, Rokos H, Morreale de Escobar G, Kohrle J. Synthetic flavonoids cross the placenta in the rat and are found in fetal brain. Am J Physiol, 1998; 274:E253-256.
  125. 125. Doerge DR, Churchwell MI, Chang HC, Newbold RR, Delclos KB. Placental transfer of the soy isoflavone genistein following dietary and gavage administration to Sprague Dawley rats. Reprod Toxicol, 2001; 15(2):105-110.
  126. 126. Alexander FE, Patheal SL, Biondi A, Brandalise S, Cabrera ME, Chan LC, Chen Z, Cimino G, Cordoba JC, Gu LJ, Hussein H, Ishii E, Kamel AM, Labra S, Magalhaes IQ , Mizutani S, Petridou E, de Oliveira MP, Yuen P, Wiemels JL, Greaves MF. Transplacental chemical exposure and risk of infant leukemia with MLL gene fusion. Cancer Res, 2001; 61(6):2542-2546.
  127. 127. Pombo-de-Oliveira MS, Koifman S. Infant acute leukemia and maternal exposures during pregnancy. Cancer Epidemiol Biomarkers Prev, 2006; 15(12):2336-2341.
  128. 128. Todaka E, Sakurai K, Fukata H, Miyagawa H, Uzuki M, Omori M, Osada H, Ikezuki Y, Tsutsumi O, Iguchi T, Mori C. Fetal exposure to phytoestrogens--the difference in phytoestrogen status between mother and fetus. Environ Res, 2005; 99(2):195-203.
  129. 129. Zandvliet DW, Hanby AM, Austin CA, Marsh KL, Clark IB, Wright NA, Poulsom R. Analysis of foetal expression sites of human type II DNA topoisomerase alpha and beta mRNAs by in situ hybridisation. Biochim Biophys Acta, 1996; 1307(2):239-247.

Written By

Donna Goodenow, Kiran Lalwani and Christine Richardson

Submitted: 28 May 2020 Reviewed: 10 December 2020 Published: 13 January 2021