Hostname: page-component-7c8c6479df-94d59 Total loading time: 0 Render date: 2024-03-26T22:33:29.239Z Has data issue: false hasContentIssue false

Flavonoids: modulators of brain function?

Published online by Cambridge University Press:  01 May 2008

Jeremy P. E. Spencer*
Affiliation:
Molecular Nutrition Group, School of Chemistry, Food and Pharmacy, University of Reading, ReadingRG2 6AP, UK
*
*Corresponding author: Dr J. P. E. Spencer, fax +44 118 931 0080, email j.p.e.spencer@reading.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Emerging evidence suggests that dietary phytochemicals, in particular flavonoids, may exert beneficial effects on the central nervous system by protecting neurons against stress-induced injury, by suppressing neuroinflammation and by improving cognitive function. It is likely that flavonoids exert such effects, through selective actions on different components of a number of protein kinase and lipid kinase signalling cascades, such as the phosphatidylinositol-3 kinase (PI3K)/Akt, protein kinase C and mitogen-activated protein kinase (MAPK) pathways. This review explores the potential inhibitory or stimulatory actions of flavonoids within these pathways, and describes how such interactions are likely to underlie neurological effects through their ability to affect the activation state of target molecules and/or by modulating gene expression. Future research directions are outlined in relation to the precise site(s) of action of flavonoids within signalling pathways and the sequence of events that allow them to regulate neuronal function.

Type
Full Papers
Copyright
Copyright © The Author 2008

Representing one of the most important lifestyle factors, diet can strongly influence the incidence and onset of cardiovascular and neurodegenerative diseases, and thus a healthy diet is an essential factor for healthy ageing. Various phytochemical constituents of foods and beverages, in particular a class of photochemicals known as flavonoids, have been avidly investigated in recent years. A number of dietary intervention studies in humans and animals, in particular those with foods and beverages derived from Vitis vinifera (grape), Camellia sinensis (tea), Theobroma cacao (cocoa) and Vaccinium spp. (blueberry) have demonstrated beneficial effects on vascular function and mental performance. While such foods and beverages differ greatly in chemical composition, macro- and micro-nutrient content and caloric load per serving, they have in common that they are amongst the major dietary sources of flavonoids. Dietary intervention studies in several mammalian species, including humans, using flavonoid-rich plant or food extracts have indicated an ability of these dietary components to improve memory and learning(Reference Youdim and Joseph1Reference Wang, Wang, Wu and Cai7), by protecting vulnerable neurons, enhancing existing neuronal function or by stimulating neuronal regeneration. Their neuroprotective potential has also been demonstrated in both oxidative stress-(Reference Inanami, Watanabe, Syuto, Nakano, Tsuji and Kuwabara8) and Aβ-induced-neuronal death models(Reference Luo, Smith, Paramasivam, Burdick, Curry, Buford, Khan, Netzer, Xu and Butko9) and evidence supports the beneficial and neuromodulatory effects of flavonoid-rich ginkgo biloba extracts, particularly in connection with age-related dementias and Alzheimer's disease(Reference Luo, Smith, Paramasivam, Burdick, Curry, Buford, Khan, Netzer, Xu and Butko9Reference Zimmermann, Colciaghi, Cattabeni and Di Luca11). Furthermore, the citrus flavanone, tangeretin, has been observed to help maintain nigrostriatal integrity and functionality following lesioning with 6-hydroxydopamine, suggesting that flavonoids may also serve as potential neuroprotective agents against the underlying pathology associated with Parkinson's disease(Reference Datla, Christidou, Widmer, Rooprai and Dexter12).

Historically, the biological actions of flavonoids have been attributed to their antioxidant properties, either through their reducing capacities per se or through their possible influences on intracellular redox status(Reference Rice-Evans, Miller and Paganga13). However, it has been speculated that their classical hydrogen donating antioxidant activity cannot explain the bioactivity of flavonoids in vivo. Indeed, it has become evident that flavonoids are more likely to exert their neuroprotective actions by (1) the modulation of intracellular signalling cascades which control neuronal survival, death and differentiation; (2) affecting gene expression and (3) interactions with mitochondria(Reference Williams, Spencer and Rice-Evans14Reference Schroeter, Boyd, Spencer, Williams, Cadenas and Rice-Evans16). This review will emphasize the potential of flavonoids to exert beneficial effects in the brain by preventing neurodegeneration, inhibiting neuroinflammation and reducing age-related cognitive decline. In particular, it will highlight probable mechanisms which underpin such actions in the brain, including their interactions with neuronal intracellular signalling pathways vital in determining neuronal death, survival, differentiation and proliferation.

Flavonoid structure, source and metabolism

Flavonoids comprise the most common group of polyphenolic compounds in the human diet and are found ubiquitously in plants. Major dietary sources of flavonoids include fruits, vegetables, cereals, tea, wine and fruit juices [reviewed in (Reference Manach, Scalbert, Morand, Remesy and Jimenez17)]. Flavonoids consist of two aromatic carbon rings, benzopyran (A and C rings) and benzene (B ring), and may be divided in six subgroups based on the degree of the oxidation of the C-ring, the hydroxylation pattern of the ring structure and the substitution of the 3-position. The main dietary groups of flavonoids are (1) flavonols (e.g. kaempferol, quercetin), which are found in onions, leeks, broccoli, (2) flavones (e.g. apigenin, luteolin), which are found in parsley and celery, (3) isoflavones (e.g. daidzein, genistein), which are mainly found on soy and soy products, (4) flavanones (e.g. hesperetin, naringenin), which are mainly found in citrus fruit and tomatoes, (5) flavanols (e.g. catechin, epicatechin, epigallocatechin, epigallocatechin gallate (EGCG), which are abundant in green tea, red wine, chocolate, and (6) anthocyanidins (e.g. pelargonidin, cyanidin, malvidin), whose sources include red wine and berry fruits. Further information regarding the structure and classes of flavonoids may be found in the thorough review by Manach et al. (Reference Manach, Scalbert, Morand, Remesy and Jimenez17)

Although flavonoids display potent antioxidant capacity in vitro (Reference Rice-Evans, Miller and Paganga13, Reference Rice-Evans18, Reference Rice-Evans19), during absorption they are extensively metabolised, resulting in significant alteration of their redox potentials. For example, the majority of flavonoid glycosides, and in some instances the aglycones, present in plant-derived foods, are extensively conjugated and metabolised during absorption [reviewed in (Reference Spencer, Schroeter, Rechner and Rice-Evans20Reference Donovan, Waterhouse, Rice-Evans and Packer24)]. In particular, there is much evidence for the extensive phase I de-glycosylation and phase II metabolism of the resulting aglycones to glucuronides, sulphates and O-methylated forms during transfer across the small intestine(Reference Spencer, Schroeter, Rechner and Rice-Evans20, Reference Spencer, Chowrimootoo, Choudhury, Debnam, Srai and Rice-Evans25) and then again in the liver. Further transformation has been reported in the colon where the enzymes of the gut microflora degrade flavonoids to simple phenolic acids(Reference Scheline26). In addition, flavonoids may undergo at least three types of intracellular metabolism: (1) oxidative metabolism, (2) P450-related metabolism and (3) conjugation with thiols, particularly GSH(Reference Spencer, Kuhnle, Williams and Rice-Evans27, Reference Spencer, Schroeter, Crossthwaithe, Kuhnle, Williams and Rice-Evans28). Circulating metabolites of flavonoids, such as glucuronides, sulphates and conjugated O-methylated forms, or intracellular metabolites like flavonoid-GSH adducts, have greatly reduced antioxidant potential(Reference Spencer, Schroeter, Crossthwaithe, Kuhnle, Williams and Rice-Evans28). Indeed, studies have indicated that although such conjugates and metabolites may participate directly in plasma antioxidant reactions and in scavenging reactive oxygen and nitrogen species in the circulation their effectiveness is reduced relative to their parent aglycones(Reference Miyake, Shimoi, Kumazawa, Yamamoto, Kinae and Osawa29Reference da Silva, Piskula, Yamamoto, Moon and Terao33).

Do flavonoids access the brain?

In order to understand whether flavonoids and their metabolic derivatives are capable acting as neuromodulators, it is crucial to ascertain whether they are able to enter the central nervous system. In order for flavonoids to enter into the brain, they must first cross the blood–brain barrier (BBB). The functions of the BBB include controlling the entry of xenobiotics into the brain and maintenance of the brain's microenvironment(Reference Abbott34). In vitro and in vivo studies have indicated that the flavanones hesperetin, naringenin and their relevant in vivo metabolites, as well as some dietary anthocyanins, cyanidin-3-rutinoside and pelargonidin-3-glucoside, are able to traverse the BBB(Reference Youdim, Dobbie, Kuhnle, Proteggente, Abbott and Rice-Evans35Reference Youdim, Qaiser, Begley, Rice-Evans and Abbott37). Furthermore, it appears that the potential for flavonoid penetration is dependent on compound lipophilicity(Reference Youdim, Dobbie, Kuhnle, Proteggente, Abbott and Rice-Evans35). Accordingly, it is plausible that the uptake of the less polar O-methylated metabolites, such as the O-methylated epicatechin metabolites (formed in the small intestine and liver), may be greater than the parent aglycone. For the same reason, the more polar glucuronidated metabolites, which have lower BBB permeability values(Reference Youdim, Dobbie, Kuhnle, Proteggente, Abbott and Rice-Evans35), may not be able to access the brain. However, evidence exists to suggest that certain drug glucuronides may cross the BBB(Reference Aasmundstad, Morland and Paulsen38) and exert pharmacological effects(Reference Sperker, Backman and Kroemer39, Reference Kroemer and Klotz40), suggesting that there may be a specific uptake mechanism for glucuronides in vivo. Apart from the flavonoids lipophilicity, their ability to enter the brain may also be influenced by their interactions with specific efflux transporters expressed in the BBB. One such transporter is P-glycoprotein, which plays an important role in drug absorption and brain uptake(Reference Lin and Yamazaki41) and appears to be responsible for the differences between naringenin and quercetin flux into the brain in situ (Reference Youdim, Qaiser, Begley, Rice-Evans and Abbott37).

Animal feeding studies also indicate that flavonoids may access the brain, with the tea flavanol, EGCG, being reported to access the brain after oral administration to mice(Reference Suganuma, Okabe, Oniyama, Tada, Ito and Fujiki42). Furthermore, oral ingestion of pure epicatechin resulted in the detection of epicatechin glucuronide and 3′-O-methyl-epicatechin glucuronide in rat brain tissue(Reference Abd El Mohsen, Kuhnle, Rechner, Schroeter, Rose, Jenner and Rice-Evans43). Anthocyanidins have also been detected in the brain after oral administration(Reference Talavera, Felgines, Texier, Besson, Gil-Izquierdo, Lamaison and Remesy44, Reference El Mohsen, Marks, Kuhnle, Moore, Debnam, Kaila, Rice-Evans and Spencer45), with several anthocyanidins being identified in different regions of rat brain after the animals were fed with blueberry(Reference Andres-Lacueva, Shukitt-Hale, Galli, Jauregui, Lamuela-Raventos and Joseph46). Such flavonoid localisation has been correlated with increased cognitive performance, suggesting a central neuroprotective role of these components. Despite their apparent ability to access the brain, it is clear that the concentrations of flavonoids and their metabolite forms accumulated in vivo (Reference Abd El Mohsen, Kuhnle, Rechner, Schroeter, Rose, Jenner and Rice-Evans43) are lower (high nm, low μm) than those recorded for small molecule antioxidant nutrients such as ascorbic acid and α-tocopherol(Reference Halliwell, Zhao and Whiteman47). Consequently, the beneficial effects of flavonoid metabolites in vivo are unlikely to result due to their ability to out-compete antioxidants such as ascorbate, which are present at higher concentrations (high μm to mm). However, evidence has accumulated to suggest that the cellular effects of flavonoids may be mediated by their interactions with specific proteins central to intracellular signalling cascades(Reference Schroeter, Boyd, Spencer, Williams, Cadenas and Rice-Evans16), such as the mitogen-activated protein kinase (MAP kinase) signalling pathway and the phosphoinositide 3-kinase (PI3 kinase/Akt) signalling cascade (Fig. 1).

Fig. 1 Overview of MAP kinase and Akt/PKB signalling cascades in neurons. Flavonoid-induced activation of ERK1/2 or PI3K/Akt pathways acts to stimulate neuronal survival and/or enhance synaptic plasticity and long-term potentiation relevant to the laying down of memory. In addition, inhibitory actions within JNK and p38 pathways are likely to be neuroprotective in the presence of stress.

Improvement in memory, learning and cognitive performance

There is a growing interest in the potential of phytochemicals to improve memory, learning and general cognitive ability. Previous studies have indicated that phytochemical-rich foods such as berries and spinach are effective at reversing age-related deficits in motor function and spatial working memory(Reference Galli, Shukitt-Hale, Youdim and Joseph3, Reference Joseph, Shukitt-Hale, Denisova, Prior, Cao, Martin, Taglialatela and Bickford48Reference Joseph, Denisova, Arendash, Gordon, Diamond, Shukitt-Hale and Morgan54). For example, the latency period to find a platform and the distance swam to a platform in a Morris water maze task, are significantly reduced following blueberry supplementation(Reference Joseph, Shukitt-Hale, Denisova, Prior, Cao, Martin, Taglialatela and Bickford48, Reference Joseph, Shukitt-Hale, Denisova, Bielinski, Martin, McEwen and Bickford49). Such results may suggest favourable effects of the blueberry diet on locomotor activity in old animals(Reference Andres-Lacueva, Shukitt-Hale, Galli, Jauregui, Lamuela-Raventos and Joseph46, Reference Ramirez, Izquierdo, do Carmo Bassols, Zuanazzi, Barros and Henriques55). However, reductions in the time taken to make a choice may also reflect an improved memory component, where rats ‘remembered’ more rapidly and thus responded quicker. Presently, it is unclear how phytochemicals may exert such effects, although they may be linked to antioxidant actions, the modulation of neurotransmitter release(Reference Joseph, Shukitt-Hale, Denisova, Prior, Cao, Martin, Taglialatela and Bickford48, Reference Joseph, Shukitt-Hale, Denisova, Bielinski, Martin, McEwen and Bickford49), the stimulation of hippocampal neurogenesis(Reference Casadesus, Shukitt-Hale, Stellwagen, Zhu, Lee, Smith and Joseph50) via the modulation of signalling(Reference Joseph, Shukitt-Hale and Casadesus52, Reference Goyarzu, Malin and Lau53) or an ability to improve cerebrovascular blood flow(Reference Schroeter, Heiss, Balzer, Kleinbongard, Keen, Hollenberg, Sies, Kwik-Uribe, Schmitz and Kelm56).

Flavonoid-rich foods, in particular those containing flavanols, have been observed to improve peripheral blood flow and surrogate markers of cardiovascular function in humans(Reference Schroeter, Heiss, Balzer, Kleinbongard, Keen, Hollenberg, Sies, Kwik-Uribe, Schmitz and Kelm56). In the context of the CNS, brain imaging studies in humans have demonstrated that the consumption of flavanol-rich cocoa may enhance cortical blood flow(Reference Francis, Head, Morris and Macdonald57, Reference Fisher, Sorond and Hollenberg58). Increased cerebrovascular function, especially in the hippocampus, a brain region important for memory, may facilitate adult neurogenesis(Reference Gage59). Indeed, new hippocampal cells are clustered near blood vessels, proliferate in response to vascular growth factors and may influence memory(Reference Palmer, Willhoite and Gage60). As well as new neuronal growth, increases in neuronal spine density and morphology are considered vital for learning and memory(Reference Harris and Kater61). Changes in spine density, morphology and motility have been shown to occur with paradigms that induce synaptic, as well as altered sensory experience, and lead to alterations in synaptic connectivity and strength between neuronal partners, affecting the efficacy of synaptic communication. These events are mediated at the cellular and molecular level and are strongly correlated with memory and learning (Fig. 2).

Fig. 2 Formation of stable long-term potentiation at synapses. (1) Increased expression and release of BDNF from the synapse through enhanced CREB activation. BDNF binds to pre- and post-synaptic TrkB receptors (2), triggering glutamate release and PI3K/mTOR signalling and Arc synthesis (3). Sustained activation of mTOR leads to enhanced translational efficiency whilst Arc, in association with Cofilin, triggers F-actin expansion and synapse growth (mushroom synapse) (4).

The enhancement of both short-term and long-term memory is controlled at the molecular level(Reference Carew62). Whereas short-term memory involves covalent modifications of pre-existing proteins, long-term memory requires the synthesis of new mRNAs and proteins(Reference Martin, Barad and Kandel63, Reference Kelleher, Govindarajan and Tonegawa64). Four signalling pathways control this process: (i) cAMP-dependent protein kinase (protein kinase A), (ii) calcium-calmodulin kinases, (iii) protein kinase C, and (iv) mitogen-activated protein kinase (MAPK). All four pathways converge to signal to the cAMP-response element-binding protein (CREB), a transcription factor which binds to the promoter regions of many genes associated with memory and synaptic plasticity(Reference Impey, McCorkle, Cha-Molstad, Dwyer, Yochum, Boss, McWeeney, Dunn, Mandel and Goodman65, Reference Barco, Bailey and Kandel66) (Fig. 2). The importance of CREB activation in the induction of long-lasting changes in plasticity and memory are highlighted by studies which show that disruption of CREB activity specifically blocks the formation of long-term memory(Reference Bourtchuladze, Frenguelli, Blendy, Cioffi, Schutz and Silva67), whereas agents that increase the amount or activity of CREB accelerate the process(Reference Tully, Bourtchouladze, Scott and Tallman68). Furthermore, robust CREB phosphorylation and CRE-reporter gene expression are detected in cortical neurons during developmental plasticity(Reference Pham, Impey, Storm and Stryker69) and in hippocampal neurons in response to both LTP-inducing stimuli and memory training tasks(Reference Impey, Smith, Obrietan, Donahue, Wade and Storm70, Reference Impey, Mark, Villacres, Poser, Chavkin and Storm71). There is considerable interest in identifying safe effective agents that enhance the activity of CREB, as these may lead to an improvement in memory.

Previous studies have suggested that phytochemicals, especially flavonoids, may exert cellular effects via direct modulation of protein and lipid kinase signalling pathways(Reference Williams, Spencer and Rice-Evans14). Interactions within the MAPK pathway are thought to be central to mediating the cellular effects of flavonoids such as those found in berries, tea and cocoa(Reference Spencer, Rice-Evans and Williams15, Reference Schroeter, Spencer, Rice-Evans and Williams72). For example, the flavanol, ( − )-epicatechin, induces both ERK1/2 and CREB activation in cortical neurons and subsequently increases CREB regulated gene expression(Reference Schroeter, Bahia, Spencer, Sheppard, Rattray, Rice-Evans and Williams73) (Fig. 1). Furthermore, another flavonoid, fisetin, has recently been shown to improve long-term potentiation (LTP) and memory through a CREB/ERK mechanism(Reference Maher, Akaishi and Abe74). Thus, one potential mechanism action of flavonoids in modulating neuronal function, synaptic plasticity and synaptogenesis may proceed via signalling through CREB. If flavonoids are able to promote neuronal activation of CREB in vivo they may be capable of influencing the neuronal expression of a number of genes which contain cAMP-response element (CRE) sequences in their promoter regions(Reference Conkright, Guzman, Flechner, Su, Hogenesch and Montminy75). Particular emphasis has been given to the regulation of BDNF(Reference Tao, Finkbeiner, Arnold, Shaywitz and Greenberg76, Reference Shieh, Hu, Bobb, Timmusk and Ghosh77), which has been implicated in synaptic plasticity and long-term memory(Reference Bramham and Messaoudi78) and is robustly induced in hippocampal neurons upon synaptic stimulation(Reference Patterson, Grover, Schwartzkroin and Bothwell79). BDNF belongs to the neurotrophin family of growth factors and affects the survival and function of neurons in the central nervous system. Its secretion from neurons is under activity-dependent control and is crucial for the formation of appropriate synaptic connections during development and for learning and memory in adults(Reference Thomas and Davies80). Decreases in BDNF and pro-BDNF have been reported in Alzheimer's disease(Reference Peng, Wuu, Mufson and Fahnestock81, Reference Michalski and Fahnestock82) and the importance of pro-BDNF has been emphasized by the finding that a polymorphism that replaces valine for methionine at position 66 of the pro-domain is associated with memory defects and abnormal hippocampal function in humans(Reference Egan, Kojima and Callicott83). In addition, genetic(Reference Linnarsson, Bjorklund and Ernfors84) as well as pharmacological inhibition(Reference Mu, Li, Yao and Zhou85) of BDNF or its receptor, tropomyosin receptor kinase B (TrkB)(Reference Minichiello, Korte, Wolfer, Kuhn, Unsicker, Cestari, Rossi-Arnaud, Lipp, Bonhoeffer and Klein86), impairs learning and memory. On the other hand, agents that increase in BDNF levels may lead to improvements in spatial working memory, in part through the regulation of protein translation via the mTOR signalling pathway(Reference Wullschleger, Loewith and Hall87) (Fig. 2). BDNF is known to bind to the TrkB receptor either pre- or post-synaptically causing activation of the PI3 kinase/Akt signalling pathway, the phosphorylation of mTOR at Ser2448, the phosphorylation of NMDA receptors and the release of neurotransmitters from pre-synaptic sites(Reference Schratt, Nigh, Chen, Hu and Greenberg88).

Many of the BDNF-regulated mRNAs in mature neurons encode proteins that function at synapses(Reference Bramham and Messaoudi78). One such protein which is associated with LTP is activity-regulated cytoskeletal-associated protein, Arc/Arg31, which has been proposed to be under regulatory control of both BDNF(Reference Yin, Edelman and Vanderklish89) and the ERK signalling pathway(Reference Waltereit, Dammermann, Wulff, Scafidi, Staubli, Kauselmann, Bundman and Kuhl90). Sustained synthesis of Arc/Arg3.1 during a protracted time-window is necessary to consolidate LTP, whilst translation of pre-existing Arc/Arg3.1 mRNA contributes to early LTP expression and translation of new Arc/Arg3.1 mRNA mediates consolidation(Reference Soule, Messaoudi and Bramham91). Increased Arc/Arg3.1 expression may facilitate changes in synaptic strength, and the induction of morphological changes, such as that observed when small spines are converted into large mushroom-shaped spines through a mechanism dependent on actin polymerization(Reference Lyford, Yamagata, Kaufmann, Barnes, Sanders, Copeland, Gilbert, Jenkins, Lanahan and Worley92) (Fig. 2). Whether flavonoids are able to promote changes in neuronal morphology in vivo, via signalling through these pathways is currently unknown. However, the ability of flavonoids and their in vivo metabolites to exert effects on neuronal signalling cascades (dealt with in greater depth later in this review) suggest that they may be capable of inducing behavioural changes in memory, learning and cognitive performance through interactions with such pathways. In agreement with this, previous studies have indicated that certain flavonoids may influence neuronal dentrite outgrowth in vitro (Reference Reznichenko, Amit, Youdim and Mandel93).

Inhibition of neuroinflammation

There is increasing evidence to suggest that neuroinflammatory processes may contribute to the cascade of events leading to the progressive neuronal damage observed in Parkinson's disease and Alzheimer's disease(Reference Hirsch, Hunot and Hartmann94, Reference McGeer and McGeer95), and also with the neuronal injury associated with stroke(Reference Zheng, Lee and Yenari96). In support of this, observations suggest that the use of non-steroidal anti-inflammatory drugs, such as ibuprofen, may delay or even prevent the onset of neurodegenerative disorders, such as Parkinson disease(Reference Casper, Yaparpalvi, Rempel and Werner97, Reference Chen, Zhang, Hernan, Schwarzschild, Willett, Colditz, Speizer and Ascherio98). Activation of glial cells (astrocytes and microglia) plays a key role in the development of inflammatory neurodegeneration(Reference McGeer and McGeer99). Glial cells occupy the majority of the brain volume and play a key role in the maintenance of brain integrity and upon appropriate activation, glia respond to invading pathogens, eliminate cellular debris and promote cell repair and recovery(Reference Vila, Jackson-Lewis, Guegan, Wu, Teismann, Choi, Tieu and Przedborski100, Reference Kim and Joh101). However, excessive and chronic activation of glial cells may have harmful effects by triggering an inflammatory response that ultimately leads to progressive neuronal degeneration(Reference Allan and Rothwell102). Central to glial-induced neurotoxicity is the generation of NO via increases in the expression of iNOS (Fig. 3). The production of NO by glial cells is mediated by iNOS, and excessive NO may diffuse away from glial cells and induce neuronal cell damage by disrupting neuronal mitochondrial electron transport chain (ETC) function(Reference Stewart and Heales103) (Fig. 3). In particular, NO may selectively inhibit mitochondrial respiration at cytochrome c oxidase (complex IV) resulting in a disruption of neuronal adenosine 5′-triphosphate (ATP) synthesis and an increased generation of ROS(Reference Moncada and Bolanos104). Therefore, the uncontrolled activation of iNOS in glial cells is a critical step in inflammatory-mediated neurodegeneration. Inflammation is also characterised by increased cytokine production, such as IL-1β and TNF-α, which also act to stimulate iNOS and NO production(Reference Kozuka, Itofusa, Kudo and Morita105), and by the activation of NADPH oxidase which generates superoxide and hydrogen peroxide (H2O2)(Reference Bal-Price, Matthias and Brown106).

Fig. 3 Involvement of activated glial cells in neuroinflammation-induced neurodegeneration. Central to glial-induced neurotoxicity is the generation of NO via increases in the expression of iNOS. iNOS itself is induced by the cell surface CD23 receptor which is in turn activated by cytokines such as TNF-α and IL-1γ. NO may diffuse to neighbouring neurons where it inhibits mitochondrial respiration at cytochrome c oxidase. NO may also react with to generate ONOO−  which can cause damage to proteins, inhibit mitochondrial respiration and activate cell death genes and signalling pathways which may ultimately lead to neuronal death. Furthermore, cytokines such as TNF-α may cause direct cell death by binding to specific receptors expressed in neurons and subsequently activate genes that trigger the apoptotic pathway.

Importantly, the transcriptional and post-transcriptional regulation of iNOS and cytokines in activated glial cells is dependent on signalling through pathways such as MAPK, specifically through activation of ERK1/2(Reference Fiebich, Lieb, Engels and Heinrich107Reference Pawate and Bhat110) (Fig. 4). These observations suggest that pharmacological control of such signalling pathways may be a useful tool in the prevention/treatment of neurodegenerative diseases through their ability to modulate glial cell activation. The MEK inhibitor PD98059, which has significant structural homology with flavonoids, has been shown to effectively block iNOS expression and generation of NO∙(Reference Bhat, Zhang, Lee and Hogan108), suggesting that flavonoids may also be capable of exerting anti-inflammatory actions via inhibitory actions on MEK1 within the ERK signalling pathway (Fig. 4). Thus far, studies have indicated that flavanols(Reference Li, Huang, Fang and Le111, Reference Huang, Wu, Tashiro, Gao, Onodera and Ikejima112), flavones(Reference Lee, Kim, Kim, Kim, Noh, Kang, Cho, Choi and Suk113Reference Chen, Raung, Liao and Chen117), and flavonols(Reference Chen, Ho, Pei-Dawn, Chen, Jeng, Hsu, Lee, Wen and Lin118) are capable of inhibiting the release of NO by activated microglia via the down-regulation of iNOS gene expression. However, it is not known if such effects are mediated by changes in signalling through ERK, or any other MAPK, thus further investigation is warranted. The modulation of glial signalling cascades and pro-inflammatory transcription factors as well as cytokines and NO production may result in a suppression of neuroinflammation and ultimately to protection against neurodegeneration. It is plausible that the development of novel therapeutic agents or a cocktail of drugs that target neuroinflammation at various stages may act to reduce neurodegeneration and thus delay the progression of neurodegenerative disease.

Fig. 4 Structural homology of flavonoids with specific pathway inhibitors. Use of specific MAPK inhibitors such as SB203580 and PD98059 inhibit the transcriptional regulation of iNOS in activated glial cells. Interestingly, the structure of PD98059 and other kinase inhibitors have close structural homology to that of flavonoids. It is therefore possible that flavonoids may modulate neuroinflammation by interfering with cell signalling pathways such as MAPK.

Modulation of neuronal function through interaction with signalling pathways

As mentioned above, flavonoids have been shown to exert neuronal effects through their interactions with a number of protein kinase and lipid kinase signalling cascades, such as the PI3 kinase (PI3K)/Akt, tyrosine kinase, protein kinase C (PKC) and mitogen-activated protein kinase (MAP kinase) signalling pathways(Reference Spencer, Rice-Evans and Williams15, Reference Schroeter, Spencer, Rice-Evans and Williams72, Reference Matter, Brown and Vlahos119Reference Kong, Yu, Chen, Mandlekar and Primiano123) (Fig. 1). Inhibitory or stimulatory actions at these pathways are likely to profoundly affect neuronal function by altering the phosphorylation state of target molecules and/or by modulating gene expression. Although selective inhibitory actions at these kinase cascades may be beneficial in cancer, proliferative diseases, inflammation and neurodegeneration they could be detrimental during development particularly in the immature nervous system where protein and lipid kinase signalling regulates survival, synaptogenesis and neurite outgrowth. In the mature brain, post-mitotic neurones utilise MAP kinase and PI3K cascades in the regulation of key functions such as synaptic plasticity and memory formation(Reference Lin, Yeh, Lin, Lu, Leu, Chang and Gean124, Reference Sweatt125) (Fig. 1), thus flavonoid interactions within these pathways could have unpredictable outcomes and will be dependent both on the cell type and disease studied.

Flavonoids have the potential to bind to the ATP-binding sites of a large number of proteins(Reference Conseil, Baubichon-Cortay, Dayan, Jault, Barron and Di Pietro126) including, mitochondrial ATPase(Reference Di Pietro, Godinot, Bouillant and Gautheron127), calcium plasma membrane ATPase(Reference Barzilai and Rahamimoff128), protein kinase A(Reference Revuelta, Cantabrana and Hidalgo129), protein kinase C(Reference Gamet-Payrastre, Manenti, Gratacap, Tulliez, Chap and Payrastre122, Reference Lee and Lin130Reference Rosenblat, Belinky, Vaya, Levy, Hayek, Coleman, Merchav and Aviram133) and topoisomerase(Reference Boege, Straub, Kehr, Boesenberg, Christiansen, Andersen, Jakob and Kohrle134). In addition, interactions with the benzodiazepine binding sites of GABAA receptors and with adenosine receptors(Reference Medina, Viola, Wolfman, Marder, Wasowski, Calvo and Paladini135, Reference Dekermendjian, Kahnberg, Witt, Sterner, Nielsen and Liljefors136) have been reported. For example, the stilbene resveratrol and the citrus flavanones, hesperetin and naringenin, have been reported to have inhibitory activity at a number of protein kinases(Reference Fischer and Lane137Reference So, Guthrie, Chambers, Moussa and Carroll139). This inhibition is mediated via the binding of the polyphenols to the ATP binding site, presumably causing three-dimensional structural changes in the kinase leading to its inactivity. They may also interact directly with mitochondria, for example, by modulating the mitochondrial transition pore (mPT), which controls cytochrome c release during apoptosis(Reference Green and Reed140, Reference Tatton and Olanow141), or by modulating other mitochondrial associated pro-apoptotic factors such as DIABLO/smac(Reference Goyal142, Reference Srinivasula, Hegde and Saleh143). Potential interactions with the mPT are especially interesting, as the transition pore possesses a benzodiazepine-binding site where flavonoids may bind(Reference Medina, Viola, Wolfman, Marder, Wasowski, Calvo and Paladini135, Reference Dekermendjian, Kahnberg, Witt, Sterner, Nielsen and Liljefors136) and influence pore opening and cytochrome c release during apoptosis.

Flavonoids may also be capable of modulating glutamate excitotoxicity via direct scavenging of ROS or by the modulation of calcium influx. Abnormal influx of Ca2+ through AMPA-type glutamate receptors has been strongly implicated in neuronal death associated with a number of brain disorders through activation of Ca2+-dependent proteases, phospholipases and stress-activated kinases. Flavonoids may be capable of rendering heteromeric AMPA receptor assemblies Ca2+-impermeable by up-regulating GluR2 subunit expression(Reference Schroeter, Bahia, Spencer, Sheppard, Rattray, Rice-Evans and Williams73). Alternatively, flavonoids and their metabolites may prevent neuronal injury by scavenging of reactive intermediates such as superoxide and peroxynitrite derived from calcium mediated activation of xanthine oxidase and nitric oxide synthase, respectively. Lastly, modulation of signalling pathways and inhibition of calcium-activated kinases may also act to prevent excitotoxic death in neurons.

Interactions within the MAP kinase signalling cascade

Mitogen-activated protein kinases (MAPK) belong to the super-family of serine/threonine kinases and play a central role in transducing various extracellular signals into intracellular responses(Reference Cobb and Goldsmith144, Reference Goldsmith and Cobb145). MAPK cascades are organised into three main levels of regulation: (1) a MAP kinase kinase kinase (MAPKKK), which phosphorylates and activates; (2) a MAP kinase kinase (MAPKK), which in turn, phosphorylates and activates; (3) a MAP kinase (MAPK)(Reference Cobb and Goldsmith144, Reference Marshall146) (Fig. 5). The best characterised MAPK pathways are the mitogenic, extracellular signal-regulated protein kinase (ERK) pathway and the stress activated, c-Jun N-terminal kinase (JNK) (Fig. 6) and p38 cascades (Fig. 5). Once activated, ERK, JNK and p38 phosphorylate a number of cytosolic proteins and transcription factors resulting in the enhancement of their transcriptional activities and activation of dependent genes(Reference Karin147).

Fig. 5 Potential points of action of flavonoids within MAPK signalling cascades in neurons. Activation of ERK1/2 or ERK5 are generally pro-survival, whilst inhibitory actions on JNK and p38 pathways are also likely to be neuroprotective.

Fig. 6 Potential points of action of flavonoids within PI3K/Akt signalling pathway. Active PI3K catalyzes the production of phosphatidylinositol-3,4,5-triphosphate (PIP3) which activates phosphoinositide-dependent protein kinase 1 and 2 (PDK1 and PDK2) and Akt. Through its effects on these kinases, PI3K is involved in the regulation of a wide variety of processes, including cell growth, cell proliferation, differentiation, cell cycle entry, cell migration and apoptosis. Flavonoids have been proposed to act on this pathway via direct modulation of PI3K activity via binding to its ATP binding pocket, in a similar manner to that of LY294002. Alternatively, they may act to modulate the activity of the tumour suppressor, PTEN.

ERK and JNK are generally considered as having opposing actions, in particular in neuronal apoptosis(Reference Xia, Dickens, Raingeaud, Davis and Greenberg148). ERK1/2 are usually associated with pro-survival signalling(Reference Anderson and Tolkovsky149Reference Kaplan and Miller151) through mechanisms that may involve activation of the cAMP response element binding protein (CREB)(Reference Bonni, Brunet, West, Datta, Takasu and Greenberg150, Reference Crossthwaite, Hasan and Williams152) (Fig. 5), the up-regulation of the anti-apoptotic protein Bcl-2 and non-transcriptional inhibition of BAD(Reference Bonni, Brunet, West, Datta, Takasu and Greenberg150, Reference Kaplan and Miller151). On the other hand, JNK has been strongly linked to transcription-dependent apoptotic signalling(Reference Mielke and Herdegen153, Reference Yuan and Yankner154), through the activation of c-Jun(Reference Behrens, Sibilia and Wagner155) and other AP-1 proteins including JunB, JunD and ATF-2(Reference Davis156) (Fig. 5). Many investigations have indicated that flavonoids and their metabolites may interact selectively within the MAPK signalling pathways(Reference Kong, Yu, Chen, Mandlekar and Primiano123, Reference Kobuchi, Roy, Sen, Nguyen and Packer157). This modulation of MAPK signalling by flavonoids is significant as ERK1/2 and JNK are involved in growth factor induced mitogenesis, differentiation, apoptosis and various forms of cellular plasticity(Reference Mielke and Herdegen153, Reference Yuan and Yankner154, Reference Herdegen, Skene and Bahr158Reference Castagne, Gautschi, Lefevre, Posada and Clarke160).

Extracellular signal-regulated protein kinase (ERK) pathway

Although most investigations have centred on the potential of flavonoids to modulate the phosphorylation state of ERK1/2(Reference Spencer, Rice-Evans and Williams15, Reference Schroeter, Boyd, Spencer, Williams, Cadenas and Rice-Evans16, Reference Schroeter, Spencer, Rice-Evans and Williams72, Reference Hung, Hsu, Lo, Huang, Wang and Chen161, Reference Llorens, Garcia, Itarte and Gomez162), it is likely that their actions on this MAPK isoform result from effects on upstream kinases, such as MEK1 and MEK2 (Fig. 5), and potentially membrane receptors(Reference Schroeter, Boyd, Spencer, Williams, Cadenas and Rice-Evans16). This appears likely as flavonoids have close structural homology to specific inhibitors of ERK signalling, such as PD98059 (2′-amino-3′-methoxyflavone) (Fig. 4). PD98059 is a flavone that has been shown to act in vivo as a highly selective non-competitive inhibitor of MEK1 activation and the MAP kinase cascade(Reference Dudley, Pang, Decker, Bridges and Saltiel163Reference Lazar, Wiese, Brady, Mastick, Waters, Yamauchi, Pessin, Cuatrecasas and Saltiel166). PD98059 acts via its ability to bind to the inactive forms of MEK so preventing its activation by upstream activators such as c-Raf(Reference Alessi, Cuenda, Cohen, Dudley and Saltiel165). This raises the possibility that flavonoids, and their metabolites, may also act on this pathway in a similar manner. In support of this, the flavonol quercetin, and to a lesser extent its O-methylated metabolites have been shown to induce neuronal apoptosis via a mechanism involving the inhibition of ERK, rather than by induction of pro-apoptotic signalling through JNK(Reference Spencer, Rice-Evans and Williams15). The potent inhibition of ERK activation, and indeed Akt/PKB phosphorylation, was also accompanied by downstream activation of BAD and a subsequent strong activation of caspase-3.

On the other hand, some flavonoids have been observed to exert a stimulatory effect on ERK1/2. For example, the flavan-3-ol, ( − )-epicatechin, and one of its metabolites, 3′-O-methyl-( − )-epicatechin, have been shown to stimulate phosphorylation of ERK1/2 and the downstream transcription factor CREB at physiologically relevant concentrations(Reference Schroeter, Bahia, Spencer, Sheppard, Rattray, Rice-Evans and Williams73). Interestingly, this activation of the ERK pathway was no longer apparent at higher concentrations suggesting that effects on this pathway are concentration specific. Furthermore, stimulation of the ERK1/2 and CREB was not observed with ( − )-epicatechin-5-O-β-d-glucuronide suggesting that effects on the ERK pathway may be dependent on cell or membrane permeability, as has been previously reported(Reference Spencer, Schroeter, Crossthwaithe, Kuhnle, Williams and Rice-Evans28). In support of these observations, the protective action of another flavanol, EGCG, against 6-hydroxy dopamine toxicity and serum deprivation has been shown to involve the restoration of both protein kinase C and ERK1/2 activities(Reference Levites, Amit, Youdim and Mandel167, Reference Reznichenko, Amit, Youdim and Mandel168).

One explanation for the concentration-specific regulation of the ERK pathway, and indeed other MAP kinase cascades (JNK and p38), may be related to the ability of flavonoids to exert high affinity receptor agonist-like actions at low concentrations and direct enzyme inhibition at higher concentrations(Reference Agullo, Gamet-Payrastre, Manenti, Viala, Remesy, Chap and Payrastre121, Reference Walker, Pacold, Perisic, Stephens, Hawkins, Wymann and Williams169), or by inducing receptor desensitization. The identity of the primary flavonoid interacting sites in neurons is unknown and could be either at the cell surface or intracellular, although the ERK and PI3K dependence to CREB phosphorylation is reminiscent of ionotropic receptor signalling(Reference Perkinton, Sihra and Williams170). Receptors reported to act as flavonoid-binding sites, that are present in cortical neurons, are adenosine(Reference Jacobson, Moro, Manthey, West and Ji171) and GABAA receptors(Reference Johnston172, Reference Adachi, Tomonaga, Tachibana, Denbow and Furuse173). However, a specific plasma membrane binding site for polyphenols has recently been described in rat brain(Reference Han, Bastianetto, Dumont and Quirion174). In addition, monomeric and dimeric flavanols show nanomolar affinity and efficacy at testosterone receptors(Reference Nifli, Bosson-Kouame, Papadopoulou, Kogia, Kampa, Castagnino, Stournaras, Vercauteren and Castanas175) and resveratrol rapidly activates ERK signalling through alpha and beta oestrogen receptors(Reference Klinge, Blankenship, Risinger, Bhatnagar, Noisin, Sumanasekera, Zhao, Brey and Keynton176). Collectively, this raises the possibility that flavonoids may act on the ERK pathway via acting through steroid-like receptors in neurons to modulate ERK and CREB-mediated gene expression.

In addition to a receptor-mediated mechanism, it is equally plausibly that changes in ERK activation and related transcription factors (i.e. CREB) may result from flavonoid-induced modulation of phosphatase activity. Phosphatases act in opposition to kinases by de-phosphorylating specific kinases and in the process either activate or de-activate them. Consequently, phosphatases are integral to many signalling pathways. Because ERK and other MAPK require both Thr and Tyr phosphorylation for full activity, dual specificity phosphatases (DSPs) that de-phosphorylate both sites are uniquely positioned to regulate MAPK signal transduction cascades. At least nine DSPs, also termed MAPK phosphatases (MKPs), have been identified in mammalian cells(Reference Camps, Nichols and Arkinstall177). DSPs frequently associated with ERK inactivation include MKP3, MKP4, and phosphatase of activated cells 1 (PAC1), although MKP3 (also termed PYST1) is probably the best studied and the most specific for ERK1/2 versus other MAPK(Reference Kim, Rice and Denu178). The finding that multiple phosphatases inactivate the ERK pathway suggests that the duration and extent of ERK activation is controlled by a balance of the activities of upstream MAPKK, such as MEK1, and phosphatases, such MKP3. Although there has been intense interest in the ability of flavonoids to modulate kinases, thus far there is no indication that they may affect signalling pathways via a modulation of phosphatase activity. If flavonoids are capable of interacting with phosphatases, such as MKP3, then this is likely to have a dramatic effect on the activation states of important kinases like ERK1/2. Future investigations in this area should consider the potential of flavonoids to inhibit, or activate phosphatases, the concentration-dependency of these effects and the mechanism by which they do so.

Stress-activated protein kinases: c-Jun-N-terminal kinase (JNK) and p38

There is strong evidence linking the activation of JNK to neuronal loss in response to a wide array of pro-apoptotic stimuli in both developmental and degenerative death signalling(Reference Mielke and Herdegen153, Reference Davis156, Reference Davis179). The activation of the JNK pathway and the death of specific neuronal populations is crucial during early brain development(Reference Leppa and Bohmann180). As with the other MAP kinases, the core signalling unit is composed of a MAPKKK, typically MEKK1-4, which phosphorylate and activate MKK4-7, which then phosphorylate and activate the JNK(Reference Davis179, Reference Ichijo181) (Fig. 5). Another MAPKKK, apoptosis signal-regulating kinase 1 (ASK1), also plays an essential role in stress-induced apoptosis(Reference Ichijo, Nishida, Irie, ten Dijke, Saitoh, Moriguchi, Takagi, Matsumoto, Miyazono and Gotoh182, Reference Wang, Diener, Jannuzzi, Trollinger, Tan, Lichenstein, Zukowski and Yao183). ASK1 can be activated in response to a variety of stress-related stimuli, including oxidative stress and activates MKK4, which in turn activates JNK (Fig. 5) and indeed p38(Reference Matsuzawa and Ichijo184). Overexpression of ASK1 has been shown to induce the activation of both JNK and p38 and lead to apoptosis via signals involving the mitochondrial cell death pathway(Reference Leppa and Bohmann180, Reference Ichijo, Nishida, Irie, ten Dijke, Saitoh, Moriguchi, Takagi, Matsumoto, Miyazono and Gotoh182).

Investigation has indicated that oxidative-induced activation of caspase-3 in neurons is blocked by flavonoids, providing compelling evidence in support of a potent anti-apoptotic action of flavonoids in these cells(Reference Spencer, Schroeter, Crossthwaithe, Kuhnle, Williams and Rice-Evans28, Reference Schroeter, Spencer, Rice-Evans and Williams72, Reference Schroeter, Williams, Matin, Iversen and Rice-Evans185). The flavanols, epicatechin and 3′-O-methyl-epicatechin have been shown to protect neurons against oxidative damage via a mechanism involving the suppression of JNK, and downstream partners, c-Jun and pro-caspase-3(Reference Schroeter, Spencer, Rice-Evans and Williams72, Reference Spencer, Schroeter, Kuhnle, Srai, Tyrrell, Hahn and Rice-Evans186). Similarly, the flavone, baicalein, has been shown to significantly inhibit 6-hydroxydopamine-induced JNK activation and neuronal cell death and quercetin may suppress JNK activity and apoptosis induced by hydrogen peroxide(Reference Wang, Matsushita, Araki and Takeda187, Reference Ishikawa and Kitamura188), 4-hydroxy-2-nonenal(Reference Uchida, Shiraishi, Naito, Torii, Nakamura and Osawa189) and tumour necrosis factor-alpha (TNF-alpha)(Reference Kobuchi, Roy, Sen, Nguyen and Packer157). There are a number of potential sites where flavonoids may interact with the JNK pathway. For instance, flavonoid-mediated inhibition of oxidative stress-induced apoptosis may occur by preventing the activation of JNK by influencing one of the many upstream MAPKKK activating proteins that transduce signals to JNK (Fig. 5). For example, their ability to inhibit JNK activation may proceed via flavonoid-induced modulation of the ASK1 phosphorylation state, and its association with 14-3-3 protein, which is essential for suppression of cellular apoptosis(Reference Zhang, Chen and Fu190). Other potential mechanisms include an ability to preserve Ca2+ homeostasis, thereby preventing Ca2+-dependent activation of JNK(Reference Schroeter, Spencer, Rice-Evans and Williams72, Reference Davis179) or an attenuation of the pro-apoptotic signalling cascade lying downstream of JNK.

Another potential site of action may be specific redox-sensitive motifs, notably cysteine residues, similar to those reported for JNK(Reference Park, Park, Kim, Ahn, Kim and Choi191). JNK redox regulation has been proposed to proceed through its binding to redox sensitive proteins such as glutathione-S-transferase (GST)(Reference Adler, Yin and Fuchs192Reference Monaco, Friedman, Hyde, Chen, Manolatus, Adler, Ronai, Koslosky and Pincus194). It has been shown that under unstressed conditions, JNK is associated with GST resulting in the inhibition of JNK activity, but that JNK dissociates from GST following UV or oxidative stress(Reference Adler, Yin and Fuchs192, Reference Yin, Ivanov, Habelhah, Tew and Ronai195). Flavonoids also may act to inhibit JNK activity, and possibly other MAPKs, via the nucleophilic addition of flavonoid-o-quinones, formed during the intracellular oxidation of flavonoids(Reference Spencer, Kuhnle, Williams and Rice-Evans27, Reference Spencer, Abd, Mohsen and Rice-Evans196), to cysteine residues on JNK.

PI3 kinase signalling pathway

In addition to MAPK pathway, flavonoids have been shown to modulate signalling through the serine/threonine kinase, Akt/PKB, one of the main downstream effectors of PI3K, a pivotal kinase in neuronal survival(Reference Kennedy, Wagner, Conzen, Jordan, Bellacosa, Tsichlis and Hay197Reference Crowder and Freeman200) (Fig. 6). Active PI3K catalyzes the production of phosphatidylinositol-3,4,5-triphosphate (PIP3) by phosphorylating phosphatidylinositol (PI), phosphatidylinositol-4-phosphate (PIP) and phosphatidylinositol-4,5-bisphosphate (PIP2). PIP3 may then activate phosphoinositide-dependent protein kinase 1 (PDK1), which plays a central role in many signal transduction pathways(Reference Carpenter and Cantley201, Reference Simpson and Parsons202), activating Akt and the PKC isoenzymes p70 S6 kinase and RSK(Reference Neri, Borgatti, Capitani and Martelli203). Through its effects on these kinases, PDK1 is involved in the regulation of a wide variety of processes, including cell growth, cell proliferation, differentiation, cell cycle entry, cell migration and apoptosis(Reference Carpenter and Cantley201). One of the most important targets of PI3K and PDK1 is Akt (also known as Protein Kinase B), as this kinase plays a critical role in controlling cellular survival and apoptosis(Reference Franke, Kaplan and Cantley204Reference Franke, Yang, Chan, Datta, Kazlauskas, Morrison, Kaplan and Tsichlis206) (Fig. 6). Akt promotes cell survival by inhibiting apoptosis through its ability to phosphorylate and inactivate several important targets, including Bad(Reference Cardone, Roy, Stennicke, Salvesen, Franke, Stanbridge, Frisch and Reed207), Forkhead transcription factors(Reference Burgering and Kops208, Reference Brunet, Bonni, Zigmond, Lin, Juo, Hu, Anderson, Arden, Blenis and Greenberg209) and caspase-9(Reference Burgering and Coffer205). Indeed, activation of Akt/PKB in neurons has been shown to lead to an inhibition of proteins central to neuronal death machinery, such as the pro-apoptotic Bcl-2 family member, Bad(Reference Zha, Harada, Yang, Jockel and Korsmeyer210), and members of the caspase family(Reference Bonni, Brunet, West, Datta, Takasu and Greenberg150, Reference Kennedy, Wagner, Conzen, Jordan, Bellacosa, Tsichlis and Hay197) that specifically cleave poly(ADP-ribose) polymerase(Reference Kennedy, Wagner, Conzen, Jordan, Bellacosa, Tsichlis and Hay197, Reference Coffer, Jin and Woodgett198), thus promoting cell survival. Akt is activated by phospholipid binding and activation loop phosphorylation at Thr308 by PDK1(Reference Alessi, Andjelkovic, Caudwell, Cron, Morrice, Cohen and Hemmings211) and by phosphorylation within the carboxy terminus at Ser473 and the activation of Akt is a pro-survival event in many cell types due to its ability to inactivate Bad via phosphorylation at Ser136(Reference Datta, Dudek, Tao, Masters, Fu, Gotoh and Greenberg212, Reference del Peso, Gonzalez-Garcia, Page, Herrera and Nunez213).

There is good evidence that flavonoids inhibit PI3K via direct interactions with its ATP binding site. Indeed, a number of studies have demonstrated that the structure of flavonoids determines whether or not they act as potent inhibitors of PI3K(Reference Agullo, Gamet-Payrastre, Manenti, Viala, Remesy, Chap and Payrastre121, Reference Ferriola, Cody and Middleton214). One of the most selective PI3K inhibitors available, LY294002 (Fig. 4), was modelled on the structure of quercetin(Reference Matter, Brown and Vlahos119, Reference Vlahos, Matter, Hui and Brown120). LY294002 and quercetin fit into the ATP binding pocket of the enzyme although with surprisingly different orientations(Reference Walker, Pacold, Perisic, Stephens, Hawkins, Wymann and Williams169). It appears that the number and substitution of hydroxyl groups on the B-ring and the degree of un-saturation of the C2–C3 bond are important determinants of this particular bioactivity. Interestingly in this regard quercetin and some of its in vivo metabolites inhibit pro-survival Akt/PKB signalling pathways(Reference Spencer, Rice-Evans and Williams15) by a mechanism of action consistent with quercetin and its metabolites acting at and inhibiting PI3K activity. Prior to inducing measurable losses of neuronal viability, quercetin stimulates a strong inhibition of basal Akt phosphorylation at both the regulatory serine473 and catalytic threonine308 sites, rendering it inactive. The inhibition of Akt/PKB phosphorylation in this way may reflect potential inhibition of its upstream partner PI3K, as has previously been described(Reference Matter, Brown and Vlahos119). If Akt/PKB inhibition is sustained, which has been reported to occur during neuronal exposure to quercetin, this leads to extensive caspase-3 activation and subsequent caspase-dependent cleavage of Akt/PKB, an event that effectively ‘switches off’ a major survival signal and results in the acceleration of apoptotic death(Reference Spencer, Rice-Evans and Williams15). However, at lower concentrations, quercetin has also been shown to trigger CREB activation in neurons indicating that exposure concentration is pivotal in determining either pro-apoptotic or anti-apoptotic effects(Reference Spencer, Rice-Evans and Williams15). Indeed, low concentrations of quercetin, may activate the MAPK pathway (ERK2, JNK1 and p38) leading to expression of survival genes (c-Fos, c-Jun) and defensive genes (Phase II detoxifying enzymes; glutathione-S-transferase, quinone reductase) resulting in survival and protective mechanisms (homeostasis response), whereas high concentrations stimulate pro-apoptotic pathways and caspase activation(Reference Kong, Yu, Chen, Mandlekar and Primiano123).

Another potential mechanism by which flavonoids may modulate the PI3 kinase/Akt signalling pathway is by their ability to modulate the expression or activity of PTEN (phosphatase and tensin homologue deleted on chromosome ten), also referred to as MMAC (mutated in multiple advanced cancers) phosphatase(Reference Cao, Jin and Zhou215Reference Dave, Eason, Till, Geng, Velarde, Badger and Simmen217). PTEN is a tumour suppressor implicated in a wide variety of human cancers(Reference Cantley and Neel218) and the main substrates of PTEN are inositol phospholipids generated by the activation of the PI3K(Reference Myers, Pass, Batty, Van der KJ, Stolarov, Hemmings, Wigler, Downes and Tonks219). PTEN acts a major negative regulator of the PI3K/Akt signalling pathway(Reference Cantley and Neel218, Reference Wu, Senechal, Neshat, Whang and Sawyers220) and thus a modulation of its expression or activation by flavonoids will have a profound effect of cellular function (Fig. 6). For example, if flavonoids are capable of inhibiting PTEN in cancer cells this may lead to an increase in cancer cell proliferation and tumour growth. On the other hand, its activation in post-mitotic cells, such as neurons, may have a positive effect by increasing Akt and CREB activity leading to a promotion of neuronal survival and synaptic plasticity (Fig. 2).

The Nrf-Keap1/ARE pathway and interactions with MAP kinase and PI3 kinase

The regulation of γ-GCS and other detoxification proteins, such as glutathione peroxidase, has been shown to involve the transcription factor NF-E2-related factor 2, Nrf2(Reference Wild, Moinova and Mulcahy221, Reference Moinova and Mulcahy222) and Nrf1(Reference Venugopal and Jaiswal223, Reference Venugopal and Jaiswal224). Nrf2 is known to regulate the gene expression of phase II detoxification enzymes and antioxidant proteins through an enhancer sequence referred to as the ‘electrophile responsive element’ or ‘antioxidant responsive element’ (EpRE/ARE)(Reference Kang, Lee and Kim225, Reference Itoh, Tong and Yamamoto226). Deficiencies in Nrf1 and Nrf2 have been found to largely abolish the constitutive and/or inducible expression of defence enzyme genes in response to oxidative and xenobiotic stress(Reference Leung, Kwong, Hou, Lee and Chan227, Reference Bloom, Dhakshinamoorthy and Jaiswal228). Recent reports have suggested that signalling through Nrf2 is involved in HO-1 induction by polyphenols such as epigallocatechin-3-gallate(Reference Andreadi, Howells, Atherfold and Manson229), whilst sulforaphane and curcumin have been shown to exert anti-inflammatory and anti-carcinogenic effects through activation of Nrf2 and subsequent up-regulation of gastrointestinal GPx(Reference Banning, Deubel, Kluth, Zhou and Brigelius-Flohe230). The protective effects of the isoflavone, genistein has also been shown to depend primarily on the activation of glutathione peroxidase mediated by Nrf1 activation(Reference Hernandez-Montes, Pollard, Vauzour, Jofre-Montseny, Rota, Rimbach, Weinberg and Spencer231). However, there is limited information regarding the effects of flavonoids on Nrf1 and Nrf2 activation or whether their ability to modulate this important signalling pathway mediates their cellular beneficial effects.

The actin-binding protein, Keap1, has been identified as a docking site for Nrf2 that is responsible for sequestering Nrf2 in the cytoplasmic compartment of unstressed cells(Reference Itoh, Wakabayashi, Katoh, Ishii, Igarashi, Engel and Yamamoto232). The association between the two proteins is between the C-terminal DGR domain of Keap1 and the N-terminal Neh2 domain of Nrf2. Although the exact mechanism of cytosolic sequestration of Nrf1 is unknown, it is also known to contain a Neh2 domain, and therefore it too may interact with Keap1 in vivo (Reference Hayes and McMahon233). One way in which flavnoids may induce Nrf1 and Nrf2 are via the rapid but non-toxic increases in intracellular oxidative species as has been suggested for 3′,4′,5′,3,4,5-hexamethoxy-chalcone(Reference Alcaraz, Vicente, Araico, Dominguez, Terencio and Ferrandiz234). Increases in reactive oxygen species may induce the disruption of the Keap1/Nrf2 complex through oxidation of Cys273 and Cys288 residues on Keap1. Cellular uptake and metabolism of flavonoids has been shown to lead to the formation of intracellular flavonoid-o-quinone species(Reference Spencer, Kuhnle, Williams and Rice-Evans27, Reference Spencer, Abd, Mohsen and Rice-Evans196, Reference Hernandez-Montes, Pollard, Vauzour, Jofre-Montseny, Rota, Rimbach, Weinberg and Spencer231), which increase intracellular oxidative stress, or alternatively may react directly with Cys273 and Cys288 and/or other cysteine residues located on Keap1(Reference Itoh, Tong and Yamamoto226, Reference Kobayashi and Yamamoto235). Interactions with cysteinyl residues would be expected to trigger the release Nrf1/Nrf2 from Keap1, so allowing their phosphorylation and translocation to the nucleus, where they may interact with the electrophile response element. Future studies are required to determine the precise nature of interactions between intracellular flavonoid metabolites, such as o-quinones, and the Keap1/Nrf1 complex.

Besides the direct oxidation or covalent modification of thiol groups of Keap1, the Nrf2-Keap1-ARE signalling can be modulated by post-transcriptional modification of Nrf2. Three major signal transduction pathways have been implicated in the regulation of ARE/EpRE motifs and nuclear translocation of Nrf2. These are the MAPK cascade, protein kinase C (PKC) and PI3 kinase(Reference Nguyen, Sherratt, Huang, Yang and Pickett236Reference Sherratt, Huang, Nguyen and Pickett238) (Fig. 7). For example, the activation of ERK and JNK has been shown to induce ARE-mediated gene expression via the recruitment of a co-activator to the transcription initiation complex and an increase in Nrf2 transcriptional activity, whereas, p38 had the opposite effect(Reference Yu, Mandlekar, Lei, Fahl, Tan and Kong239, Reference Shen, Hebbar, Nair, Xu, Li, Lin, Keum, Han, Gallo and Kong240). Furthermore, PI3 kinase also appears to be involved in Nrf2 activation in cells exposed to peroxynitrite(Reference Kang, Lee, Park and Kim241) and hemin(Reference Nakaso, Yano, Fukuhara, Takeshima, Wada-Isoe and Nakashima242). In response to oxidative stress, the activation of signalling cascades mediated by PI3K results in de-polymerization of actin microfilaments thereby facilitating Nrf2 translocation to the nucleus(Reference Kang, Lee, Park and Kim241). As flavonoids are known to modulate MAP kinase and PI3 kinase/Akt pathways (outlined above), it is highly likely that they will also affect the activation and nuclear translocation of Nrf1 and/or Nrf2 indirectly through the modulation of these upstream pathways (Fig. 7).

Fig. 7 Involvement of MAP kinase and PI3 kinase signalling in regulation of the Keap1-Nrf2 pathway. Inactive Nrf2 is retained in the cytosol by association a complex with the cytoskeletal protein Keap1. Cytosolic Nrf2 may be phosphorylated in response to MAP kinase, PI3 kinase and protein kinase C pathways. Following phosphorylation, Nrf2 translocates to the nucleus, where it activates gene expression through binding to the ARE, following its interaction with other transcription factors in the bZIP family (CREB, ATF-4 and fos or c-Jun). Nrf2 activation of genes is opposed by small maf proteins, including MafG and MafK, maintaining a counterbalance to Nrf2 and balancing the oxidation level of the intracellular environment.

Conclusions

Evidence suggests that dietary phytochemicals, in particular flavonoids, may exert beneficial effects in the CNS by protecting neurons against stress induced injury, by suppressing the activation of microglia and astrocytes, which mediate neuroinflammation, and by promoting synaptic plasticity, memory and cognitive function. Evidence also supports the localization of flavonoids within the brain, thus these phytochemicals may be regarded as a potential neuroprotective, neuromodulatory or anti-neuroinflammatory agents. It appears highly likely that such beneficial properties are mediated by their abilities to interact with both protein and lipid kinase signalling cascades, rather then via their potential to act as classical antioxidants. The concentrations of flavonoids encountered in vivo are sufficiently high to exert pharmacological activity at receptors, kinases and transcription factors. Presently the precise sites of action are unknown, although it is likely that their activity depends on their ability to (1) bind to ATP sites on enzymes and receptors; (2) modulate the activity of kinases directly, i.e. MAPKKK, MAPKK or MAPK; (3) affect the function of important phosphatases, which act in opposition to kinases; (4) preserve Ca2+ homeostasis, thereby preventing Ca2+-dependent activation of kinases in neurons; and (5) modulate signalling cascades lying downstream of kinases, i.e. transcription factor activation and binding to promoter sequences.

However, at present, more information is required in order to understand the precise cellular site(s) of action of flavonoids. For instance, it is still unclear whether flavonoid action requires cellular uptake or if they are capable of mediating effects via extracellular receptor binding. Presently, there is no certainty either way, although flavonoid glucuronides, which are unable to enter cells to any significant degree, do not appear to express cellular effects. This may suggest a requirement for cytosolic localisation, although it could equally signify that the conjugation of flavonoids with glucuronide or sulphate moieties blocks receptor binding and therefore their cellular activity. It seems likely that the inhibition of Akt is almost certainly mediated via actions at PI3K, thus requiring cellular uptake. However, actions at ERK1/2 could result from either flavonoid modulation of upstream regulatory kinases or by binding directly to receptors. The challenge now is to determine the precise site(s) of action of flavonoids within the signalling pathways and the sequence of events that allow them to regulate neuronal function in the central nervous system.

Ultimately actions within these neuronal signalling cascades may be beneficial or negative in the context of the brain. For example, whilst they may be positive in the treatment of proliferative diseases, they could be detrimental to the nervous system, at least at high concentrations, where these same pathways act to control neuronal survival and synaptic plasticity. Thus, flavonoid interactions with intracellular signalling pathways could have unpredictable outcomes and will be dependent on the cell type (i.e. neurons, astrocytes, microglia, oligodendrocytes), the disease studied and the stimulus applied. In summary, it is evident that flavonoids are potent bioactive molecules and a clear understanding of their mechanisms of action as modulators of cell signalling will be crucial in the evaluation of their potential to act as inhibitors of neurodegeneration or as modulators of brain function.

Acknowledgements

The author is sponsored by the Biotechnology and Biological Sciences Research Council (BB/C518222/1 and BB/F008953/1) and the Medical Research Council (G0400278/NI02). The publication of this paper was made possible by the financial support of the European Co-operation in the field of Scientific and Technical (COST) Research Action 926 ‘Impact of new technologies on the health benefits and safety of bioactive plant compounds’ (2004–2008). The author had no conflicts of interest to disclose.

References

1Youdim, KA & Joseph, JA (2001) A possible emerging role of phytochemicals in improving age-related neurological dysfunctions: a multiplicity of effects. Free Radic Biol Med 30, 583594.CrossRefGoogle ScholarPubMed
2Youdim, KA, Spencer, JP, Schroeter, H & Rice-Evans, C (2002) Dietary flavonoids as potential neuroprotectants. Biol Chem 383, 503519.CrossRefGoogle ScholarPubMed
3Galli, RL, Shukitt-Hale, B, Youdim, KA & Joseph, JA (2002) Fruit polyphenolics and brain aging: nutritional interventions targeting age-related neuronal and behavioral deficits. Ann N Y Acad Sci 959, 128132.CrossRefGoogle ScholarPubMed
4Unno, K, Takabayashi, F, Kishido, T & Oku, N (2004) Suppressive effect of green tea catechins on morphologic and functional regression of the brain in aged mice with accelerated senescence (SAMP10). Exp Gerontol 39, 10271034.CrossRefGoogle ScholarPubMed
5Haque, AM, Hashimoto, M, Katakura, M, Tanabe, Y, Hara, Y & Shido, O (2006) Long-term administration of green tea catechins improves spatial cognition learning ability in rats. J Nutr 136, 10431047.CrossRefGoogle ScholarPubMed
6Kuriyama, S, Hozawa, A, Ohmori, K, Shimazu, T, Matsui, T, Ebihara, S, Awata, S, Nagatomi, R, Arai, H & Tsuji, I (2006) Green tea consumption and cognitive function: a cross-sectional study from the Tsurugaya Project 1. Am J Clin Nutr 83, 355361.CrossRefGoogle ScholarPubMed
7Wang, Y, Wang, L, Wu, J & Cai, J (2006) The in vivo synaptic plasticity mechanism of EGb 761-induced enhancement of spatial learning and memory in aged rats. Br J Pharmacol 148, 147153.CrossRefGoogle ScholarPubMed
8Inanami, O, Watanabe, Y, Syuto, B, Nakano, M, Tsuji, M & Kuwabara, M (1998) Oral administration of ( − )catechin protects against ischemia-reperfusion-induced neuronal death in the gerbil. Free Radic Res 29, 359365.CrossRefGoogle ScholarPubMed
9Luo, Y, Smith, JV, Paramasivam, V, Burdick, A, Curry, KJ, Buford, JP, Khan, I, Netzer, WJ, Xu, H & Butko, P (2002) Inhibition of amyloid-beta aggregation and caspase-3 activation by the Ginkgo biloba extract EGb761. Proc Natl Acad Sci U S A 99, 1219712202.CrossRefGoogle ScholarPubMed
10Bastianetto, S, Zheng, WH & Quirion, R (2000) The Ginkgo biloba extract (EGb 761) protects and rescues hippocampal cells against nitric oxide-induced toxicity: involvement of its flavonoid constituents and protein kinase C. J Neurochem 74, 22682277.CrossRefGoogle ScholarPubMed
11Zimmermann, M, Colciaghi, F, Cattabeni, F & Di Luca, M (2002) Ginkgo biloba extract: from molecular mechanisms to the treatment of Alzhelmer's disease. Cell Mol Biol (Noisy-le-grand) 48, 613623.Google Scholar
12Datla, KP, Christidou, M, Widmer, WW, Rooprai, HK & Dexter, DT (2001) Tissue distribution and neuroprotective effects of citrus flavonoid tangeretin in a rat model of Parkinson's disease. Neuroreport 12, 38713875.CrossRefGoogle Scholar
13Rice-Evans, CA, Miller, NJ & Paganga, G (1996) Structure–antioxidant activity relationships of flavonoids and phenolic acids. Free Radic Biol Med 20, 933956.CrossRefGoogle ScholarPubMed
14Williams, RJ, Spencer, JPE & Rice-Evans, C (2004) Flavonoids: antioxidants or signalling molecules? Free Radic Biol Med 36, 838849.CrossRefGoogle ScholarPubMed
15Spencer, JPE, Rice-Evans, C & Williams, RJ (2003) Modulation of pro-survival Akt/PKB and ERK1/2 signalling cascades by quercetin and its in vivo metabolites underlie their action on neuronal viability. J Biol Chem 278, 3478334793.CrossRefGoogle ScholarPubMed
16Schroeter, H, Boyd, C, Spencer, JPE, Williams, RJ, Cadenas, E & Rice-Evans, C (2002) MAPK signaling in neurodegeneration: influences of flavonoids and of nitric oxide. Neurobiol Aging 23, 861880.CrossRefGoogle ScholarPubMed
17Manach, C, Scalbert, A, Morand, C, Remesy, C & Jimenez, L (2004) Polyphenols: food sources and bioavailability. Am J Clin Nutr 79, 727747.CrossRefGoogle ScholarPubMed
18Rice-Evans, C (2001) Flavonoid antioxidants. Curr Med Chem 8, 797807.CrossRefGoogle ScholarPubMed
19Rice-Evans, C (1995) Plant polyphenols: free radical scavengers or chain-breaking antioxidants? Biochem Soc Symp 61, 103116.CrossRefGoogle ScholarPubMed
20Spencer, JPE, Schroeter, H, Rechner, AR & Rice-Evans, C (2001) Bioavailability of flavan-3-ols and procyanidins: gastrointestinal tract influences and their relevance to bioactive forms in vivo. Antioxid Redox Signal 3, 10231039.CrossRefGoogle ScholarPubMed
21Spencer, JPE, Srai, SK & Rice-Evans, C (2003) Metabolism in the small intestine and gastrointestinal tract. In Flavonoids in Health and Disease, pp. 363390 [Rice-Evans, C and Packer, L, editors]. New York: Marcel Dekker.Google Scholar
22Walle, T, Walgren, RA, Walle, UK, Galijatovic, A & Vaidyanathan, JB (2003) Understanding the bioavailability of flavanoids through studies in Caco-2 cells. In Flavonoids in Health and Disease, pp. 349362 [Rice-Evans, C and Packer, L, editors]. New York: Marcel Dekker.Google Scholar
23Day, AJ & Williamson, G (2003) Absorption of quercetin glycosides. In Flavonoids in Health and Disease, pp. 391412 [Rice-Evans, C and Packer, L, editors]. New York: Marcel Dekker.Google Scholar
24Donovan, JL & Waterhouse, AL (2003) Bioavailability of flavanol monomers. In Flavonoids in Health and Disease, pp. 413440 [Rice-Evans, C and Packer, L, editors]. New York: Marcel Dekker.Google Scholar
25Spencer, JPE, Chowrimootoo, G, Choudhury, R, Debnam, ES, Srai, SK & Rice-Evans, C (1999) The small intestine can both absorb and glucuronidate luminal flavonoids. FEBS Lett 458, 224230.CrossRefGoogle ScholarPubMed
26Scheline, RR (1999) Metabolism of oxygen heterocyclic compounds. In CRC Handbook of Mammalian Metabolism of Plant Compounds, pp. 243295Boca Ranton: CRC Press.Google Scholar
27Spencer, JPE, Kuhnle, GG, Williams, RJ & Rice-Evans, C (2003) Intracellular metabolism and bioactivity of quercetin and its in vivo metabolites. Biochem J 372, 173181.CrossRefGoogle ScholarPubMed
28Spencer, JPE, Schroeter, H, Crossthwaithe, AJ, Kuhnle, G, Williams, RJ & Rice-Evans, C (2001) Contrasting influences of glucuronidation and O-methylation of epicatechin on hydrogen peroxide-induced cell death in neurons and fibroblasts. Free Radic Biol Med 31, 11391146.CrossRefGoogle ScholarPubMed
29Miyake, Y, Shimoi, K, Kumazawa, S, Yamamoto, K, Kinae, N & Osawa, T (2000) Identification and antioxidant activity of flavonoid metabolites in plasma and urine of eriocitrin-treated rats. J Agric Food Chem 48, 32173224.CrossRefGoogle ScholarPubMed
30Terao, J, Yamaguchi, S, Shirai, M, Miyoshi, M, Moon, JH, Oshima, S, Inakuma, T, Tsushida, T & Kato, Y (2001) Protection by quercetin and quercetin 3-O-beta-d-glucuronide of peroxynitrite-induced antioxidant consumption in human plasma low-density lipoprotein. Free Radic Res 35, 925931.CrossRefGoogle ScholarPubMed
31Shirai, M, Moon, JH, Tsushida, T & Terao, J (2001) Inhibitory effect of a quercetin metabolite, quercetin 3-O-beta-d-glucuronide, on lipid peroxidation in liposomal membranes. J Agric Food Chem 49, 56025608.CrossRefGoogle ScholarPubMed
32Yamamoto, N, Moon, JH, Tsushida, T, Nagao, A & Terao, J (1999) Inhibitory effect of quercetin metabolites and their related derivatives on copper ion-induced lipid peroxidation in human low-density lipoprotein. Arch Biochem Biophys 372, 347354.CrossRefGoogle ScholarPubMed
33da Silva, EL, Piskula, MK, Yamamoto, N, Moon, JH & Terao, J (1998) Quercetin metabolites inhibit copper ion-induced lipid peroxidation in rat plasma. FEBS Lett 430, 405408.CrossRefGoogle ScholarPubMed
34Abbott, NJ (2002) Astrocyte–endothelial interactions and blood–brain barrier permeability. J Anat 200, 629638.CrossRefGoogle ScholarPubMed
35Youdim, KA, Dobbie, MS, Kuhnle, G, Proteggente, AR, Abbott, NJ & Rice-Evans, C (2003) Interaction between flavonoids and the blood–brain barrier: in vitro studies. J Neurochem 85, 180192.CrossRefGoogle ScholarPubMed
36Youdim, KA, Shukitt-Hale, B & Joseph, JA (2004) Flavonoids and the brain: interactions at the blood–brain barrier and their physiological effects on the central nervous system. Free Radic Biol Med 37, 16831693.CrossRefGoogle ScholarPubMed
37Youdim, KA, Qaiser, MZ, Begley, DJ, Rice-Evans, CA & Abbott, NJ (2004) Flavonoid permeability across an in situ model of the blood–brain barrier. Free Radic Biol Med 36, 592604.CrossRefGoogle Scholar
38Aasmundstad, TA, Morland, J & Paulsen, RE (1995) Distribution of morphine 6-glucuronide and morphine across the blood–brain barrier in awake, freely moving rats investigated by in vivo microdialysis sampling. J Pharmacol Exp Ther 275, 435441.Google ScholarPubMed
39Sperker, B, Backman, JT & Kroemer, HK (1997) The role of beta-glucuronidase in drug disposition and drug targeting in humans. Clin Pharmacokinet 33, 1831.CrossRefGoogle ScholarPubMed
40Kroemer, HK & Klotz, U (1992) Glucuronidation of drugs. A re-evaluation of the pharmacological significance of the conjugates and modulating factors. Clin Pharmacokinet 23, 292310.CrossRefGoogle ScholarPubMed
41Lin, JH & Yamazaki, M (2003) Role of P-glycoprotein in pharmacokinetics: clinical implications. Clin Pharmacokinet 42, 5998.CrossRefGoogle ScholarPubMed
42Suganuma, M, Okabe, S, Oniyama, M, Tada, Y, Ito, H & Fujiki, H (1998) Wide distribution of [3H]( − )-epigallocatechin gallate, a cancer preventive tea polyphenol, in mouse tissue. Carcinogenesis 19, 17711776.CrossRefGoogle ScholarPubMed
43Abd El Mohsen, MM, Kuhnle, G, Rechner, AR, Schroeter, H, Rose, S, Jenner, P & Rice-Evans, CA (2002) Uptake and metabolism of epicatechin and its access to the brain after oral ingestion. Free Radic Biol Med 33, 16931702.CrossRefGoogle Scholar
44Talavera, S, Felgines, C, Texier, O, Besson, C, Gil-Izquierdo, A, Lamaison, JL & Remesy, C (2005) Anthocyanin metabolism in rats and their distribution to digestive area, kidney, and brain. J Agric Food Chem 53, 39023908.CrossRefGoogle ScholarPubMed
45El Mohsen, MA, Marks, J, Kuhnle, G, Moore, K, Debnam, E, Kaila, SS, Rice-Evans, C & Spencer, JP (2006) Absorption, tissue distribution and excretion of pelargonidin and its metabolites following oral administration to rats. Br J Nutr 95, 5158.CrossRefGoogle ScholarPubMed
46Andres-Lacueva, C, Shukitt-Hale, B, Galli, RL, Jauregui, O, Lamuela-Raventos, RM & Joseph, JA (2005) Anthocyanins in aged blueberry-fed rats are found centrally and may enhance memory. Nutr Neurosci 8, 111120.CrossRefGoogle ScholarPubMed
47Halliwell, B, Zhao, K & Whiteman, M (2000) The gastrointestinal tract: a major site of antioxidant action? Free Radic Res 33, 819830.CrossRefGoogle Scholar
48Joseph, JA, Shukitt-Hale, B, Denisova, NA, Prior, RL, Cao, G, Martin, A, Taglialatela, G & Bickford, PC (1998) Long-term dietary strawberry, spinach, or vitamin E supplementation retards the onset of age-related neuronal signal-transduction and cognitive behavioral deficits. J Neurosci 18, 80478055.CrossRefGoogle ScholarPubMed
49Joseph, JA, Shukitt-Hale, B, Denisova, NA, Bielinski, D, Martin, A, McEwen, JJ & Bickford, PC (1999) Reversals of age-related declines in neuronal signal transduction, cognitive, and motor behavioral deficits with blueberry, spinach, or strawberry dietary supplementation. J Neurosci 19, 81148121.CrossRefGoogle ScholarPubMed
50Casadesus, G, Shukitt-Hale, B, Stellwagen, HM, Zhu, X, Lee, HG, Smith, MA & Joseph, JA (2004) Modulation of hippocampal plasticity and cognitive behavior by short-term blueberry supplementation in aged rats. Nutr Neurosci 7, 309316.CrossRefGoogle ScholarPubMed
51Shukitt-Hale, B, Smith, DE, Meydani, M & Joseph, JA (1999) The effects of dietary antioxidants on psychomotor performance in aged mice. Exp Gerontol 34, 797808.Google ScholarPubMed
52Joseph, JA, Shukitt-Hale, B & Casadesus, G (2005) Reversing the deleterious effects of aging on neuronal communication and behavior: beneficial properties of fruit polyphenolic compounds. Am J Clin Nutr 81, 313S316S.CrossRefGoogle ScholarPubMed
53Goyarzu, P, Malin, DH, Lau, FC, et al. (2004) Blueberry supplemented diet: effects on object recognition memory and nuclear factor-kappa B levels in aged rats. Nutr Neurosci 7, 7583.CrossRefGoogle ScholarPubMed
54Joseph, JA, Denisova, NA, Arendash, G, Gordon, M, Diamond, D, Shukitt-Hale, B & Morgan, D (2003) Blueberry supplementation enhances signaling and prevents behavioral deficits in an Alzheimer disease model. Nutr Neurosci 6, 153162.CrossRefGoogle Scholar
55Ramirez, MR, Izquierdo, I, do Carmo Bassols, RM, Zuanazzi, JA, Barros, D & Henriques, AT (2005) Effect of lyophilised Vaccinium berries on memory, anxiety and locomotion in adult rats. Pharmacol Res 52, 457462.CrossRefGoogle ScholarPubMed
56Schroeter, H, Heiss, C, Balzer, J, Kleinbongard, P, Keen, CL, Hollenberg, NK, Sies, H, Kwik-Uribe, C, Schmitz, HH & Kelm, M (2006) ( − )-Epicatechin mediates beneficial effects of flavanol-rich cocoa on vascular function in humans. Proc Natl Acad Sci U S A 103, 10241029.CrossRefGoogle ScholarPubMed
57Francis, ST, Head, K, Morris, PG & Macdonald, IA (2006) The effect of flavanol-rich cocoa on the fMRI response to a cognitive task in healthy young people. J Cardiovasc Pharmacol 47, Suppl. 2, S215S220.CrossRefGoogle ScholarPubMed
58Fisher, ND, Sorond, FA & Hollenberg, NK (2006) Cocoa flavanols and brain perfusion. J Cardiovasc Pharmacol 47, Suppl. 2, S210S214.CrossRefGoogle ScholarPubMed
59Gage, FH (2000) Mammalian neural stem cells. Science 287, 14331438.CrossRefGoogle ScholarPubMed
60Palmer, TD, Willhoite, AR & Gage, FH (2000) Vascular niche for adult hippocampal neurogenesis. J Comp Neurol 425, 479494.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
61Harris, KM & Kater, SB (1994) Dendritic spines: cellular specializations imparting both stability and flexibility to synaptic function. Annu Rev Neurosci 17, 341371.CrossRefGoogle ScholarPubMed
62Carew, TJ (1996) Molecular enhancement of memory formation. Neuron 16, 58.CrossRefGoogle ScholarPubMed
63Martin, KC, Barad, M & Kandel, ER (2000) Local protein synthesis and its role in synapse-specific plasticity. Curr Opin Neurobiol 10, 587592.CrossRefGoogle ScholarPubMed
64Kelleher, RJ III, Govindarajan, A & Tonegawa, S (2004) Translational regulatory mechanisms in persistent forms of synaptic plasticity. Neuron 44, 5973.CrossRefGoogle ScholarPubMed
65Impey, S, McCorkle, SR, Cha-Molstad, H, Dwyer, JM, Yochum, GS, Boss, JM, McWeeney, S, Dunn, JJ, Mandel, G & Goodman, RH (2004) Defining the CREB regulon: a genome-wide analysis of transcription factor regulatory regions. Cell 119, 10411054.Google ScholarPubMed
66Barco, A, Bailey, CH & Kandel, ER (2006) Common molecular mechanisms in explicit and implicit memory. J Neurochem 97, 15201533.CrossRefGoogle ScholarPubMed
67Bourtchuladze, R, Frenguelli, B, Blendy, J, Cioffi, D, Schutz, G & Silva, AJ (1994) Deficient long-term memory in mice with a targeted mutation of the cAMP-responsive element-binding protein. Cell 79, 5968.CrossRefGoogle ScholarPubMed
68Tully, T, Bourtchouladze, R, Scott, R & Tallman, J (2003) Targeting the CREB pathway for memory enhancers. Nat Rev Drug Discov 2, 267277.CrossRefGoogle ScholarPubMed
69Pham, TA, Impey, S, Storm, DR & Stryker, MP (1999) CRE-mediated gene transcription in neocortical neuronal plasticity during the developmental critical period. Neuron 22, 6372.CrossRefGoogle ScholarPubMed
70Impey, S, Smith, DM, Obrietan, K, Donahue, R, Wade, C & Storm, DR (1998) Stimulation of cAMP response element (CRE)-mediated transcription during contextual learning. Nat Neurosci 1, 595601.CrossRefGoogle ScholarPubMed
71Impey, S, Mark, M, Villacres, EC, Poser, S, Chavkin, C & Storm, DR (1996) Induction of CRE-mediated gene expression by stimuli that generate long-lasting LTP in area CA1 of the hippocampus. Neuron 16, 973982.CrossRefGoogle ScholarPubMed
72Schroeter, H, Spencer, JP, Rice-Evans, C & Williams, RJ (2001) Flavonoids protect neurons from oxidized low-density-lipoprotein-induced apoptosis involving c-Jun N-terminal kinase (JNK), c-Jun and caspase-3. Biochem J 358, 547557.CrossRefGoogle ScholarPubMed
73Schroeter, H, Bahia, P, Spencer, JPE, Sheppard, O, Rattray, M, Rice-Evans, C & Williams, RJ (2007) ( − )-Epicatechin stimulates ERK-dependent cyclic AMP response element activity and upregulates GLUR2 in cortical neurons J Neurochem 101, 15961606..CrossRefGoogle ScholarPubMed
74Maher, P, Akaishi, T & Abe, K (2006) Flavonoid fisetin promotes ERK-dependent long-term potentiation and enhances memory. Proc Natl Acad Sci U S A 103, 1656816573.CrossRefGoogle ScholarPubMed
75Conkright, MD, Guzman, E, Flechner, L, Su, AI, Hogenesch, JB & Montminy, M (2003) Genome-wide analysis of CREB target genes reveals a core promoter requirement for cAMP responsiveness. Mol Cell 11, 11011108.CrossRefGoogle ScholarPubMed
76Tao, X, Finkbeiner, S, Arnold, DB, Shaywitz, AJ & Greenberg, ME (1998) Ca2+ influx regulates BDNF transcription by a CREB family transcription factor-dependent mechanism. Neuron 20, 709726.CrossRefGoogle ScholarPubMed
77Shieh, PB, Hu, SC, Bobb, K, Timmusk, T & Ghosh, A (1998) Identification of a signaling pathway involved in calcium regulation of BDNF expression. Neuron 20, 727740.CrossRefGoogle ScholarPubMed
78Bramham, CR & Messaoudi, E (2005) BDNF function in adult synaptic plasticity: the synaptic consolidation hypothesis. Prog Neurobiol 76, 99125.CrossRefGoogle ScholarPubMed
79Patterson, SL, Grover, LM, Schwartzkroin, PA & Bothwell, M (1992) Neurotrophin expression in rat hippocampal slices: a stimulus paradigm inducing LTP in CA1 evokes increases in BDNF and NT-3 mRNAs. Neuron 9, 10811088.CrossRefGoogle ScholarPubMed
80Thomas, K & Davies, A (2005) Neurotrophins: a ticket to ride for BDNF. Curr Biol 15, R262R264.CrossRefGoogle ScholarPubMed
81Peng, S, Wuu, J, Mufson, EJ & Fahnestock, M (2005) Precursor form of brain-derived neurotrophic factor and mature brain-derived neurotrophic factor are decreased in the pre-clinical stages of Alzheimer's disease. J Neurochem 93, 14121421.CrossRefGoogle ScholarPubMed
82Michalski, B & Fahnestock, M (2003) Pro-brain-derived neurotrophic factor is decreased in parietal cortex in Alzheimer's disease. Brain Res Mol Brain Res 111, 148154.CrossRefGoogle ScholarPubMed
83Egan, MF, Kojima, M, Callicott, JH, et al. (2003) The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function. Cell 112, 257269.CrossRefGoogle ScholarPubMed
84Linnarsson, S, Bjorklund, A & Ernfors, P (1997) Learning deficit in BDNF mutant mice. Eur J Neurosci 9, 25812587.CrossRefGoogle ScholarPubMed
85Mu, JS, Li, WP, Yao, ZB & Zhou, XF (1999) Deprivation of endogenous brain-derived neurotrophic factor results in impairment of spatial learning and memory in adult rats. Brain Res 835, 259265.CrossRefGoogle ScholarPubMed
86Minichiello, L, Korte, M, Wolfer, D, Kuhn, R, Unsicker, K, Cestari, V, Rossi-Arnaud, C, Lipp, HP, Bonhoeffer, T & Klein, R (1999) Essential role for TrkB receptors in hippocampus-mediated learning. Neuron 24, 401414.CrossRefGoogle ScholarPubMed
87Wullschleger, S, Loewith, R & Hall, MN (2006) TOR signaling in growth and metabolism. Cell 124, 471484.CrossRefGoogle ScholarPubMed
88Schratt, GM, Nigh, EA, Chen, WG, Hu, L & Greenberg, ME (2004) BDNF regulates the translation of a select group of mRNAs by a mammalian target of rapamycin-phosphatidylinositol 3-kinase-dependent pathway during neuronal development. J Neurosci 24, 73667377.CrossRefGoogle ScholarPubMed
89Yin, Y, Edelman, GM & Vanderklish, PW (2002) The brain-derived neurotrophic factor enhances synthesis of Arc in synaptoneurosomes. Proc Natl Acad Sci U S A 99, 23682373.CrossRefGoogle ScholarPubMed
90Waltereit, R, Dammermann, B, Wulff, P, Scafidi, J, Staubli, U, Kauselmann, G, Bundman, M & Kuhl, D (2001) Arg3.1/Arc mRNA induction by Ca2+ and cAMP requires protein kinase A and mitogen-activated protein kinase/extracellular regulated kinase activation. J Neurosci 21, 54845493.CrossRefGoogle ScholarPubMed
91Soule, J, Messaoudi, E & Bramham, CR (2006) Brain-derived neurotrophic factor and control of synaptic consolidation in the adult brain. Biochem Soc Trans 34, 600604.CrossRefGoogle ScholarPubMed
92Lyford, GL, Yamagata, K, Kaufmann, WE, Barnes, CA, Sanders, LK, Copeland, NG, Gilbert, DJ, Jenkins, NA, Lanahan, AA & Worley, PF (1995) Arc, a growth factor and activity-regulated gene, encodes a novel cytoskeleton-associated protein that is enriched in neuronal dendrites. Neuron 14, 433445.CrossRefGoogle ScholarPubMed
93Reznichenko, L, Amit, T, Youdim, MB & Mandel, S (2005) Green tea polyphenol ( − )-epigallocatechin-3-gallate induces neurorescue of long-term serum-deprived PC12 cells and promotes neurite outgrowth. J Neurochem 93, 11571167.CrossRefGoogle ScholarPubMed
94Hirsch, EC, Hunot, S & Hartmann, A (2005) Neuroinflammatory processes in Parkinson's disease. Parkinsonism Relat Disord 11, Suppl. 1, S9S15.CrossRefGoogle ScholarPubMed
95McGeer, EG & McGeer, PL (2003) Inflammatory processes in Alzheimer's disease. Prog Neuropsychopharmacol Biol Psychiatry 27, 741749.CrossRefGoogle ScholarPubMed
96Zheng, Z, Lee, JE & Yenari, MA (2003) Stroke: molecular mechanisms and potential targets for treatment. Curr Mol Med 3, 361372.CrossRefGoogle ScholarPubMed
97Casper, D, Yaparpalvi, U, Rempel, N & Werner, P (2000) Ibuprofen protects dopaminergic neurons against glutamate toxicity in vitro. Neurosci Lett 289, 201204.CrossRefGoogle ScholarPubMed
98Chen, H, Zhang, SM, Hernan, MA, Schwarzschild, MA, Willett, WC, Colditz, GA, Speizer, FE & Ascherio, A (2003) Nonsteroidal anti-inflammatory drugs and the risk of Parkinson disease. Arch Neurol 60, 10591064.CrossRefGoogle ScholarPubMed
99McGeer, EG & McGeer, PL (1997) The role of the immune system in neurodegenerative disorders. Mov Disord 12, 855858.CrossRefGoogle ScholarPubMed
100Vila, M, Jackson-Lewis, V, Guegan, C, Wu, DC, Teismann, P, Choi, DK, Tieu, K & Przedborski, S (2001) The role of glial cells in Parkinson's disease. Curr Opin Neurol 14, 483489.CrossRefGoogle ScholarPubMed
101Kim, YS & Joh, TH (2006) Microglia, major player in the brain inflammation: their roles in the pathogenesis of Parkinson's disease. Exp Mol Med 38, 333347.CrossRefGoogle ScholarPubMed
102Allan, SM & Rothwell, NJ (2003) Inflammation in central nervous system injury. Philos Trans R Soc Lond B Biol Sci 358, 16691677.CrossRefGoogle ScholarPubMed
103Stewart, VC & Heales, SJ (2003) Nitric oxide-induced mitochondrial dysfunction: implications for neurodegeneration. Free Radic Biol Med 34, 287303.CrossRefGoogle ScholarPubMed
104Moncada, S & Bolanos, JP (2006) Nitric oxide, cell bioenergetics and neurodegeneration. J Neurochem 97, 16761689.CrossRefGoogle ScholarPubMed
105Kozuka, N, Itofusa, R, Kudo, Y & Morita, M (2005) Lipopolysaccharide and proinflammatory cytokines require different astrocyte states to induce nitric oxide production. J Neurosci Res 82, 717728.CrossRefGoogle ScholarPubMed
106Bal-Price, A, Matthias, A & Brown, GC (2002) Stimulation of the NADPH oxidase in activated rat microglia removes nitric oxide but induces peroxynitrite production. J Neurochem 80, 7380.CrossRefGoogle ScholarPubMed
107Fiebich, BL, Lieb, K, Engels, S & Heinrich, M (2002) Inhibition of LPS-induced p42/44 MAP kinase activation and iNOS/NO synthesis by parthenolide in rat primary microglial cells. J Neuroimmunol 132, 1824.CrossRefGoogle ScholarPubMed
108Bhat, NR, Zhang, P, Lee, JC & Hogan, EL (1998) Extracellular signal-regulated kinase and p38 subgroups of mitogen-activated protein kinases regulate inducible nitric oxide synthase and tumor necrosis factor-alpha gene expression in endotoxin-stimulated primary glial cultures. J Neurosci 18, 16331641.CrossRefGoogle ScholarPubMed
109Marcus, JS, Karackattu, SL, Fleegal, MA & Sumners, C (2003) Cytokine-stimulated inducible nitric oxide synthase expression in astroglia: role of Erk mitogen-activated protein kinase and NF-kappaB. Glia 41, 152160.CrossRefGoogle ScholarPubMed
110Pawate, S & Bhat, NR (2006) C-Jun N-terminal kinase (JNK) regulation of iNOS expression in glial cells: predominant role of JNK1 isoform. Antioxid Redox Signal 8, 903909.CrossRefGoogle ScholarPubMed
111Li, R, Huang, YG, Fang, D & Le, WD (2004) ( − )-Epigallocatechin gallate inhibits lipopolysaccharide-induced microglial activation and protects against inflammation-mediated dopaminergic neuronal injury. J Neurosci Res 78, 723731.CrossRefGoogle ScholarPubMed
112Huang, Q, Wu, LJ, Tashiro, S, Gao, HY, Onodera, S & Ikejima, T (2005) (+)-Catechin, an ingredient of green tea, protects murine microglia from oxidative stress-induced DNA damage and cell cycle arrest. J Pharmacol Sci 98, 1624.CrossRefGoogle ScholarPubMed
113Lee, H, Kim, YO, Kim, H, Kim, SY, Noh, HS, Kang, SS, Cho, GJ, Choi, WS & Suk, K (2003) Flavonoid wogonin from medicinal herb is neuroprotective by inhibiting inflammatory activation of microglia. FASEB J 17, 19431944.CrossRefGoogle ScholarPubMed
114Shen, SC, Lee, WR, Lin, HY, Huang, HC, Ko, CH, Yang, LL & Chen, YC (2002) In vitro and in vivo inhibitory activities of rutin, wogonin, and quercetin on lipopolysaccharide-induced nitric oxide and prostaglandin E(2) production. Eur J Pharmacol 446, 187194.CrossRefGoogle ScholarPubMed
115Woo, KJ, Lim, JH, Suh, SI, Kwon, YK, Shin, SW, Kim, SC, Choi, YH, Park, JW & Kwon, TK (2006) Differential inhibitory effects of baicalein and baicalin on LPS-induced cyclooxygenase-2 expression through inhibition of C/EBPbeta DNA-binding activity. Immunobiology 211, 359368.CrossRefGoogle ScholarPubMed
116Kim, H, Kim, YS, Kim, SY & Suk, K (2001) The plant flavonoid wogonin suppresses death of activated C6 rat glial cells by inhibiting nitric oxide production. Neurosci Lett 309, 6771.CrossRefGoogle ScholarPubMed
117Chen, CJ, Raung, SL, Liao, SL & Chen, SY (2004) Inhibition of inducible nitric oxide synthase expression by baicalein in endotoxin/cytokine-stimulated microglia. Biochem Pharmacol 67, 957965.CrossRefGoogle ScholarPubMed
118Chen, JC, Ho, FM, Pei-Dawn, LC, Chen, CP, Jeng, KC, Hsu, HB, Lee, ST, Wen, TW & Lin, WW (2005) Inhibition of iNOS gene expression by quercetin is mediated by the inhibition of IkappaB kinase, nuclear factor-kappa B and STAT1, and depends on heme oxygenase-1 induction in mouse BV-2 microglia. Eur J Pharmacol 521, 920.CrossRefGoogle ScholarPubMed
119Matter, WF, Brown, RF & Vlahos, CJ (1992) The inhibition of phosphatidylinositol 3-kinase by quercetin and analogs. Biochem Biophys Res Commun 186, 624631.CrossRefGoogle ScholarPubMed
120Vlahos, CJ, Matter, WF, Hui, KY & Brown, RF (1994) A specific inhibitor of phosphatidylinositol 3-kinase, 2-(4-morpholinyl)-8-phenyl-4H-1-benzopyran-4-one (LY294002). J Biol Chem 269, 52415248.CrossRefGoogle ScholarPubMed
121Agullo, G, Gamet-Payrastre, L, Manenti, S, Viala, C, Remesy, C, Chap, H & Payrastre, B (1997) Relationship between flavonoid structure and inhibition of phosphatidylinositol 3-kinase: a comparison with tyrosine kinase and protein kinase C inhibition. Biochem Pharmacol 53, 16491657.CrossRefGoogle ScholarPubMed
122Gamet-Payrastre, L, Manenti, S, Gratacap, MP, Tulliez, J, Chap, H & Payrastre, B (1999) Flavonoids and the inhibition of PKC and PI 3-kinase. Gen Pharmacol 32, 279286.CrossRefGoogle ScholarPubMed
123Kong, AN, Yu, R, Chen, C, Mandlekar, S & Primiano, T (2000) Signal transduction events elicited by natural products: role of MAPK and caspase pathways in homeostatic response and induction of apoptosis. Arch Pharm Res 23, 116.CrossRefGoogle ScholarPubMed
124Lin, CH, Yeh, SH, Lin, CH, Lu, KT, Leu, TH, Chang, WC & Gean, PW (2001) A role for the PI-3 kinase signaling pathway in fear conditioning and synaptic plasticity in the amygdala. Neuron 31, 841851.CrossRefGoogle ScholarPubMed
125Sweatt, JD (2001) Memory mechanisms: the yin and yang of protein phosphorylation. Curr Biol 11, R391R394.CrossRefGoogle ScholarPubMed
126Conseil, G, Baubichon-Cortay, H, Dayan, G, Jault, JM, Barron, D & Di Pietro, A (1998) Flavonoids: a class of modulators with bifunctional interactions at vicinal ATP- and steroid-binding sites on mouse P-glycoprotein. Proc Natl Acad Sci U S A 95, 98319836.CrossRefGoogle ScholarPubMed
127Di Pietro, A, Godinot, C, Bouillant, ML & Gautheron, DC (1975) Pig heart mitochondrial ATPase: properties of purified and membrane-bound enzyme. Effects of flavonoids. Biochimie 57, 959967.CrossRefGoogle ScholarPubMed
128Barzilai, A & Rahamimoff, H (1983) Inhibition of Ca2+-transport ATPase from synaptosomal vesicles by flavonoids. Biochim Biophys Acta 730, 245254.CrossRefGoogle ScholarPubMed
129Revuelta, MP, Cantabrana, B & Hidalgo, A (1997) Depolarization-dependent effect of flavonoids in rat uterine smooth muscle contraction elicited by CaCl2. Gen Pharmacol 29, 847857.CrossRefGoogle ScholarPubMed
130Lee, SF & Lin, JK (1997) Inhibitory effects of phytopolyphenols on TPA-induced transformation, PKC activation, and c-jun expression in mouse fibroblast cells. Nutr Cancer 28, 177183.CrossRefGoogle ScholarPubMed
131Ursini, F, Maiorino, M, Morazzoni, P, Roveri, A & Pifferi, G (1994) A novel antioxidant flavonoid (IdB 1031) affecting molecular mechanisms of cellular activation. Free Radic Biol Med 16, 547553.CrossRefGoogle ScholarPubMed
132Kantengwa, S & Polla, BS (1991) Flavonoids, but not protein kinase C inhibitors, prevent stress protein synthesis during erythrophagocytosis. Biochem Biophys Res Commun 180, 308314.CrossRefGoogle Scholar
133Rosenblat, M, Belinky, P, Vaya, J, Levy, R, Hayek, T, Coleman, R, Merchav, S & Aviram, M (1999) Macrophage enrichment with the isoflavan glabridin inhibits NADPH oxidase-induced cell-mediated oxidation of low density lipoprotein. A possible role for protein kinase C. J Biol Chem 274, 1379013799.CrossRefGoogle ScholarPubMed
134Boege, F, Straub, T, Kehr, A, Boesenberg, C, Christiansen, K, Andersen, A, Jakob, F & Kohrle, J (1996) Selected novel flavones inhibit the DNA binding or the DNA religation step of eukaryotic topoisomerase I. J Biol Chem 271, 22622270.CrossRefGoogle ScholarPubMed
135Medina, JH, Viola, H, Wolfman, C, Marder, M, Wasowski, C, Calvo, D & Paladini, AC (1997) Overview—flavonoids: a new family of benzodiazepine receptor ligands. Neurochem Res 22, 419425.CrossRefGoogle ScholarPubMed
136Dekermendjian, K, Kahnberg, P, Witt, MR, Sterner, O, Nielsen, M & Liljefors, T (1999) Structure–activity relationships and molecular modeling analysis of flavonoids binding to the benzodiazepine site of the rat brain GABA(A) receptor complex. J Med Chem 42, 43434350.CrossRefGoogle Scholar
137Fischer, PM & Lane, DP (2000) Inhibitors of cyclin-dependent kinases as anti-cancer therapeutics. Curr Med Chem 7, 12131245.CrossRefGoogle ScholarPubMed
138Huang, YT, Hwang, JJ, Lee, PP, Ke, FC, Huang, JH, Huang, CJ, Kandaswami, C, Middleton, E Jr & Lee, MT (1999) Effects of luteolin and quercetin, inhibitors of tyrosine kinase, on cell growth and metastasis-associated properties in A431 cells overexpressing epidermal growth factor receptor. Br J Pharmacol 128, 9991010.CrossRefGoogle ScholarPubMed
139So, FV, Guthrie, N, Chambers, AF, Moussa, M & Carroll, KK (1996) Inhibition of human breast cancer cell proliferation and delay of mammary tumorigenesis by flavonoids and citrus juices. Nutr Cancer 26, 167181.CrossRefGoogle ScholarPubMed
140Green, DR & Reed, JC (1998) Mitochondria and apoptosis. Science 281, 13091312.CrossRefGoogle ScholarPubMed
141Tatton, WG & Olanow, CW (1999) Apoptosis in neurodegenerative diseases: the role of mitochondria. Biochim Biophys Acta 1410, 195213.CrossRefGoogle ScholarPubMed
142Goyal, L (2001) Cell death inhibition: keeping caspases in check. Cell 104, 805808.CrossRefGoogle ScholarPubMed
143Srinivasula, SM, Hegde, R, Saleh, A, et al. (2001) A conserved XIAP-interaction motif in caspase-9 and Smac/DIABLO regulates caspase activity and apoptosis. Nature 410, 112116.CrossRefGoogle ScholarPubMed
144Cobb, MH & Goldsmith, EJ (1995) How MAP kinases are regulated. J Biol Chem 270, 1484314846.CrossRefGoogle ScholarPubMed
145Goldsmith, EJ & Cobb, MH (1994) Protein kinases. Curr Opin Struct Biol 4, 833840.CrossRefGoogle ScholarPubMed
146Marshall, CJ (1994) Signal transduction. Hot lips and phosphorylation of protein kinases. Nature 367, 686.CrossRefGoogle ScholarPubMed
147Karin, M (1995) The regulation of AP-1 activity by mitogen-activated protein kinases. J Biol Chem 270, 1648316486.CrossRefGoogle ScholarPubMed
148Xia, Z, Dickens, M, Raingeaud, J, Davis, RJ & Greenberg, ME (1995) Opposing effects of ERK and JNK-p38 MAP kinases on apoptosis. Science 270, 13261331.CrossRefGoogle ScholarPubMed
149Anderson, CN & Tolkovsky, AM (1999) A role for MAPK/ERK in sympathetic neuron survival: protection against a p53-dependent, JNK-independent induction of apoptosis by cytosine arabinoside. J Neurosci 19, 664673.CrossRefGoogle ScholarPubMed
150Bonni, A, Brunet, A, West, AE, Datta, SR, Takasu, MA & Greenberg, ME (1999) Cell survival promoted by the Ras-MAPK signaling pathway by transcription-dependent and -independent mechanisms. Science 286, 13581362.CrossRefGoogle ScholarPubMed
151Kaplan, DR & Miller, FD (2000) Neurotrophin signal transduction in the nervous system. Curr Opin Neurobiol 10, 381391.CrossRefGoogle ScholarPubMed
152Crossthwaite, AJ, Hasan, S & Williams, RJ (2002) Hydrogen peroxide-mediated phosphorylation of ERK1/2, Akt/PKB and JNK in cortical neurones: dependence on Ca(2+) and PI3-kinase. J Neurochem 80, 2435.CrossRefGoogle ScholarPubMed
153Mielke, K & Herdegen, T (2000) JNK and p38 stresskinases—degenerative effectors of signal-transduction-cascades in the nervous system. Prog Neurobiol 61, 4560.CrossRefGoogle ScholarPubMed
154Yuan, J & Yankner, BA (2000) Apoptosis in the nervous system. Nature 407, 802809.CrossRefGoogle ScholarPubMed
155Behrens, A, Sibilia, M & Wagner, EF (1999) Amino-terminal phosphorylation of c-Jun regulates stress-induced apoptosis and cellular proliferation. Nat Genet 21, 326329.CrossRefGoogle ScholarPubMed
156Davis, RJ (2000) Signal transduction by the JNK group of MAP kinases. Cell 103, 239252.CrossRefGoogle ScholarPubMed
157Kobuchi, H, Roy, S, Sen, CK, Nguyen, HG & Packer, L (1999) Quercetin inhibits inducible ICAM-1 expression in human endothelial cells through the JNK pathway. Am J Physiol 277, C403C411.CrossRefGoogle ScholarPubMed
158Herdegen, T, Skene, P & Bahr, M (1997) The c-Jun transcription factor—bipotential mediator of neuronal death, survival and regeneration. Trends Neurosci 20, 227231.CrossRefGoogle ScholarPubMed
159Castagne, V & Clarke, PG (1999) Inhibitors of mitogen-activated protein kinases protect axotomized developing neurons. Brain Res 842, 215219.CrossRefGoogle ScholarPubMed
160Castagne, V, Gautschi, M, Lefevre, K, Posada, A & Clarke, PG (1999) Relationships between neuronal death and the cellular redox status. Focus on the developing nervous system. Prog Neurobiol 59, 397423.CrossRefGoogle ScholarPubMed
161Hung, SP, Hsu, JR, Lo, CP, Huang, HJ, Wang, JP & Chen, ST (2005) Genistein-induced neuronal differentiation is associated with activation of extracellular signal-regulated kinases and upregulation of p21 and N-cadherin. J Cell Biochem 96, 10611070.CrossRefGoogle ScholarPubMed
162Llorens, F, Garcia, L, Itarte, E & Gomez, N (2002) Apigenin and LY294002 prolong EGF-stimulated ERK1/2 activation in PC12 cells but are unable to induce full differentiation. FEBS Lett 510, 149153.CrossRefGoogle ScholarPubMed
163Dudley, DT, Pang, L, Decker, SJ, Bridges, AJ & Saltiel, AR (1995) A synthetic inhibitor of the mitogen-activated protein kinase cascade. Proc Natl Acad Sci U S A 92, 76867689.CrossRefGoogle ScholarPubMed
164Pang, L, Sawada, T, Decker, SJ & Saltiel, AR (1995) Inhibition of MAP kinase kinase blocks the differentiation of PC-12 cells induced by nerve growth factor. J Biol Chem 270, 1358513588.CrossRefGoogle ScholarPubMed
165Alessi, DR, Cuenda, A, Cohen, P, Dudley, DT & Saltiel, AR (1995) PD 098059 is a specific inhibitor of the activation of mitogen-activated protein kinase kinase in vitro and in vivo. J Biol Chem 270, 2748927494.CrossRefGoogle ScholarPubMed
166Lazar, DF, Wiese, RJ, Brady, MJ, Mastick, CC, Waters, SB, Yamauchi, K, Pessin, JE, Cuatrecasas, P & Saltiel, AR (1995) Mitogen-activated protein kinase kinase inhibition does not block the stimulation of glucose utilization by insulin. J Biol Chem 270, 2080120807.CrossRefGoogle Scholar
167Levites, Y, Amit, T, Youdim, MB & Mandel, S (2002) Involvement of protein kinase C activation and cell survival/cell cycle genes in green tea polyphenol ( − )-epigallocatechin 3-gallate neuroprotective action. J Biol Chem 277, 3057430580.CrossRefGoogle ScholarPubMed
168Reznichenko, L, Amit, T, Youdim, MB & Mandel, S (2005) Green tea polyphenol ( − )-epigallocatechin-3-gallate induces neurorescue of long-term serum-deprived PC12 cells and promotes neurite outgrowth. J Neurochem 93, 11571167.CrossRefGoogle ScholarPubMed
169Walker, EH, Pacold, ME, Perisic, O, Stephens, L, Hawkins, PT, Wymann, MP & Williams, RL (2000) Structural determinants of phosphoinositide 3-kinase inhibition by wortmannin, LY294002, quercetin, myricetin, and staurosporine. Mol Cell 6, 909919.CrossRefGoogle ScholarPubMed
170Perkinton, MS, Sihra, TS & Williams, RJ (1999) Ca(2+)-permeable AMPA receptors induce phosphorylation of cAMP response element-binding protein through a phosphatidylinositol 3-kinase-dependent stimulation of the mitogen-activated protein kinase signaling cascade in neurons. J Neurosci 19, 58615874.CrossRefGoogle ScholarPubMed
171Jacobson, KA, Moro, S, Manthey, JA, West, PL & Ji, XD (2002) Interactions of flavones and other phytochemicals with adenosine receptors. Adv Exp Med Biol 505, 163171.CrossRefGoogle ScholarPubMed
172Johnston, GA (2005) GABA(A) receptor channel pharmacology. Curr Pharm Des 11, 18671885.CrossRefGoogle ScholarPubMed
173Adachi, N, Tomonaga, S, Tachibana, T, Denbow, DM & Furuse, M (2006) ( − )-Epigallocatechin gallate attenuates acute stress responses through GABAergic system in the brain. Eur J Pharmacol 531, 171175.CrossRefGoogle ScholarPubMed
174Han, YS, Bastianetto, S, Dumont, Y & Quirion, R (2006) Specific plasma membrane binding sites for polyphenols, including resveratrol, in the rat brain. J Pharmacol Exp Ther 318, 238245.CrossRefGoogle ScholarPubMed
175Nifli, AP, Bosson-Kouame, A, Papadopoulou, N, Kogia, C, Kampa, M, Castagnino, C, Stournaras, C, Vercauteren, J & Castanas, E (2005) Monomeric and oligomeric flavanols are agonists of membrane androgen receptors. Exp Cell Res 309, 329339.CrossRefGoogle ScholarPubMed
176Klinge, CM, Blankenship, KA, Risinger, KE, Bhatnagar, S, Noisin, EL, Sumanasekera, WK, Zhao, L, Brey, DM & Keynton, RS (2005) Resveratrol and estradiol rapidly activate MAPK signaling through estrogen receptors alpha and beta in endothelial cells. J Biol Chem 280, 74607468.CrossRefGoogle ScholarPubMed
177Camps, M, Nichols, A & Arkinstall, S (2000) Dual specificity phosphatases: a gene family for control of MAP kinase function. FASEB J 14, 616.CrossRefGoogle ScholarPubMed
178Kim, Y, Rice, AE & Denu, JM (2003) Intramolecular dephosphorylation of ERK by MKP3. Biochemistry 42, 1519715207.CrossRefGoogle ScholarPubMed
179Davis, RJ (1999) Signal transduction by the c-Jun N-terminal kinase. Biochem Soc Symp 64, 112.Google ScholarPubMed
180Leppa, S & Bohmann, D (1999) Diverse functions of JNK signaling and c-Jun in stress response and apoptosis. Oncogene 18, 61586162.CrossRefGoogle ScholarPubMed
181Ichijo, H (1999) From receptors to stress-activated MAP kinases. In Oncogene 18, pp. 60876093.Google Scholar
182Ichijo, H, Nishida, E, Irie, K, ten Dijke, P, Saitoh, M, Moriguchi, T, Takagi, M, Matsumoto, K, Miyazono, K & Gotoh, Y (1997) Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275, 9094.CrossRefGoogle ScholarPubMed
183Wang, XS, Diener, K, Jannuzzi, D, Trollinger, D, Tan, TH, Lichenstein, H, Zukowski, M & Yao, Z (1996) Molecular cloning and characterization of a novel protein kinase with a catalytic domain homologous to mitogen-activated protein kinase kinase kinase. J Biol Chem 271, 3160731611.Google ScholarPubMed
184Matsuzawa, A & Ichijo, H (2001) Molecular mechanisms of the decision between life and death: regulation of apoptosis by apoptosis signal-regulating kinase 1. J Biochem (Tokyo) 130, 18.CrossRefGoogle ScholarPubMed
185Schroeter, H, Williams, RJ, Matin, R, Iversen, L & Rice-Evans, CA (2000) Phenolic antioxidants attenuate neuronal cell death following uptake of oxidized low-density lipoprotein. Free Radic Biol Med 29, 12221233.CrossRefGoogle ScholarPubMed
186Spencer, JPE, Schroeter, H, Kuhnle, G, Srai, SK, Tyrrell, RM, Hahn, U & Rice-Evans, C (2001) Epicatechin and its in vivo metabolite, 3′-O-methyl epicatechin, protect human fibroblasts from oxidative-stress-induced cell death involving caspase-3 activation. Biochem J 354, 493500.CrossRefGoogle Scholar
187Wang, L, Matsushita, K, Araki, I & Takeda, M (2002) Inhibition of c-Jun N-terminal kinase ameliorates apoptosis induced by hydrogen peroxide in the kidney tubule epithelial cells (NRK-52E). Nephron 91, 142147.CrossRefGoogle ScholarPubMed
188Ishikawa, Y & Kitamura, M (2000) Anti-apoptotic effect of quercetin: intervention in the JNK- and ERK-mediated apoptotic pathways. Kidney Int 58, 10781087.CrossRefGoogle ScholarPubMed
189Uchida, K, Shiraishi, M, Naito, Y, Torii, Y, Nakamura, Y & Osawa, T (1999) Activation of stress signaling pathways by the end product of lipid peroxidation. 4-Hydroxy-2-nonenal is a potential inducer of intracellular peroxide production. J Biol Chem 274, 22342242.CrossRefGoogle ScholarPubMed
190Zhang, L, Chen, J & Fu, H (1999) Suppression of apoptosis signal-regulating kinase 1-induced cell death by 14-3-3 proteins. Proc Natl Acad Sci U S A 96, 85118515.CrossRefGoogle ScholarPubMed
191Park, HS, Park, E, Kim, MS, Ahn, K, Kim, IY & Choi, EJ (2000) Selenite inhibits the c-Jun N-terminal kinase/stress-activated protein kinase (JNK/SAPK) through a thiol redox mechanism. J Biol Chem 275, 25272531.CrossRefGoogle ScholarPubMed
192Adler, V, Yin, Z, Fuchs, SY, et al. (1999) Regulation of JNK signaling by GSTp. EMBO J 18, 13211334.CrossRefGoogle ScholarPubMed
193Adler, V, Yin, Z, Tew, KD & Ronai, Z (1999) Role of redox potential and reactive oxygen species in stress signaling. Oncogene 18, 61046111.CrossRefGoogle ScholarPubMed
194Monaco, R, Friedman, FK, Hyde, MJ, Chen, JM, Manolatus, S, Adler, V, Ronai, Z, Koslosky, W & Pincus, MR (1999) Identification of a glutathione-S-transferase effector domain for inhibition of jun kinase, by molecular dynamics. J Protein Chem 18, 859866.CrossRefGoogle ScholarPubMed
195Yin, Z, Ivanov, VN, Habelhah, H, Tew, K & Ronai, Z (2000) Glutathione S-transferase p elicits protection against H2O2-induced cell death via coordinated regulation of stress kinases. Cancer Res 60, 40534057.Google ScholarPubMed
196Spencer, JPE, Abd, El, Mohsen, MM & Rice-Evans, C (2003) Cellular uptake and metabolism of flavonoids and their metabolites: implications for their bioactivity. Arch Biochem Biophys 423, 148161.CrossRefGoogle Scholar
197Kennedy, SG, Wagner, AJ, Conzen, SD, Jordan, J, Bellacosa, A, Tsichlis, PN & Hay, N (1997) The PI 3-kinase/Akt signaling pathway delivers an anti-apoptotic signal. Genes Dev 11, 701713.CrossRefGoogle ScholarPubMed
198Coffer, PJ, Jin, J & Woodgett, JR (1998) Protein kinase B (c-Akt): a multifunctional mediator of phosphatidylinositol 3-kinase activation. Biochem J 335, [Pt 1], 113.CrossRefGoogle ScholarPubMed
199Miller, FD & Kaplan, DR (2001) Neurotrophin signalling pathways regulating neuronal apoptosis. Cell Mol Life Sci 58, 10451053.CrossRefGoogle ScholarPubMed
200Crowder, RJ & Freeman, RS (1998) Phosphatidylinositol 3-kinase and Akt protein kinase are necessary and sufficient for the survival of nerve growth factor-dependent sympathetic neurons. J Neurosci 18, 29332943.CrossRefGoogle ScholarPubMed
201Carpenter, CL & Cantley, LC (1990) Phosphoinositide kinases. Biochemistry 29, 1114711156.CrossRefGoogle ScholarPubMed
202Simpson, L & Parsons, R (2001) PTEN: life as a tumor suppressor. Exp Cell Res 264, 2941.CrossRefGoogle ScholarPubMed
203Neri, LM, Borgatti, P, Capitani, S & Martelli, AM (2002) The nuclear phosphoinositide 3-kinase/AKT pathway: a new second messenger system. Biochim Biophys Acta 1584, 7380.CrossRefGoogle ScholarPubMed
204Franke, TF, Kaplan, DR & Cantley, LC (1997) PI3K: downstream AKTion blocks apoptosis. Cell 88, 435437.CrossRefGoogle ScholarPubMed
205Burgering, BM & Coffer, PJ (1995) Protein kinase B (c-Akt) in phosphatidylinositol-3-OH kinase signal transduction. Nature 376, 599602.CrossRefGoogle ScholarPubMed
206Franke, TF, Yang, SI, Chan, TO, Datta, K, Kazlauskas, A, Morrison, DK, Kaplan, DR & Tsichlis, PN (1995) The protein kinase encoded by the Akt proto-oncogene is a target of the PDGF-activated phosphatidylinositol 3-kinase. Cell 81, 727736.CrossRefGoogle ScholarPubMed
207Cardone, MH, Roy, N, Stennicke, HR, Salvesen, GS, Franke, TF, Stanbridge, E, Frisch, S & Reed, JC (1998) Regulation of cell death protease caspase-9 by phosphorylation. Science 282, 13181321.CrossRefGoogle ScholarPubMed
208Burgering, BM & Kops, GJ (2002) Cell cycle and death control: long live Forkheads. Trends Biochem Sci 27, 352360.CrossRefGoogle ScholarPubMed
209Brunet, A, Bonni, A, Zigmond, MJ, Lin, MZ, Juo, P, Hu, LS, Anderson, MJ, Arden, KC, Blenis, J & Greenberg, ME (1999) Akt promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell 96, 857868.CrossRefGoogle ScholarPubMed
210Zha, J, Harada, H, Yang, E, Jockel, J & Korsmeyer, SJ (1996) Serine phosphorylation of death agonist BAD in response to survival factor results in binding to 14-3-3 not BCL-X(L). Cell 87, 619628.CrossRefGoogle Scholar
211Alessi, DR, Andjelkovic, M, Caudwell, B, Cron, P, Morrice, N, Cohen, P & Hemmings, BA (1996) Mechanism of activation of protein kinase B by insulin and IGF-1. EMBO J 15, 65416551.CrossRefGoogle ScholarPubMed
212Datta, SR, Dudek, H, Tao, X, Masters, S, Fu, H, Gotoh, Y & Greenberg, ME (1997) Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery. Cell 91, 231241.CrossRefGoogle Scholar
213del Peso, L, Gonzalez-Garcia, M, Page, C, Herrera, R & Nunez, G (1997) Interleukin-3-induced phosphorylation of BAD through the protein kinase Akt. Science 278, 687689.CrossRefGoogle ScholarPubMed
214Ferriola, PC, Cody, V & Middleton, E Jr (1989) Protein kinase C inhibition by plant flavonoids. Kinetic mechanisms and structure–activity relationships. Biochem Pharmacol 38, 16171624.CrossRefGoogle ScholarPubMed
215Cao, F, Jin, TY & Zhou, YF (2006) Inhibitory effect of isoflavones on prostate cancer cells and PTEN gene. Biomed Environ Sci 19, 3541.Google ScholarPubMed
216Gulati, N, Laudet, B, Zohrabian, VM, Murali, R & Jhanwar-Uniyal, M (2006) The antiproliferative effect of Quercetin in cancer cells is mediated via inhibition of the PI3K-Akt/PKB pathway. Anticancer Res 26, 11771181.Google ScholarPubMed
217Dave, B, Eason, RR, Till, SR, Geng, Y, Velarde, MC, Badger, TM & Simmen, RC (2005) The soy isoflavone genistein promotes apoptosis in mammary epithelial cells by inducing the tumor suppressor PTEN. Carcinogenesis 26, 17931803.CrossRefGoogle ScholarPubMed
218Cantley, LC & Neel, BG (1999) New insights into tumor suppression: PTEN suppresses tumor formation by restraining the phosphoinositide 3-kinase/AKT pathway. Proc Natl Acad Sci U S A 96, 42404245.CrossRefGoogle ScholarPubMed
219Myers, MP, Pass, I, Batty, IH, Van der KJ, KJ, Stolarov, JP, Hemmings, BA, Wigler, MH, Downes, CP & Tonks, NK (1998) The lipid phosphatase activity of PTEN is critical for its tumor supressor function. Proc Natl Acad Sci U S A 95, 1351313518.CrossRefGoogle ScholarPubMed
220Wu, X, Senechal, K, Neshat, MS, Whang, YE & Sawyers, CL (1998) The PTEN/MMAC1 tumor suppressor phosphatase functions as a negative regulator of the phosphoinositide 3-kinase/Akt pathway. Proc Natl Acad Sci U S A 95, 1558715591.CrossRefGoogle ScholarPubMed
221Wild, AC, Moinova, HR & Mulcahy, RT (1999) Regulation of gamma-glutamylcysteine synthetase subunit gene expression by the transcription factor Nrf2. J Biol Chem 274, 3362733636.CrossRefGoogle ScholarPubMed
222Moinova, HR & Mulcahy, RT (1999) Up-regulation of the human gamma-glutamylcysteine synthetase regulatory subunit gene involves binding of Nrf-2 to an electrophile responsive element. Biochem Biophys Res Commun 261, 661668.CrossRefGoogle Scholar
223Venugopal, R & Jaiswal, AK (1998) Nrf2 and Nrf1 in association with Jun proteins regulate antioxidant response element-mediated expression and coordinated induction of genes encoding detoxifying enzymes. Oncogene 17, 31453156.CrossRefGoogle ScholarPubMed
224Venugopal, R & Jaiswal, AK (1996) Nrf1 and Nrf2 positively and c-Fos and Fra1 negatively regulate the human antioxidant response element-mediated expression of NAD(P)H:quinone oxidoreductase1 gene. Proc Natl Acad Sci U S A 93, 1496014965.CrossRefGoogle ScholarPubMed
225Kang, KW, Lee, SJ & Kim, SG (2005) Molecular mechanism of nrf2 activation by oxidative stress. Antioxid Redox Signal 7, 16641673.CrossRefGoogle ScholarPubMed
226Itoh, K, Tong, KI & Yamamoto, M (2004) Molecular mechanism activating Nrf2-Keap1 pathway in regulation of adaptive response to electrophiles. Free Radic Biol Med 36, 12081213.CrossRefGoogle ScholarPubMed
227Leung, L, Kwong, M, Hou, S, Lee, C & Chan, JY (2003) Deficiency of the Nrf1 and Nrf2 transcription factors results in early embryonic lethality and severe oxidative stress. J Biol Chem 278, 4802148029.CrossRefGoogle ScholarPubMed
228Bloom, D, Dhakshinamoorthy, S & Jaiswal, AK (2002) Site-directed mutagenesis of cysteine to serine in the DNA binding region of Nrf2 decreases its capacity to upregulate antioxidant response element-mediated expression and antioxidant induction of NAD(P)H:quinone oxidoreductase1 gene. Oncogene 21, 21912200.CrossRefGoogle ScholarPubMed
229Andreadi, CK, Howells, LM, Atherfold, PA & Manson, MM (2006) Involvement of Nrf2, p38, B-Raf, and nuclear factor-kappaB, but not phosphatidylinositol 3-kinase, in induction of hemeoxygenase-1 by dietary polyphenols. Mol Pharmacol 69, 10331040.CrossRefGoogle Scholar
230Banning, A, Deubel, S, Kluth, D, Zhou, Z & Brigelius-Flohe, R (2005) The GI-GPx gene is a target for Nrf2. Mol Cell Biol 25, 49144923.CrossRefGoogle ScholarPubMed
231Hernandez-Montes, E, Pollard, SE, Vauzour, D, Jofre-Montseny, L, Rota, C, Rimbach, G, Weinberg, PD & Spencer, JP (2006) Activation of glutathione peroxidase via Nrf1 mediates genistein's protection against oxidative endothelial cell injury. Biochem Biophys Res Commun 346, 851859.CrossRefGoogle ScholarPubMed
232Itoh, K, Wakabayashi, N, Katoh, Y, Ishii, T, Igarashi, K, Engel, JD & Yamamoto, M (1999) Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev 13, 7686.CrossRefGoogle Scholar
233Hayes, JD & McMahon, M (2001) Molecular basis for the contribution of the antioxidant responsive element to cancer chemoprevention. Cancer Lett 174, 103113.CrossRefGoogle ScholarPubMed
234Alcaraz, MJ, Vicente, AM, Araico, A, Dominguez, JN, Terencio, MC & Ferrandiz, ML (2004) Role of nuclear factor-kappaB and heme oxygenase-1 in the mechanism of action of an anti-inflammatory chalcone derivative in RAW 264.7 cells. Br J Pharmacol 142, 11911199.CrossRefGoogle ScholarPubMed
235Kobayashi, M & Yamamoto, M (2005) Molecular mechanisms activating the Nrf2-Keap1 pathway of antioxidant gene regulation. Antioxid Redox Signal 7, 385394.CrossRefGoogle ScholarPubMed
236Nguyen, T, Sherratt, PJ, Huang, HC, Yang, CS & Pickett, CB (2003) Increased protein stability as a mechanism that enhances Nrf2-mediated transcriptional activation of the antioxidant response element. Degradation of Nrf2 by the 26 S proteasome. J Biol Chem 278, 45364541.CrossRefGoogle ScholarPubMed
237Nguyen, T, Huang, HC & Pickett, CB (2000) Transcriptional regulation of the antioxidant response element. Activation by Nrf2 and repression by MafK. J Biol Chem 275, 1546615473.CrossRefGoogle ScholarPubMed
238Sherratt, PJ, Huang, HC, Nguyen, T & Pickett, CB (2004) Role of protein phosphorylation in the regulation of NF-E2-related factor 2 activity. Methods Enzymol 378, 286301.CrossRefGoogle ScholarPubMed
239Yu, R, Mandlekar, S, Lei, W, Fahl, WE, Tan, TH & Kong, AN (2000) p38 mitogen-activated protein kinase negatively regulates the induction of phase II drug-metabolizing enzymes that detoxify carcinogens. J Biol Chem 275, 23222327.CrossRefGoogle ScholarPubMed
240Shen, G, Hebbar, V, Nair, S, Xu, C, Li, W, Lin, W, Keum, YS, Han, J, Gallo, MA & Kong, AN (2004) Regulation of Nrf2 transactivation domain activity. The differential effects of mitogen-activated protein kinase cascades and synergistic stimulatory effect of Raf and CREB-binding protein. J Biol Chem 279, 2305223060.CrossRefGoogle ScholarPubMed
241Kang, KW, Lee, SJ, Park, JW & Kim, SG (2002) Phosphatidylinositol 3-kinase regulates nuclear translocation of NF-E2-related factor 2 through actin rearrangement in response to oxidative stress. Mol Pharmacol 62, 10011010.CrossRefGoogle ScholarPubMed
242Nakaso, K, Yano, H, Fukuhara, Y, Takeshima, T, Wada-Isoe, K & Nakashima, K (2003) PI3K is a key molecule in the Nrf2-mediated regulation of antioxidative proteins by hemin in human neuroblastoma cells. FEBS Lett 546, 181184.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1 Overview of MAP kinase and Akt/PKB signalling cascades in neurons. Flavonoid-induced activation of ERK1/2 or PI3K/Akt pathways acts to stimulate neuronal survival and/or enhance synaptic plasticity and long-term potentiation relevant to the laying down of memory. In addition, inhibitory actions within JNK and p38 pathways are likely to be neuroprotective in the presence of stress.

Figure 1

Fig. 2 Formation of stable long-term potentiation at synapses. (1) Increased expression and release of BDNF from the synapse through enhanced CREB activation. BDNF binds to pre- and post-synaptic TrkB receptors (2), triggering glutamate release and PI3K/mTOR signalling and Arc synthesis (3). Sustained activation of mTOR leads to enhanced translational efficiency whilst Arc, in association with Cofilin, triggers F-actin expansion and synapse growth (mushroom synapse) (4).

Figure 2

Fig. 3 Involvement of activated glial cells in neuroinflammation-induced neurodegeneration. Central to glial-induced neurotoxicity is the generation of NO via increases in the expression of iNOS. iNOS itself is induced by the cell surface CD23 receptor which is in turn activated by cytokines such as TNF-α and IL-1γ. NO may diffuse to neighbouring neurons where it inhibits mitochondrial respiration at cytochrome c oxidase. NO may also react with O _{2}^{ - } to generate ONOO−  which can cause damage to proteins, inhibit mitochondrial respiration and activate cell death genes and signalling pathways which may ultimately lead to neuronal death. Furthermore, cytokines such as TNF-α may cause direct cell death by binding to specific receptors expressed in neurons and subsequently activate genes that trigger the apoptotic pathway.

Figure 3

Fig. 4 Structural homology of flavonoids with specific pathway inhibitors. Use of specific MAPK inhibitors such as SB203580 and PD98059 inhibit the transcriptional regulation of iNOS in activated glial cells. Interestingly, the structure of PD98059 and other kinase inhibitors have close structural homology to that of flavonoids. It is therefore possible that flavonoids may modulate neuroinflammation by interfering with cell signalling pathways such as MAPK.

Figure 4

Fig. 5 Potential points of action of flavonoids within MAPK signalling cascades in neurons. Activation of ERK1/2 or ERK5 are generally pro-survival, whilst inhibitory actions on JNK and p38 pathways are also likely to be neuroprotective.

Figure 5

Fig. 6 Potential points of action of flavonoids within PI3K/Akt signalling pathway. Active PI3K catalyzes the production of phosphatidylinositol-3,4,5-triphosphate (PIP3) which activates phosphoinositide-dependent protein kinase 1 and 2 (PDK1 and PDK2) and Akt. Through its effects on these kinases, PI3K is involved in the regulation of a wide variety of processes, including cell growth, cell proliferation, differentiation, cell cycle entry, cell migration and apoptosis. Flavonoids have been proposed to act on this pathway via direct modulation of PI3K activity via binding to its ATP binding pocket, in a similar manner to that of LY294002. Alternatively, they may act to modulate the activity of the tumour suppressor, PTEN.

Figure 6

Fig. 7 Involvement of MAP kinase and PI3 kinase signalling in regulation of the Keap1-Nrf2 pathway. Inactive Nrf2 is retained in the cytosol by association a complex with the cytoskeletal protein Keap1. Cytosolic Nrf2 may be phosphorylated in response to MAP kinase, PI3 kinase and protein kinase C pathways. Following phosphorylation, Nrf2 translocates to the nucleus, where it activates gene expression through binding to the ARE, following its interaction with other transcription factors in the bZIP family (CREB, ATF-4 and fos or c-Jun). Nrf2 activation of genes is opposed by small maf proteins, including MafG and MafK, maintaining a counterbalance to Nrf2 and balancing the oxidation level of the intracellular environment.