Next Article in Journal
Prognosis of Dermatitis Herpetiformis Patients with and without Villous Atrophy at Diagnosis
Next Article in Special Issue
Changes of Colonic Bacterial Composition in Parkinson’s Disease and Other Neurodegenerative Diseases
Previous Article in Journal
Associations of Maternal Vitamin D Deficiency with Pregnancy and Neonatal Complications in Developing Countries: A Systematic Review
Previous Article in Special Issue
Towards an Integrative Understanding of tRNA Aminoacylation–Diet–Host–Gut Microbiome Interactions in Neurodegeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polyphenols in Parkinson’s Disease: A Systematic Review of In Vivo Studies

by
Małgorzata Kujawska
* and
Jadwiga Jodynis-Liebert
Department of Toxicology, Poznan University of Medical Sciences, 30 Dojazd Str., 60-631 Poznań, Poland
*
Author to whom correspondence should be addressed.
Nutrients 2018, 10(5), 642; https://doi.org/10.3390/nu10050642
Submission received: 21 April 2018 / Revised: 14 May 2018 / Accepted: 16 May 2018 / Published: 19 May 2018
(This article belongs to the Special Issue Brain Aging and Gut-Brain Axis)

Abstract

:
Parkinson’s disease (PD) is the second most common neurodegenerative disorder. However, therapeutic options treating only its symptoms are very disappointing. Therefore there is an ongoing search for compounds capable of tackling the multi-dimensional features of PD. Recently natural polyphenols have gained great interest as potential therapeutic agents. Herein, we have attempted to summarize results obtained in different animal models demonstrating their neuroprotective effects. The in vivo findings presented below are supported by human subject data and reports regarding the ability of polyphenols to cross the blood-brain barrier. The beneficial effects of polyphenols are demonstrated by the results of behavioral examinations, mainly related to motor and cognitive capabilities, histopathological and immunohistochemical examination concerning the protection of dopaminergic neurons, analyses of dopamine and the concentration of its metabolites, as well as mechanistic studies regarding the modulation of oxidative stress, neuroinflammation, cellular iron management, proteinopathy, and additionally the regulation of signaling pathways. Importantly, data about brain distribution of the metabolic derivatives of the reviewed polyphenols are crucial for the justification of their nutritional intake in neuroprotective intervention, as well as for the identification of potential targets for a novel therapeutic approach to Parkinson’s disease.

1. Introduction

Parkinson’s disease (PD) is, after Alzheimer’s disease, the second most common human neurodegenerative disorder, which is characterized by motor and cognitive dysfunction associated with the degeneration or loss of dopaminergic neurons in the substantia nigra pars compacta (SNpc). The hallmark lesions called Lewy bodies are produced by the progressive accumulation of protein inclusions containing α-synuclein and ubiquitin in the cytoplasm of selected neurons, which leads to their death by necrosis and/or apoptosis. Despite intensive research, the etiology of PD remains poorly understood. Nevertheless, several underlying pathophysiological mechanisms such as oxidative stress, neuroinflammation, iron dysregulation, mitochondrial dysfunction, excitotoxicity, loss of neurotrophic factors, glial activation, and endoplasmic reticulum stress as well as protein misfolding and dysfunction in their degradation have been credited as significant pathways for the development of therapeutic approaches [1,2]. These complexities of Parkinson’s disease are accompanied by a lack of sufficient therapies, which are limited to long-term symptomatic treatment. In this respect, the concept of neuroprotection referring to the prevention of DA cell death and, hence, to retarding or halting disease progression appears to be a promising strategy. There is an ongoing search for substances targeting the underlying neurodegenerative process. Various substances exhibiting anti-inflammatory, antioxidant, and metal chelating activity in the central nervous system (CNS) have been tested for use to facilitate the management of PD. In this context, natural polyphenols have raised much attention in the recent decade. It has been established that they may exert neuroprotective effects by targeting multiple mechanisms, beyond those above mentioned. Importantly, polyphenols have been reported to inhibit the formation of α-synuclein misfolded aggregates and to reduce mitochondrial dysfunction-induced oxidative stress and inflammatory responses. Moreover, they are capable of affecting cell survival and cell cycle genes as well as signaling pathways. Polyphenolic compounds can activate extracellular signal-regulated kinase (ERK), phosphoinositide 3-kinase (PI3K)/Akt, mitogen-activated protein kinase (MAPK), serine/threonine protein kinase (AKT), protein kinase C (PKC) or inhibit NF-κB pathways. Finally, the antioxidant activity of polyphenols contributes to the activation of the nuclear factor-erythroid 2-related factor 2 (Nrf2)/antioxidant responsive element (ARE) [3,4,5]. These mechanisms have been extensively investigated in in vitro models [6,7,8,9,10,11,12]. However, the findings of the above cited studies cannot be directly translated into in vivo conditions since the action of polyphenols will depend on their biotransformation and the ability to penetrate the blood-brain barrier.
The number of studies concerning neuroprotective properties of polyphenols in animal models is markedly lower than those performed in vitro. Epidemiological and clinical data regarding the relation between polyphenol consumption and the incidence of neurodegenerative diseases, such as PD, are also limited.
The aim of our study was to review the investigation of neuroprotective properties of polyphenols in animal models of Parkinson’s disease, which were supported by human subject data as well as the current knowledge regarding the ability of polyphenols and/or their metabolites to cross the blood-brain barrier. This review does not cover other neurodegenerative disorders.

2. Studies in Animal Models

Around 95% of diagnosed Parkinson’s disease cases are sporadic, caused by a combination of environmental exposures and genetic susceptibility as well as aging. In parallel with advances in our understanding of the disease pathogenesis, various animal models of PD have been established. The majority of them are based on environmental contribution to the pathological hallmarks of Parkinson’s disease, including loss of dopaminergic neurons in the SNpc, reduced dopamine (DA) content in the striatum, neuroinflammation and aggregation pathology reminiscent of Lewy bodies as well as motor symptoms. These parkinsonian-like features are developed in animals by the use of neurotoxins targeting the mitochondrial oxidative phosphorylation such as 6-hydroxydopamine (6-OHDA), N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), rotenone (ROT) and homocysteine (Hcy). In addition, lipopolysaccharide (LPS), a potent microglial activator, is used to induce inflammatory dopaminergic neurodegeneration. These models facilitate an integrated view of a neuroprotective mechanism in regard to functional improvements with underlying changes occurring in molecular pathways, categories of cells, and brain networks. However, to address a specific molecular target of neuronal degeneration of PD, various genetic models are employed. Since the pathology of PD is multifaceted and several pathomechanisms are involved in the dopaminergic neuronal death, no single model is currently able to recapitulate Parkinson’s disease as seen in humans [1,13].

2.1. Pure Compounds

There is an ongoing search for novel therapeutic candidates showing promise for tackling the multi-dimensional features of Parkinson’s disease. Recently, a large number of individual polyphenols have gained considerable interest for their potential in developing alternative drugs or their beneficial effect from nutritional intake [1]. The following sections review the neuroprotective effects of pure polyphenolic compounds. To systematize the great number of reports we have presented them in the order of various PD models. All studies discussed below are summarized in Table 1.

2.1.1. 6−hydroxydopamine Model

The unilateral injection of 6-hydroxydopamine (6-OHDA) into the medial forebrain bundle of rodents has been commonly applied as an experimental model of PD. 6-OHDA generates oxidative stress and induces a lesion of nigrostriatal dopaminergic neurons that results in their progressive loss and consistent behavioral phenotype [81].
Flavonoids represent the most abundant group of polyphenols with well-established anti-parkinsonian effect [82]. Administration of citrus flavanone naringenin to rodents for four days, before lesioning with 6-OHDA resulted in a significant decrease in the loss of tyrosine hydroxylase (TH)—positive cells as well as in the loss of DA and its metabolites 3,4-dihydroxyphenylacetic acid (DOPAC) and homovanillic acid (HVA) in the brain [14,15]. As tyrosine hydroxylase catalyzes the formation of l-3,4-dihydroxyphenylalanine (l-DOPA), which is the rate-limiting step in the biosynthesis of dopamine, thus TH expression is one of the most frequently used markers for identifying DA neurons [83]. Lou et al. [15] studied cellular mechanisms underlying the neuroprotective effects of naringenin, and they have demonstrated that it readily activated Keap1/Nrf2/ARE axis, thereby inhibiting oxidative stress in 6-OHDA-lesioned mice. The pretreatment with glycoside of naringenin − naringin, also has been shown to provide neuroprotection in the experimental model of PD [16]. It protected DA neurons in the mouse brain by reducing neuroinflammation and microglial activation as well as by increasing mammalian target of rapamycin complex 1 (mTORC1) activity [16]. Another flavanone glycoside found in citrus fruits − hesperidin and its aglycone hesperetin were reported to exert neuroprotective effect [17,18]. In the striatum of aged mice, hesperidin attenuated the 6-OHDA-induced reduction in glutathione peroxidase (GPx) and catalase (CAT) activity and total reactive antioxidant potential, leading to increased levels of striatal dopamine and its metabolites and preventing memory impairment and depressive-like behavior [17]. The beneficial effect of hesperetin against 6-OHDA-induced behavioral abnormalities has been demonstrated to be associated with the attenuation of apoptosis, astrogliosis and oxidative stress [18]. However, the neuroprotective effect was not observed after treatment with flavonol quercetin [14]. The lack of quercetin action is in accordance with the results of Kääriäinen et al. [19], who reported that i.p. administration of quercetin for 14 days to rats lesioned with 6-OHDA did not afford protection in TH-positive nigral cell number, striatal fiber density or levels of dopamine. On the other hand, Haleagrahara et al. [20] in the same PD rat model observed increased striatal dopamine and neuronal survivability, which were accompanied by a decreased level of oxidative damage to proteins and lipids and increased content of reduced glutathione (GSH), in rats treated with quercetin, suggesting that it could provide neuroprotection through antioxidant effects. The reported different responses to quercetin could be related to the magnitude of the lesion, which was probably different since Zbarsky et al. [14] and Kääriäinen et al. [19] injected 6-OHDA into the medial forebrain bundle, whereas Haleagrahara et al. [20] did it by an intracisternal puncture. The glycoside of quercetin − rutin, has also been shown to protect against 6-OHDA-induced neurotoxicity in rats via its antioxidant as well as anti-inflammatory activity [21]. Rutin supplementation significantly reduced oxidative damages to membrane lipids and protein levels, and led to an increase in GSH content and the activity of antioxidant enzymes in the striatum of 6-OHDA infused rats. In addition it protected these rats from the rise in circulating tumor necrosis factor alpha (TNF-α) and interleukin 1 beta (IL-1β) production. These effects correlated well with: a decrease in the deficits in locomotor activity and motor coordination, an increase in DA and its metabolite content and an increase in the number of dopaminergic D2 receptors, in lesioned rats pre-treated with rutin [21]. Similar neuroprotective effects of troxerutin, a derivative of rutin were reported [22]. Pretreatment of 6-OHDA lesioned rats with troxerutin improved their motor functions, prevented the loss of nigral TH-positive neurons and decreased the striatal level of lipid peroxidation and ROS. Moreover, the flavonol effectively inhibited astrogliosis and apoptosis. The authors suggested that the beneficial effects of troxerutin could partly be mediated through the PI3K/Akt pathway [22].
Myricitrin, derived from the root bark of Myrica cerifera, and its aglycon, flavonol myricetin have also been shown to exert neuroprotective effects. In mice receiving an intrastriatal injection of 6-OHDA, myricitrin improved motor function and protected DA neurons. The beneficial effects were associated with the activation of mTORC1, maintenance of TH activity as well as the anti-neuroinflammatory action [84]. Myricetin prevented the decrease of dopamine and degeneration of TH-positive neurons induced by 6-OHDA in rats. Its protective effects were partially attributed to the suppressed iron toxicity, as was shown by the decreased number of iron-staining cells [23].
The same model of the nigrostriatal lesion was applied to examine the neuroprotective effects of isoflavone genistein given to rats intraperitoneally, 1 h before surgery. Histological study revealed that the compound tested protected nigrostriatal neurons against degenerative effects induced by 6-OHDA. It was supported by the attenuation of rotational behavior induced by apomorphine, the test for nigrostriatal dopamine depletion [24]. Similarly, another isoflavonoid puerarin was demonstrated to reduce the 6-OHDA induced decrease in TH-positive cell counts and to prevent apoptosis. Immunohistochemical analysis proved that puerarin decreased the level of Bax stimulated by 6-OHDA injection. Moreover, puerarin was capable of restoring the content of dopamine and its metabolites, and to enhance the expression level of a glial cell-derived neurotrophic factor (GDNF), which is involved in the survival of adult DA neurons in the striatum [25]. The authors suggested that the latter effect, as well as the inhibition of apoptotic signaling pathways, are the essential events in neuroprotection exerted by puerarin.
Oral administration of flavone found in Scutellaria baicalensis, baicalein for one week, re-balanced glutamate (GLu) and gamma-aminobutyric acid (GABA) levels in basal ganglia of 6-hydroxydopamine treated rats resulting in the attenuation of muscle tremor. This effect was caused by regulating the expression of gamma-aminobutyric acid transaminase (GABA-T) and glutamine synthetase (GS)—enzymes involved in the turnover of GLu and GABA as well as by inactivating the neuronal excitation induced by GLu [26].
Administration of tangeretin, a citrus flavanone, to rats p.o. for four days before 6-OHDA lesioning resulted in a significant reduction of the loss in both TH- positive cells and striatal dopamine content [27].
Epigallocatechin-3-gallate (EGCG), a major flavanol in tea, has been widely shown to exert neuroprotective effects in various PD models [85]. However, in 6-OHDA-lesioned rats, only a slight improvement in postural and motor behavior has been observed after oral 14-day pre-treatment with this representative of green tea polyphenols [28].
Intake of the anthocyanin pelargonidin has been reported to dose-dependently attenuate behavioral and structural abnormalities, due to 6-OHDA toxicity and to decrease lipid peroxidation [29].
Some stilbenes were also demonstrated to protect against 6-OHDA-induced neurotoxicity. Resveratrol (RES) given orally for ten weeks reduced neural inflammation in a 6-OHDA-induced PD in rats, as was evidenced by suppression in the production of cyclooxygenase-2 (COX-2) and TNF-α in substantia nigra as well as ameliorating 6-OHDA-induced neurobehavioral deficit [30]. These findings are in agreement with data reported by Wang et al. [31]. They demonstrated that resveratrol administered orally for 14 days decreased the abnormal rotational behavior, the loss and apoptosis of nigral cells, and ROS level, while improving the total antioxidant capacity of nigral tissues significantly in rats injected with of 6-OHDA. In another study, pretreatment with RES for 15 days led to upregulation of the antioxidant status and lowering the dopamine loss in 6-OHDA lesioned rats. That was associated with the attenuation of elevated levels of thiobarbituric acid reactive substances (TBARS) and protein carbonyls (PCs) as well as the activity of phospholipase A2 [32]. Co-treatment with another stilbenoid, piceid, was also found to attenuate 6-OHDA-induced motor defects in rats. Furthermore, the lipid peroxidation level was significantly decreased, and the activity of superoxide dismutase (SOD) was increased after co-administration of piceid. The findings support the tight correlation between the antioxidant activity of polyphenols and their protective effects on motor function [33].
An increasingly popular polyphenol tested for its neuroprotective properties is curcumin (CUR). Long-term pretreatment with this compound reversed the 6-OHDA-induced decrease in the levels of DA and its metabolite DOPAC in the striatum. Moreover, curcumin restored the number of TH-positive neurons and decreased the number of iron-positive cells, as compared with the 6-OHDA treated group [34]. Tripanichkul et al. [35] have shown that the same dose of CUR administered for seven days, starting after 6-OHDA injection, attenuated loss of striatal DA axons triggered by the neurotoxin. Inhibition of microglial and astroglial reaction, as evidenced by reduced glial fibrillary acidic protein (GFAP) and ionized calcium binding adapter molecule 1 (Iba1) immunoreactivity, as well as maintenance of the superoxide dismutase 1 (SOD1) level has been suggested to be one of the mechanisms underlying the neuroprotective effects of curcumin [35].
In rats treated with carnosic acid (CA), a phenolic diterpene found in rosemary, for three weeks before 6-OHDA injection, improvement of the locomotor activity and reduction of apomorphine-caused rotations have been observed. Moreover, pretreatment with CA provided protection against lipid peroxidation and GSH depletion in the substantia nigra. Those effects were observed concurrently with the increase in the protein expression of c-glutamate-cysteine ligase, superoxide dismutase, and glutathione reductase in the striatum. Finally, CA has been demonstrated to reverse the 6-OHDA-induced activation of c-Jun NH2-terminal kinase and p38, the down-regulation of the Bcl-2/Bax ratio, the up-regulation of cleaved caspase-3/caspase-3 and cleaved PARP/PARP ratios, and the down-regulation of tyrosine hydroxylase protein in the striatum of rats. These results suggest that the CA protective effect against 6-OHDA-induced neurotoxicity is attributable to its anti-apoptotic and anti-oxidative action [36].
Zare et al. [37] demonstrated the neuroprotective potential of sinapic acid, hydroxycinnamic acid widespread in plants, given to rats p.o. before injection of 6-OHDA as evidenced by the improved turning behavior, reduced loss of dopaminergic neurons in the substantia nigra, lowered iron reactivity, and decreased the level of MDA and nitrite [37]. Ellagic acid, another natural polyphenol antioxidant found in numerous fruits and vegetables, administered by gavage caused an improvement in locomotion and a reduction in the levels of neuroinflammatory biomarkers, TNF-α and IL-1, in the striatum and hippocampus of rats treated with 6-OHDA [38].
Studies on cannabinoids representative, delta 9-tetrahydrocannabinol, revealed that daily administration (i.p.) of the compound for two weeks to rats produced a significant attenuation in the magnitude of the following changes evoked by 6-OHDA injection: a depletion of dopamine content and a reduction in tyrosine hydroxylase activity in striatum, accompanied by a reduction in tyrosine hydroxylase mRNA levels in the substantia nigra [39].

2.1.2. N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine Model

Another toxin employed as an experimental model to study therapeutic strategy against PD is N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), which induces dopaminergic degeneration via oxidative stress [81]. The active metabolite of MPTP − MPP+, is selectively taken up by dopaminergic neurons and actively transported into mitochondria where it interferes with the respiratory chain and inhibits complex I, inducing energy depletion and ROS production. The MPTP-induced neurotoxicity is limited to the nigrostriatal dopaminergic pathway and is associated with motor dysfunctions [13].
This model was applied broadly to examine the neuroprotective properties of EGCG, one of the main active components of green tea [40,41,42,43,44,45,46]. Treatment with EGCG has been demonstrated to prevent MPP+-induced motor impairment [43,46] and dopamine neuron loss [40,41,42,45,46], and depletion of DA [40,41,42,43,44] in the substantia nigra. Moreover, EGCG has been shown to increase expression and activity of tyrosine hydroxylase [40,41]. Mandel et al. [41] demonstrated that EGCG prevented an MPTP-induced increase in the number of synuclein-positive neurons in the substantia nigra, upregulation of Bax and the depletion of protein kinase C (PKCα) in the striatum. It has also been suggested that the decrease in the expression of neuronal [42] and inducible nitric oxide synthase [45] in the SNpc and striatum in MPTP-treated mice has contributed significantly to the neuroprotective effect of EGCG. Another study showed that EGCG caused restoration of the MPTP-reduced level of the survival factor Ras [44]. Nigral iron accumulation has been demonstrated in animals treated with MPTP that may be associated with the altered expression of iron-related proteins, such as the increased expression of iron importer divalent metal transporter 1 (DMT1), or the decreased expression of the iron exporter ferroportin [86,87,88]. EGCG post-treatment for seven days has been reported to regulate the iron-export protein ferroportin in the SNpc and to reduce oxidative stress in mice. In addition, EGCG restored MPTP-induced functional and neurochemical deficits (DA and DOPAC) to a value similar to the control group [43]. Besides the influence of EGCG on the brain Zhou et al. [46] have demonstrated that its protective effect against dopaminergic neurons loss is also associated with modulation of peripheral inflammation. The authors have reported that EGCG treatment caused a decrease in the ratio of CD3+CD4+ to CD3+CD8+ T cells in peripheral blood and in concentrations of serum TNF-α and IL-6 of MPTP-treated mice [46].
In MPTP-treated mice, co-administration with natural flavone acacetin for three days (p.o.) caused a decrease in time of turning, and locomotor activity as compared to those in MPTP-only-treated mice. In the acacetin co-treated group, the inhibition of MPTP-induced degeneration of DA neurons and depletion of dopamine level in the substantia nigra and striatum of the brain were observed. Moreover, administration of acacetin inhibited microglial activation accompanied by decreased production of inducible nitric oxide synthase (iNOS) and COX-2 [47]. Baicalein administration improved the behavioral performance in the motor balance test in mice injected with MPTP. The effect was associated with the attenuation of the MPTP upregulated upsurge of glutamatergic strength that might result from the suppression of cytokine production, including IL-1β and TNF-α in the substantia nigra and striatum [48]. Similarly, in the study by Lee et al. [49], baicalein was shown to improve motor ability, prevent dopaminergic neuron loss and reduce microglial and astrocyte activations caused by MPTP. Two-week administration of 7,8-dihydroxyflavone to mice (i.p.) attenuated the neurotoxic effects of MPTP as evidenced by the preservation of tyrosine hydroxylase expression and the reduction of activated caspase-3. The compound tested has been recognized as a high-affinity agonist of tyrosine kinase receptor B that provokes downstream signaling similar to the action of a brain-derived neurotrophic factor, which mediates neuronal survival, synaptic plasticity, and neurogenesis [50]. Takano et al. [51] reported a significant reduction in the loss of both TH- positive cells and striatal dopamine content in mice given another citrus flavonone, tangeretin i.p., for four consecutive days post MPTP injection. Administration of nobiletin, a polymethoxylated flavonoid isolated from citrus fruit peel, improved MPTP-induced motor dysfunction and cognitive deficits in C57/BL6N mice. The effect was correlated with enhanced activities of striatal and hippocampal Ca2+/calmodulin-dependent protein kinase II (CaMKII) and cAMP-dependent protein kinase (PKA), which are critical for TH activity [53]. Jeong et al. [52] have demonstrated that nobiletin protected DA neurons and attenuated interleukin-1b production by inhibiting microglial activation and preserving GDNF expression in the SNpc. Similarly, it has been found that intraperitoneal injection of a flavanone glycoside naringin, prevented nigrostriatal DA neurons against the MPP+-induced degeneration through the induction of glial-derived neurotrophic factor (GDNF). Moreover, the naringin attenuated level of TNF-α in microglia increased after exposure to MPP+ [54].
Flavonol kaempferol has been found to improve motor coordination, to raise striatal dopamine and its metabolites levels, to increase SOD and GPx activity, and to reduce the content of MDA compared with mice treated with MPTP alone. Immunohistochemical studies showed that treatment with kaempferol could prevent the loss of TH-positive neurons induced by MPTP [55]. Similarly, another flavonol, quercetin, protected against oxidative stress by increasing SOD and GPx activity and against dopamine depletion as well as improved motor balance and coordination. The compound also counteracted inhibition of acetylcholinesterase (AChE) activity and maintained the resting membrane potential of neurons in MPTP-treated mice [56]. Neuroprotective activity of morin, a flavonol originally isolated from the Moraceae family, was investigated in mice administered with the compound i.p. for ten days concomitantly with MPTP given for the last five days. Morin significantly attenuated striatal dopamine depletion and a loss of TH-positive neurons caused by MPTP treatment. Moreover, as evidenced by the measurement of cataleptic time (bar test) and the number of steps (drag test), behavioral deficits induced by neurotoxin were also reduced [57].
A similar protective effect on the motor deficit and dopaminergic neuronal loss caused by MPTP has been reported in mice co-treated with silibinin (SIL), a flavonolignane derived from Silybum marianum [58]. The authors concluded that the neuroprotective effects of SIL were mediated by the stabilization of neuronal mitochondrial transmembrane potentials (MMPs) [58]. In another study, treatment with silibinin ameliorated in a dose-dependent manner the MPP+-induced neurotoxicity in the SNpc of rats, as was clearly shown by an increase in TH-positive cell number. In addition, treatment with SIL attenuated the MPP+-induced levels of inflammatory molecules such as TNF-α, IL-1β, and iNOS [59]. Co-treatment with SIL for seven days significantly protected against the progression of MPP+-induced striatal dopaminergic neurotoxicity in rats as evidenced by attenuation of motor deficits. This action was mediated through the increase in DA level in the striatum. SIL treatment improved striatal mitochondrial function and integrity as well as the enhanced activity of SOD. Anti-apoptotic and anti-inflammatory properties of SIL, as demonstrated by a decrease in expression of caspase-3 and NFκB protein, might also contribute to its neuroprotective activity [60].
Genistein pretreatment reduced neurotoxicity in MPTP-challenged mice, which was attributed to increased contents of DA, DOPAC, and HVA in striatum and number of TH-positive neurons in the SNpc. This phytoestrogen restored the decrease in MPTP-induced Bcl-2 gene expression in mice midbrain and therefore positively regulated cell survival [61].
Gastrodin, the phenolic glucoside from Gastrodia elata, showed neuroprotective effects in the subchronic MPTP mouse model of PD. It was evidenced by ameliorating bradykinesia and motor impairment as well as by its beneficial effects against dopamine depletion and by a reduction in reactive astrogliosis caused by MPTP in the substantia nigra and striatum of mice. Moreover, gastrodin efficiently prevented neuronal apoptosis through the reduction in the protein expression of proapoptotic Bax and caspase-3 protein and the increase in the antiapoptotic Bcl-2 expression. As a consequence, a MPP+-induced activation of the downstream target of caspase-3-poly (ADP-ribose) polymerase (PARP) contributing to cell death, was substantially suppressed [62].
Phenanthrenequinone tanshinone I, a constituent of Salviae miltiorrhizae, also improved motor functions, normalized striatal level of DA, and provided dopaminergic neuronal protection in mice receiving four i.p. injections of MPTP-HCl, as evidenced by the increased number of TH-positive neurons. The antiparkinsonian effect of tanshinone I was mediated by the modulation of microglial activation leading to the attenuation of the increase in TNF-α and IL-10 concentrations [63].
Ferulic acid (FA), hydroxycinnamic acid abundant in plants, administered for seven days to MPTP injected mice attenuated the formation of MPP+ and altered Bax/Bcl-2 ratio exerting an antiapoptotic effect in the substantia nigra region. The authors of the report suggested that both the antioxidant and antiapoptotic effects of FA contributed to the improvement of motor balance and coordination in mice challenged with MPTP [64].
Madecassoside, a saponin isolated from Centella asiatica, administered intragastrically for seven days, was effective in recovering MPTP-induced early signs of parkinsonism, as evidenced by improved locomotor dysfunction as well as the protection of dopaminergic neurons and dopamine loss in the striatum. The observed effects were accompanied by a decrease in MDA content and an increase in the GSH level, Bcl-2/Bax ratio, and protein expression of brain-derived neurotrophic factor (BDNF) [65].
Curcumin was demonstrated to attenuate MPTP-induced toxicity when administered to mice i.p. immediately post MPTP injection [66] or for seven days [67]. MPTP-induced depletion of striatal dopamine was markedly limited [66,67], the density of dopaminergic neurons was increased, and the MAO-B activity was inhibited [66] as a result of curcumin treatment. Recently, it has been reported that pre-treatment with curcumin derivatives significantly protected against MPTP-induced dopaminergic neurodegeneration, as assessed by attenuation of striatal DA and DA turnover as well as by a significant decrease in the loss of TH immunoreactivity [68]. Moreover, the authors have revealed that MPTP-induced inflammatory responses, as manifested by stimulated GFAP protein expression and elevated pro-inflammatory cytokines in the striatum of mice brains, were markedly suppressed by curcuminoids treatment.
Ginsenoside Rg1, a triterpenoid saponin found in ginseng species, injected to mice i.p. for eight days reduced the loss of dopaminergic neurons caused by MPTP, prevented glutathione depletion, and caused an increase in SOD activity in the substantia nigra region. As the JNK signaling cascade contributes to dopaminergic neuron apoptosis, the observed attenuation of the phosphorylation of JNK and c-Jun in the substantia nigra, as well as antioxidant activity of ginsenoside, play a significant role in its neuroprotective effects [69].
Daily oral treatment of mice with piceid, for 14 days starting from the first day of subcutaneous injection of MPTP, significantly prevented motor deficits as well as preserved SOD activity and reduced MDA level in the striatal region. Furthermore, piceid protected TH positive DA neurons and activated survival signaling (e.g., increased p-Akt expression and reduced activated caspase-3) [33].
In another model of PD, MPTP is given with probenecid (p) that blocks the renal clearance of MPTP to prolong its action [89]. Several studies using this model were performed for evaluating the antiparkinsonian effects of theaflavin (TF), a black tea polyphenol. Oral pretreatment of MPTP/p administered mice with TF for 35 days caused normalization of their motor function. The effect was accompanied by attenuation of neurodegeneration and apoptosis, as evidenced by an elevated expression of nigral TH and dopamine transporter (DAT) as well as the reduced level of apoptotic markers such as caspase-3, 8, and 9 [70]. Moreover, TF has been found to increase striatal GSH level and GPx activity in MPTP-treated mice [71]. The neuroprotective effect of theaflavin has also been demonstrated to depend on an inhibition of the release of pro-inflammatory cytokines as well as on a decrease in the expression of proapoptotic Bax and an increase in the expression of antiapoptotic protein Bcl-2 in the striatum, which processes counteract the initiation of apoptotic events [72].
Flavones apigenin and luteolin given orally for 26 days, including five days of pretreatment, have been demonstrated to exert multifunctional antiparkinsonian effects like the protection of tyrosine hydroxylase expression, alleviation of oxidative stress and inflammation, enhancement of neurotrophic activity leading to improved motor coordination and locomotor behavior in an MPTP/p mouse model. Additionally, treatment with the compounds tested caused a decrease in astrocyte activation, reflected by the reduced expression of the GFAP in the SNpc of the brain. The authors suggested that the neuroprotective effect of both flavonoids has been attributed to the enhancement of neurotrophic potential as evidenced by the increased level of the brain-derived neurotrophic factor (BDNF) in the substantia nigra region [73].
Syringic acid (SA) is one of the major naturally occurring benzoic acids exhibiting strong antioxidant activity. Oral pre-treatment of mice with SA (1 h before each MPTP injection) for 35 days was found to improve the MPTP/p-induced impaired motor functions by restoring the dopamine content as well as to enhance the antioxidant defense system in the brain. In addition, SA showed the ability to ameliorate MPTP/p-induced decrease in expressions of TH, DAT, and vesicular monoamine transporter-2 (VMAT2) and to attenuate the production of pro-inflammatory cytokines, such as TNF-α, IL-1β, and COX-2 [74].

2.1.3. Rotenone Model

Rotenone (ROT), another inhibitor of respiratory complex I, is also used to induce dopaminergic neuronal oxidative damage in experimental animal models of PD. Rats exposed to prolonged, low-dose rotenone treatment develop a PD-like phenotype characterized by motor dysfunction, a loss of dopaminergic neurons, the formation of Lewy-like inclusions, and microglial activation [90].
In a rotenone model of PD, the application of quercetin to rats for four days rescued motor dysfunction by enhancing dopamine release in the striatum and prevented the loss of nigral TH-positive neurons. In addition, this flavonol attenuated the ROT-induced loss of mitochondrial complex-I activity and GSH depletion as well as blocking programmed cell death in nigral neurons [75].
Curcumin administered for 50 days significantly alleviated motor dysfunction and increased TH activity in the SNpc of ROT-injured rats. The biochemical assessment indicated that in rats pretreated with curcumin an elevated level of GSH and a decreased content of reactive oxygen species and malondialdehyde were observed. Mechanistic studies demonstrated that curcumin was able to restore levels of heme oxygenase-1 (HO-1) and NAD(P)H: quinone oxidoreductase 1 (NQO1) expression and to activate Akt/Nrf2 signaling pathway, thus ameliorating ROT-induced damage in the brain [76].
In rats given rotenone subcutaneously for five weeks, co-administration (p.o.) of piceid attenuated rotenone-induced motor defects in a dose-dependent manner. The dose of 80 mg/kg/day caused an even better effect than l-levodopa (l-dopa)—the gold standard for treating Parkinson’s disease. Treatment with piceid prevented the rotenone-induced changes in the levels of GSH, thioredoxin, ATP, MDA, and SOD activity in the striatum. Furthermore, treatment with the stilbenoid rescued rotenone-induced dopaminergic neurodegeneration in the substantia nigra region [33].
Ferulic acid co-administered to ROT-treated rats for four weeks rescued DA neurons in substantia nigra area and nerve terminals in the striatum from rotenone damage. FA restored the activity of antioxidant enzymes and GSH content while inhibited lipid peroxidation. Following treatment with FA, the production of inflammatory mediators such as COX-2 and inducible NOS and proinflammatory cytokines was also reduced. In addition, these effects were supported by a reduction in expression of Iba1 and GFAP suggesting attenuation of microglial and astrocytic activation. Hence, it could be concluded that protective effects of FA were mediated through its antioxidant and anti-inflammatory properties [77].
Another study has demonstrated that sesamol, a phenolic compound found in sesame seeds, and naringenin, given orally for ten consecutive days, had protective effects in rotenone-lesioned rats. It was evidenced by improved motor skills and enhanced expression of neuroprotective proteins, including parkin, PARK 7 protein (DJ1), C terminus Hsp70 interacting protein (CHIP), and TH in the striatum and substantia nigra. These results correlated well with the reduction in caspase-9 and 3 and ubiquitin levels. Moreover, in the same rats, improved morphology and survivability of neurons were noticed [78].

2.1.4. Lipopolysaccharide Model

Another toxin-induced PD model consists in the treatment of rodents with lipopolysaccharide (LPS), a potent microglial activator, which is able to induce toxicity to dopaminergic neurons. Pretreatment of rats with EGCG 24 h before LPS intraperitoneal injections for seven days was found to have an anti-inflammatory effect through reducing the level of TNF-α and NO as well as to preserve the DA level in the midbrain [79].

2.1.5. Homocysteine Model

In epidemiological and longitudinal studies, a tight correlation between homocysteine (Hcy) and cognitive impairment due to apoptosis, neuronal death, oxidative stress, overactivation of glutamate receptors, mitochondrial dysfunctions, and activation of caspases for all of the neurodegenerative diseases has been demonstrated [80]. Intranigral infusion of homocysteine in rats is used to inhibit mitochondrial complex-I activity and to reduce striatal dopamine in order to induce neurodegeneration in SNpc and impaired motor activity [91]. Mansouri et al. [80] have investigated the neuroprotective effects of curcumin against Hcy-induced neurotoxicity in substantia nigra in the rats. They have found that administration (i.p.) of curcumin for ten days, beginning five days prior to homocysteine intracerebroventricular injection caused the improvement of locomotor function. The effect was accompanied by a reduction in HCy-induced apoptosis and an increase in Bcl-2 as well as a decrease in Bax proteins expression [80].

2.2. Beverages and Extracts

There is evidence that plant extracts often have greater activity than isolated constituents at a comparable dose, which is due to pharmacodynamic and/or pharmacokinetic synergy. In addition, the polyphenol–gut microbiota interactions and effects of food matrix on gut microbiomic profile and integrity have been demonstrated to contribute to the health effects of this class of compounds [92,93]. Studies concerning the comparison of the neuroprotective properties of extracts and pure compounds are very scarce. The single example is the report of Wu and Zhu [94] who revealed that contrary to the neuroprotective effect of Gingko biloba extract (EGb1), its two components ginkgolides A and B did not demonstrate any activity in the same experimental protocol.
The cocktail drink EM-X derived from the fermentation of rice, papaya, and seaweeds with effective microorganisms was examined in the 6-OHDA rat model of PD. The material tested was rich in lycopene, ubiquinone, various saponins as well as polyphenols, such as quercetin and kaempferol. Four-day administration of the test drink to rats treated with 6-OHDA, instead of drinking/tap water, attenuated the loss of TH- positive cells and the reduction of dopamine and its metabolite HVA levels in the SNpc [95].
Ginkgo biloba extracts have been used for many decades to increase cerebral blood flow and to treat dementia. The main active substances of G. biloba contributing to its neuroprotective effects belong to two groups of compounds, the flavone glycosides (quercetin, kaempferol, and isorhamnetin glycosides) and the terpene lactones (ginkgolides A, B, C, J, and bilobalide) [96]. In an early study by Wu and Zhu [94] the neuroprotective effects of a standardized extract of Ginkgo biloba (EGb 761) were investigated in mice lesioned with repeated administration of MPTP. Two-week administration of the extract tested (20, 50, 100 mg/kg/day, i.p.; 7 day pre-treatment and 6 day co-treatment) significantly protected striatal dopamine level [94]. The same extract was administered i.p. to rats for a week before 6-OHDA striatal injection and one week after surgery (50 or 100 mg/kg/day). Behavioral deficits induced by 6-OHDA were attenuated as well as dopamine neuron loss in the SNpc, and depletion of striatal dopamine were significantly reduced in rats treated with EGb 761 [97]. In the experimental model using MPTP, Gingko biloba extract administered to rats for 20 days (i.p., 50 or 100 mg/kg/day) was demonstrated to improve behavioral deficits, to reduce the decrease in dopamine level and SOD activity in the SNpc and to suppress the level of lipid peroxidation [98]. More parameters were assayed in the experiment involving 3-week pretreatment of rats with Gingko biloba extract (EGb) (50, 100, 150 mg/kg/day; p.o.) before a lesion induced by 6-OHDA. Beside the improvement of behavioral deficits, the extract dose-dependently restored the glutathione level, antioxidant enzymes activities, and lipid peroxidation level in the SNpc. The 6-OHDA-induced decrease in dopamine and its metabolites levels and the increased number of dopaminergic D2 receptors in the striatum were restored significantly following EGb treatment. Consequently, with the changes in these parameters, immunohistochemical analysis revealed a substantial increase in the density of TH-IR fibers in the substantia nigra [99].
Indisputably, green tea polyphenols (GTPs), including (−)-epigallocatechin gallate (EGCG), (−)-epicatechin gallate (ECG), (−)-epigallocatechin (EGC), and (−)-epicatechin (EC) have been one of the most extensively studied polyphenols pertaining to neuroprotective effects and not only in PD models. Seven-day pretreatment with GTPs (150 and 450 mg/kg/day; p.o.) was shown to protect midbrain and striatal dopaminergic neurons from 6-OHDA-induced cell death in rats. The effect correlated well with the decrease in the level of nitric oxide synthase (NOS), the content of reactive oxygen species (ROS) and nitric oxide (NO) as well as with attenuation of lipid peroxidation and enhancement of the free radical scavenging capability within the nigrostriatal region [100]. In MPTP-lesioned monkeys, oral administration of tea polyphenols (TPs), containing mainly EGCG, post to MPTP administration in a dose of 40 mg/kg/day for 80 days, caused the reduction in striatal synuclein oligomers’ level, an increase in the number of nigral TH-positive neurons and in the levels of DA and its metabolites DOPAC and HVA in the striatum, which was found to correlate well with improved motor deficits [101].
In another study, silymarin, a complex of flavonolignans derived from the seeds of Silybum marianum, has been demonstrated to exert neuroprotective effects through its antioxidant and anti-inflammatory properties. Silymarin treatment (50 and 100 mg/kg; i.p.) for 5 days after MPTP intoxication preserved dopamine levels as well as significantly diminished the number of apoptotic cells and preserved dopaminergic neurons in the substantia nigra region in mice [102]. Singhal et al. [103] examined neuroprotective properties of silymarin in a long-term experiment using another PD model involving two pesticides, maneb, and paraquat (MB + PQ). Silymarin was administered in a dose of 40 mg/kg (i.p.) to mice for 9 weeks, simultaneously some subsets of animals received maneb and paraquat i.p. twice a week. Silymarin treatment significantly attenuated MB+PQ-mediated reduction in locomotor activity, dopamine content, TH-immunoreactivity, and vesicular monoamine transporter (VMAT2) mRNA expression. Moreover, the substance tested reduced the number of degenerating neurons and restored the level of lipid peroxidation, nitrite content, mRNA expression of CYP2E1 and GST A, catalytic activities of CYP2E and GST as well as the expression of P-p53 and caspase-9 proteins. Hence, the authors concluded that silymarin could afford dopaminergic neuroprotection by the modulation of oxidative stress and apoptotic pathways [103].
The protective effects of Polygonum multiflorum extract (containing stilbenes, anthraquinones, and flavonoids) on the degeneration of nigrostriatal DA neurons induced by a combination of paraquat and maneb were investigated in a 6-week experiment on mice. Administration of the extract for 47 days simultaneously with neurotoxins (400 and 800 mg/kg/day; p.o.) significantly attenuated an impairment of behavioral performance and a decrease in the striatal dopamine level as well as in substantia nigra tyrosine hydroxylase-positive neurons [104].
Various extracts were investigated in the rat model of PD: tangerine peel rich in polymethoxylated flavones (35 mg/kg/day), cocoa rich in procyanidins (100 mg/kg/day), red clover rich in isoflavones (200 mg/kg/day), and grape seeds rich in catechins (100 mg/kg/day). Rats were treated p.o. for four days with the extracts, and the nigrostriatal lesion was induced by infusing 6-OHDA. Classical parameters of PD, namely the number of the dopaminergic cells in the SNpc and the levels of dopamine and its metabolites DOPAC and HVA in the striatum were quantified. Pretreatment of rats with the extracts of tangerine peel, cocoa, and red clover significantly attenuated the 6-OHDA-induced dopaminergic loss. No marked protection was seen in animals supplemented with grape seeds containing catechins [105].
Five-day pretreatment with a proanthocyanidin-rich fraction from the bark of Croton celtidifolius (10 mg/kg/day; i.p.) prevented mitochondrial complex-1 inhibition and a decrease in the expression of tyrosine hydroxylase in the striatum and olfactory bulb in rats given a single intranasal dose of MPTP. Moreover, the extract attenuated the short-term memory deficits, depressive-like behavior and a reduction in locomotor activity observed after MPTP administration. Catechins and oligomeric proanthocyanidins (formed of catechin and epicatechin units) are the main active substances found in the extract tested [102]. The authors suggested that besides antioxidant and ferric iron chelation properties the compounds of catechol-like structure inhibit the uptake of MPP+, the toxic metabolite of MPTP, and protect dopaminergic neurons. Additionally, these compounds inhibited monoaminooxidase B (MAO B), the enzyme catalyzing the oxidative deamination of monoamines including DA, thus decreasing its turnover, which resulted in the reduced formation of hydrogen peroxide and the limited generation of MPP+ from MPTP [106].
A polyphenol-rich aqueous walnut extract (JSE; an extract of Juglandis Semen), standardized based on caffeic acid content, has been demonstrated to protect dopaminergic neurons and to prevent the reduction of striatal DA and its metabolites, when it was given in a dose of 100 mg/kg/day (p.o.) for 6 days in mice exposed to MPTP. Moreover, the JSE treatment inhibited MPTP-induced PD-like symptoms, including reduced ambulation, bradykinesia, impaired motor coordination, and postural balance. The authors of this study suggested that JSE partly prevented the DA loss due to its MAO inhibitory and antioxidant activities [107].
Combined treatment with two polyphenols protocatechuic acid (PCA) and chrysin from the fruits of Alpinia oxyphylla at a dose of 10 mg/kg significantly attenuated both MPTP and 6-OHDA-induced dopaminergic neuron loss and ameliorated motor deficiencies in mice [108].

3. Epidemiological Studies

Data on dietary and lifestyle factors associated with PD are very sparse, especially in relation to dietary intake of polyphenols.
A large prospective study carried out over two decades involving almost 130,000 individuals showed that men with a high habitual intake of flavonoid-rich food and beverages were at a lower risk of PD. The association was confirmed for greater anthocyanin and anthocyanins-rich berries and apple intake and lower PD risk [109].
The early report concerning the relation of tea drinking habit to PD was published by Chan et al. [110]. Groups of 215 patients with PD and 313 controls in a Hong Kong Chinese population were enrolled in the study. It was demonstrated that black tea drinking reduced the risk of PD [110].
Further studies confirmed these findings. A case-control study conducted in Washington State in 1992–2000 involved 210 PD cases and 347 controls. Reduced risks of PD were observed for consumption of 2 cups or more of tea a day (relative risk = 0.4, 95% confidence interval: 0.2–0.9) and 2 or more cola drinks a day (relative risk = 0.6, 95% confidence interval:0.3–0.1.4) [111].
In a case-control study in a Chinese population, 310 PD patients and 500 controls were screened. It was found that 3 cups/day of tea for ten years led to a 28% reduction of PD [112].
Another study conducted in 1993–2005 on Chinese population involved 157 incident PD cases. Black tea intake showed an inverse association with PD risk (relative risk = 0.29, 95% confidence interval: 0.13–0.67) and ingredients other than caffeine were responsible for this effect. Contrary to numerous in vitro findings, green tea drinking was unrelated to PD risk [113]. The data on the inverse correlation between long-term consumption of tea and the onset of PD have been reported by Kandinow et al. [114]. Tea intake of at least 3 cups per day has been demonstrated to delay the motor symptoms onset of PD even by 7.7 years [114]. Other authors observed that consumption of one cup of tea per day or more dose-independently exerts neuroprotection in PD [115]. Finally, a meta-analysis study of tea drinkers and non-drinkers showed that the first ones have lesser risk being affected by PD [116]. This protective effect has been suggested to be attributed to the antioxidant, anti-inflammatory, and anti-excitotoxic activity of tea polyphenols [117].

4. The Permeability of Polyphenols Across the Blood-Brain Barrier

Despite evidence for the neuroprotective properties of polyphenols, information concerning their ability to enter CNS is limited. Since ingested polyphenols are extensively metabolized, it is critical to understand their metabolism as well as bioavailability and brain distribution of their metabolites to confirm the beneficial effects in PD.
To evaluate the ability of various compounds to gain entry to the brain, the in vitro blood-brain barrier (BBB) model consisting of brain endothelial cell lines and ECV304 monolayers co-cultured with C6 glioma cells was established. This cells culture showed upregulation of some features characteristic of the BBB in vivo including increased tight junction organization and elevated TERR. Using this model, the permeability of hesperetin, naringenin, and their respective glucuronides, as well as anthocyanins C3R and P3G across BBB, was demonstrated [118].
The above mentioned in vitro model does not fully mimic physiological conditions especially with respect to the effect of efflux transporters which restrict entry of xenobiotics to the brain. Hence, the same authors examined permeability of [3H] naringenin and [14C] quercetin across the blood-brain barrier using an in situ rat brain perfusion model. Test compounds were added to the perfusate injected into rat carotid artery, then their flux into different brain regions was measured. The results demonstrated that both compounds are able to traverse the BBB in vivo, although naringenin permeability was significantly higher. The authors suggested that this difference might be due to quercetin’s affinity to efflux transporters [119]. Similarly, de Boer et al. [120] have demonstrated that long-term feeding of rats with quercetin via diet (~500 mg/kg b.w./day) resulted in the deposition of quercetin and its metabolites isorhamnetin and tamarixetin in brain tissue.
Some flavonoids were demonstrated to cross the blood-brain barrier in a classical in vivo experiment. Following intravenous administration of naringenin to rats (20 mg/kg), the content of its total form (free and glucuronide conjugate) measured after 10 min in brain homogenate was 2.11 µg/g [121]. Unbound hesperetin (0.015 µg/mL) was determined in microdialysates from the brain striata of rats 30 min after intravenous injection of the compound (50 mg/kg) [122]. However, in both reports cited the route of administration was not representative of normal dietary ingestion, hence, in this experimental model processes of absorption, distribution, and gastrointestinal metabolism were not taken into consideration. More reliable are findings of experiments involving oral ingestion of compounds under investigation.
Following oral administration of tangeretin, a citrus flavonoid, (10 mg/kg/day) to rats for 28 days, significant amounts of the compound were found in various regions of the brain: 3.88 ng/mg in the hypothalamus, 2.36 ng/mg in the striatum, and 2.00 ng/mg in the hippocampus [27].
The presence of [3H](−) epigallocatechin gallate (EGCG) was found in the brain of mice administered the compound orally. Duplicate dosing of [3H] EGCG resulted in a 4-fold increase of radioactivity in the brain [123]. Another component of tea, epicatechin, was also shown to have access to the brain. After oral administration to rats of 100 mg/kg/day for 1, 5, and 10 days, epicatechin glucuronide and 3′-O-methyl epicatechin glucuronide were detected in the brain, however, their levels were too low for quantification [124]. The same epicatechin metabolite accumulated in the brain of rats after oral administration of grape-derived polyphenolic preparation as well as its components including proanthocyanidins, catechins, and epicatechins [125].
Anthocyanidins, a part of flavonoid family, widely distributed among pigmented fruit and vegetable were demonstrated to cross the blood-brain barrier. Two laboratories reported that anthocyanidins could enter the brain following dietary intake. In the experiment carried out by Talavera et al. [126] rats received a diet supplemented with 15 g blackberry extract per kg diet (14.8 mmol anthocyanidins per kg) for 15 days. The predominant compound found in the brain was cyanidin-3-glucoside which content was higher in the brain than in plasma (0.21 mmol/g tissue vs. 0.15 mmol/mL) [126]. Following a 10-week dietary supplementation with 2% blueberry extract, several anthocyanins were found in the brain of rats in unmetabolized forms. The main compounds identified in the hippocampus and the cortex were a galactoside form of cyanidin and malvidin as well as malvidin-3-O-β-arabinose [127].
Kalt et al. [128] reported the findings of the experiment in which pigs were fed diets supplemented with 1, 2, and 4% w/w of freeze-dried blueberries for four weeks. Eleven intact anthocyanidins were detected in the cortex and cerebellum, their total concentration was 0.88 and 0.66 pm/g tissue, respectively. The authors suggest that the consumption of low doses of flavonoids can lead to their accumulation in some tissues including brain even when they are not detected in the circulation [128].
Furthermore, other investigators have shown that ten-day treatment with a combination of polyphenol preparations from grapes or resveratrol led to the accumulation of proanthocyanidins and quercetin glucuronides (Gluc), including catechin-O-β-Gluc, 3′-O-methyl-catechin-O-β-Gluc, epicatechin-O-β-Gluc, 3′-O-methyl-epicatechin-O-β-Gluc, quercetin-3-OGluc, O-methyl-quercetin-Gluc, as well as free glycosides such as malvidin-3-O-glucoside (Glc), petunidin-Glc, delphinidin-Glc, peonidin-Glc, cyanidin-Glc, malvidin-3-O-Glc, petunidin-3-O-Glc, delphinidin-3-O-Glc, peonidin-3-O-Glc, and cyanidin-3-O-Glc in the brain of rats. Resveratrol was distributed into the brain, both in unchanged form and conjugated with glucuronic acid [129]. Chen et al. [130] reported that in pigs treated with polyphenols from apple/grape seed and bilberry extracts, metabolites such as catechin/epicatechin-glucuronides, 3′MeO-catechin/epicatechin-glucuronides, and quercetin glucuronides were accumulated in the cerebellum, brain stem, medial frontal cortex, hippocampus, hypothalamus, and amygdala, whereas glycosides of anthocyanidins, such as petunidin, malvidin, cyanidin, and peonidin were found only in the first three brain regions.
Besides the above-mentioned feeding with enriched foods, absorption and distribution of purified anthocyanidins was also evidenced. Passamonti et al. [131] introduced surgically a single dose of anthocyanidins extracted from grapes (8 mg/kg b.w.) into the stomach of rats. After 10 min, intact anthocyanidins, mainly malvidin-3-glucoside and its p-cumarate ester, were detected in the brain in quantities similar to that in the plasma, namely 192 ng/g and 176 ng/g, respectively [131]. Metabolism and tissue distribution of pelargonidin following oral administration to rats of a single dose of 50 mg/kg b.w. were investigated. Detectable levels of aglycone, 0.16 nmol/g, were found in the brain [132].

5. Polyphenols and the Microbiota–Gut–Brain Axis

Recent research has shown that the gut microbiota plays a key role in biotransformation and bioavailability and thus in the biological activities of phytoconstituents, particularly high-molecular-weight polyphenols. In addition, the interaction has been reported to be reciprocal, since not absorbed in the small intestine, oligomeric and polymeric polyphenols can modulate the composition of the colonic gut microbiota by inhibiting some bacterial populations or stimulating others [133]. Some polyphenols, such as flavan-3-ols, proanthocyanidins, and hydrolyzable tannins (ellagitannins), have been demonstrated to exert both selective prebiotic effects and selective antimicrobial effects against pathogenic gut bacteria [134,135]. Furthermore, commensals residing in the gut may exert a beneficial influence on the brain, since they biotransform polyphenols into metabolites with greater neuroprotective activity than their precursor structures. To achieve this health benefit, polyphenols require conversion into low-molecular-weight metabolites able to cross the blood-brain-barrier. In this respect, catabolic reactions involving C-ring opening and C–C bond breaking in polyphenols, carried out by gut microbiota, are crucial for the polyphenol’s activation. This process has been substantially studied in the catabolism of ellagitannins, mainly pomegranate’s punicalagins, to produce different urolithins by Gordonibacter spp. [136]. Ellagitannins are hydrolyzed in the upper gastrointestinal tract, releasing ellagic acid, which together with residual parent compounds is metabolized by gut microbiota to urolithins, including urolithins A, B, C, M6, and isourolithin A in the intestine. However, the ability to produce urolithins differs considerably among individuals depending on their colon microbial composition and three (A, B, and 0) human urolithin-producing metabotypes have been described. Metabotype B individuals produce urolithin B and/or isourolithin A and urolithin A (UA), metabotype A generates only urolithin A while those with metabotype 0 do not produce urolithins at all [137]. The absorbed urolithins undergo subsequent phase I and II metabolism including methylation, glucuronidation, and sulfation [138]. The colonic metabolites, due to their anti-inflammatory, antioxidant, and antiglycative activity, have recently attracted considerable interest in the context of neuroprotection [138,139,140]. However, the neuroprotective potential of urolithins have been mainly demonstrated in vitro. Only Yuan et al. [138] studied the neuroprotective effects of urolithins against Alzheimer’s disease in vivo and have shown that methyl-urolithin B protected Caenorhabditis elegants against Aβ1–42-induced neurotoxicity and paralysis. Neuroprotective activity of urolithins is also supported by human data. In middle-aged and older adults with mild memory complaints, drinking 8 ounces of punicalagin-rich pomegranate juice for 4 weeks caused a significant improvement in verbal and visual memory that correlated well with plasma urolithin A-glucuronide concentration [141]. Notably, UA has been recently reported to promote mitochondrial health and induce mitophagy in vivo, which has substantial implications for neuroprotection in Parkinson’s disease [142,143]. However, data about pharmacokinetics and tissue distribution, especially into the brain, of urolithins and their conjugates are somewhat scarce. There are only a few available reports in the literature confirming the distribution of urolithin A to the brain. Gasperotti et al. [144] have detected UA in the brains of rats injected with a mixture of polyphenol microbial metabolites including urolithin A. Seeram et al. [145] have found methylated-UA in the brains of mice administered intraperitoneally with UA. Therefore, Da Silva et al. [146] have recommended conducting animal studies using relevant doses of purified urolithins in order to elucidate their pharmacokinetics, pharmacodynamics, and mechanism of neuroprotective action.

6. Summary and Conclusions

A lot of in vivo data presented in the current review suggests that many natural polyphenols may protect against the neurodegenerative characteristics of Parkinson disease.
A wide array of assays reflecting the multifactorial etiology of PD has been employed to evidence the neuroprotective activity of the compounds under investigation. As dopamine depletion in the nigrostriatal region is a major pathological aspect of PD, in the majority of studies, the increase in dopamine and its metabolites’ concentrations as well as in the activity, protein, and mRNA levels of tyrosine hydroxylase, the enzyme responsible for DA synthesis, have been demonstrated.
Protection of dopaminergic neurons and attenuation of the loss of TH-positive cells by polyphenols have also been evidenced on the basis of histopathologic and immunohistochemical examination.
The beneficial effects of the compounds tested were supported by the results of behavioral examinations which indicated the improvement of locomotor activity, reduction of apomorphine-induced rotation, reduction in time of turning, improvement of motor balance and coordination, as well as attenuation of bradykinesia.
Multiple mechanisms of the anti-parkinsonian effects of polyphenols have been suggested in the reports cited in the current review. It is well established that oxidative stress and inflammation are significant factors underlying the process of neurodegeneration. Hence, the parameters of oxidative stress, including oxidative damage to essential cellular macromolecules, the content of GSH, as well as the activity of antioxidant enzymes, have been examined, demonstrating the potential of polyphenols to counteract oxidative damage in the striatum. There is abundant evidence that α-synuclein oligomers trigger activation of astrocytes and microglia cells which secrete proinflammatory cytokines leading to neuroinflammation and mitochondrial dysfunction [147]. A number of polyphenols have been shown to inhibit microglial activation accompanied with the suppressed production of TNF-α, IL-6, IL-1β, as well as COX-2, iNOS, and NO. Simultaneously, some of the compounds discussed enhanced the expression of neurotrophic factors such as glial cell-derived neurotrophic factor (GDNF) which is involved in the survival of DA neurons.
Recently, growing attention is being directed towards some molecular pathways underlying the neuroprotective activity of natural compounds in vivo. Many polyphenols have been shown to restore the reduced activities of protein kinases C (PKCs) which play a fundamental role in programmed cell death and cell survival, thus inhibiting apoptotic signaling pathways. These compounds may modulate events associated with apoptosis by lowering the expression of several proapoptotic genes, such as Bax and Bad, and by increasing the level of antiapoptotic Bcl-2 as well as the inhibition of caspases.
Another signaling pathway involved in apoptotic events which was found to be modulated by polyphenols are Jun N-terminal kinases (JNK). JNKs, belonging to the superfamily of MAP kinases, are responsive to stress stimuli such as cytokines or ROS and play a significant role in stimulating apoptosis. As mentioned above, some polyphenols have been shown to reduce activation of JNKs and transcription factor c-Jun which result in the suppression of dopaminergic neurons’ apoptosis. Thus, the JNK pathway might be considered one of the novel therapeutic targets for neurodegenerative diseases [148].
The promising results of the above presented studies make the continuing search for phytochemicals exerting neuroprotective effects on DA neurons and thus retarding their degeneration highly desirable. Especially significant in future investigations will be the understanding of the molecular and cellular pathways involved in the neurodegenerative process, which will enable the identification of potential targets for effective treatments. The ultimate deliverable is also expected from ongoing research on metabolic derivatives of polyphenols, in particular, bioavailable gut microbiota metabolites, which offer a novel therapeutic approach to the treatment of Parkinson’s disease.

Author Contributions

Both authors contributed to the gathering of source articles and writing of this paper.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

4-HNE4-hydroxynonenal
6-OHDA6-Hydroxydopamine
AChEAcetylcholinesterase
AKTserine/threonine protein kinase
AREAntioxidant responsive element
BaxBcl-2-like protein 4
Bcl-2B-cell lymphoma 2
BDNFBrain-derived neurotrophic factor.
b.w.Body weight
CaMKIICa2+/calmodulin-dependent protein kinase II
Casp-3Caspase-3
Casp-8Caspase-8
Casp-9Caspase-9
CATCatalase
CHIPC terminus Hsp70 interacting protein
CNSCentral nervous system
COICytochrome c oxidase I
COX-2Cyclooxygenase-2
CURCurcumin
DADopamine
DARPP-32Dopamine- and cAMP-regulated phosphoprotein
DATDopamine transporter
DJ-1Protein deglycase
DOPAC3,4-dihydroxyphenylacetic acid
EC(−)-Epicatechin
ECGEpicatechin gallate
EGC(−)-Epigallocatechin
EGCGEpigallocatechin-3-gallate
ERKExtracellular signal-regulated kinase protein-serine/threonine kinase
FAFerulic acid
GAPDHGlyceraldehyde 3-phosphate dehydrogenase
GABAGamma-aminobutyric acid
GCLCGlutamate cysteine ligase catalytic subunit
GCLMGlutamate cysteine ligase modifier subunit
GDNFGlial cell-derived neurotrophic factor
GFAPGlial fibrillary acidic protein
GLuGlutamate (or Glutamic acid)
GPxGlutathione peroxidase
GRGlutathione reductase
GRP-7878 kDa Glucose-regulated protein
GSGlutamine synthetase
GSHReduced glutathione
GSSGOxidized glutathione
GTPsGreen tea polyphenols
HcyHomocysteine
HO-1Heme oxygenase-1
Hsp70sThe 70 kDa heat shock proteins
HVAHomovanillic acid
Iba1Ionised calcium binding adapter molecule 1
i.p.Intraperitoneally
IL-1βInterleukin 1 beta
IL-6Interleukin 6
IL-10Interleukin 10
iNOSInducible nitric oxide synthase
JNKc-Jun N-terminal kinase
JSEJuglandis Semen extract
L-DOPAl-3,4-dihydroxyphenylalanine
LPSLipopolysaccharide
MAOMonoamine oxidase
MAPKMitogen-activated protein kinase
MBManeb
MDAMalondialdehyde
MMPsN-methyl-4-phenyl-1,2,3,6-tetrahydropyridine
MPTPMitochondrial transmembrane potentials
mTORC1Mammalian target of rapamycin complex 1
NF-кBNuclear factor-кB
nNOSNuclear nitric oxide synthase
NONitric oxide
Nrf2Nuclear factor (erythroid-derived 2)-like 2
NQO1NAD(P)H: quinone oxidoreductase 1
(p)probenocid
p-4E-BP1Phosphorylated 4E-binding protein 1
p38MAP Kinase (MAPK), CSBP Cytokinin-Specific Binding Protein or RK
PARPPoly(ADP-ribose) polymerase
PCAProtocatechuic acid
PCsProtein carbonyls
PDParkinson disease
PGE2Prostaglandin E2
PI3Kphosphoinositide 3-kinase
PKCProtein kinase C
PLA2Phospholipases A2
p.o.Orally
PQResveratrol
RESParaquat
ROSReactive oxygen species
ROTRotenone
SASyringic acid
SILSilibinin
SNpcSubstantia nigra pars compacta
SODSuperoxide dismutase
TBARSThiobarbituric acid reactive substances
THTyrosine hydroxylase
TNF-αTumor necrosis factor alpha
TRAPTotal reactive antioxidant potential
TrxThioredoxin
TFTheaflavin
TPsTea polyphenols
TrkBTropomyosin receptor kinase B
TUNELTerminal deoxynucleotidyl transferase dUTP nick end labeling
UAUrolithin A
VMAT2Vesicular monoamine transporter-2

References

  1. Reglodi, D.; Renaud, J.; Tamas, A.; Tizabi, Y.; Socías, S.B.; Del-Bel, E.; Raisman-Vozari, R. Novel tactics for neuroprotection in parkinson’s disease: Role of antibiotics, polyphenols and neuropeptides. Prog. Neurobiol. 2017, 155, 120–148. [Google Scholar] [CrossRef] [PubMed]
  2. Costa, S.L.; Silva, V.D.; Dos Santos Souza, C.; Santos, C.C.; Paris, I.; Muñoz, P.; Segura-Aguilar, J. Impact of plant-derived flavonoids on neurodegenerative diseases. Neurotox. Res. 2016, 30, 41–52. [Google Scholar] [CrossRef] [PubMed]
  3. Mandel, S.; Youdim, M.B. Catechin polyphenols: Neurodegeneration and neuroprotection in neurodegenerative diseases. Free Radic. Biol. Med. 2004, 37, 304–317. [Google Scholar] [CrossRef] [PubMed]
  4. Costa, C.; Tsatsakis, A.; Mamoulakis, C.; Teodoro, M.; Briguglio, G.; Caruso, E.; Tsoukalas, D.; Margina, D.; Dardiotis, E.; Kouretas, D.; et al. Current evidence on the effect of dietary polyphenols intake on chronic diseases. Food Chem. Toxicol. 2017, 110, 286–299. [Google Scholar] [CrossRef] [PubMed]
  5. Moosavi, F.; Hosseini, R.; Saso, L.; Firuzi, O. Modulation of neurotrophic signaling pathways by polyphenols. Drug Des. Dev. Ther. 2016, 10, 23–42. [Google Scholar]
  6. Pan, T.; Jankovic, J.; Le, W. Potential therapeutic properties of green tea polyphenols in parkinson’s disease. Drugs Aging 2003, 20, 711–721. [Google Scholar] [CrossRef] [PubMed]
  7. Youdim, K.A.; Shukitt-Hale, B.; Joseph, J.A. Flavonoids and the brain: Interactions at the blood-brain barrier and their physiological effects on the central nervous system. Free Radic. Biol. Med. 2004, 37, 1683–1693. [Google Scholar] [CrossRef] [PubMed]
  8. Youdim, K.A.; Spencer, J.P.; Schroeter, H.; Rice-Evans, C. Dietary flavonoids as potential neuroprotectants. Biol. Chem. 2002, 383, 503–519. [Google Scholar] [CrossRef] [PubMed]
  9. Vauzour, D.; Vafeiadou, K.; Rodriguez-Mateos, A.; Rendeiro, C.; Spencer, J.P. The neuroprotective potential of flavonoids: A multiplicity of effects. Genes Nutr. 2008, 3, 115–126. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, B. Natural antioxidants protect neurons in alzheimer’s disease and parkinson’s disease. Neurochem. Res. 2009, 34, 630–638. [Google Scholar] [CrossRef] [PubMed]
  11. Weinreb, O.; Amit, T.; Mandel, S.; Youdim, M.B. Neuroprotective molecular mechanisms of (−)-epigallocatechin-3-gallate: A reflective outcome of its antioxidant, iron chelating and neuritogenic properties. Genes Nutr. 2009, 4, 283–296. [Google Scholar] [CrossRef] [PubMed]
  12. Mandel, S.; Amit, T.; Reznichenko, L.; Weinreb, O.; Youdim, M.B. Green tea catechins as brain-permeable, natural iron chelators-antioxidants for the treatment of neurodegenerative disorders. Mol. Nutr. Food Res. 2006, 50, 229–234. [Google Scholar] [CrossRef] [PubMed]
  13. Vingill, S.; Connor-Robson, N.; Wade-Martins, R. Are rodent models of parkinson’s disease behaving as they should? Behav Brain Res. 2017. [Google Scholar] [CrossRef] [PubMed]
  14. Zbarsky, V.; Datla, K.P.; Parkar, S.; Rai, D.K.; Aruoma, O.I.; Dexter, D.T. Neuroprotective properties of the natural phenolic antioxidants curcumin and naringenin but not quercetin and fisetin in a 6-ohda model of parkinson’s disease. Free Radic. Res. 2005, 39, 1119–1125. [Google Scholar] [CrossRef] [PubMed]
  15. Lou, H.; Jing, X.; Wei, X.; Shi, H.; Ren, D.; Zhang, X. Naringenin protects against 6-ohda-induced neurotoxicity via activation of the nrf2/are signaling pathway. Neuropharmacology 2014, 79, 380–388. [Google Scholar] [CrossRef] [PubMed]
  16. Kim, H.D.; Jeong, K.H.; Jung, U.J.; Kim, S.R. Naringin treatment induces neuroprotective effects in a mouse model of parkinson’s disease in vivo, but not enough to restore the lesioned dopaminergic system. J. Nutr. Biochem. 2016, 28, 140–146. [Google Scholar] [CrossRef] [PubMed]
  17. Antunes, M.S.; Goes, A.T.; Boeira, S.P.; Prigol, M.; Jesse, C.R. Protective effect of hesperidin in a model of parkinson’s disease induced by 6-hydroxydopamine in aged mice. Nutrition 2014, 30, 1415–1422. [Google Scholar] [CrossRef] [PubMed]
  18. Kiasalari, Z.; Khalili, M.; Baluchnejadmojarad, T.; Roghani, M. Protective effect of oral hesperetin against unilateral striatal 6-hydroxydopamine damage in the rat. Neurochem. Res. 2016, 41, 1065–1072. [Google Scholar] [CrossRef] [PubMed]
  19. Kääriäinen, T.M.; Piltonen, M.; Ossola, B.; Kekki, H.; Lehtonen, S.; Nenonen, T.; Lecklin, A.; Raasmaja, A.; Männistö, P.T. Lack of robust protective effect of quercetin in two types of 6-hydroxydopamine-induced parkinsonian models in rats and dopaminergic cell cultures. Brain Res. 2008, 1203, 149–159. [Google Scholar] [CrossRef] [PubMed]
  20. Haleagrahara, N.; Siew, C.J.; Mitra, N.K.; Kumari, M. Neuroprotective effect of bioflavonoid quercetin in 6-hydroxydopamine-induced oxidative stress biomarkers in the rat striatum. Neurosci. Lett. 2011, 500, 139–143. [Google Scholar] [CrossRef] [PubMed]
  21. Khan, M.M.; Raza, S.S.; Javed, H.; Ahmad, A.; Khan, A.; Islam, F.; Safhi, M.M. Rutin protects dopaminergic neurons from oxidative stress in an animal model of parkinson’s disease. Neurotox. Res. 2012, 22, 1–15. [Google Scholar] [CrossRef] [PubMed]
  22. Baluchnejadmojarad, T.; Jamali-Raeufy, N.; Zabihnejad, S.; Rabiee, N.; Roghani, M. Troxerutin exerts neuroprotection in 6-hydroxydopamine lesion rat model of parkinson’s disease: Possible involvement of pi3k/erβ signaling. Eur. J. Pharmacol. 2017, 801, 72–78. [Google Scholar] [CrossRef] [PubMed]
  23. Ma, Z.G.; Wang, J.; Jiang, H.; Liu, T.W.; Xie, J.X. Myricetin reduces 6-hydroxydopamine-induced dopamine neuron degeneration in rats. Neuroreport 2007, 18, 1181–1185. [Google Scholar] [CrossRef] [PubMed]
  24. Baluchnejadmojarad, T.; Roghani, M.; Nadoushan, M.R.; Bagheri, M. Neuroprotective effect of genistein in 6-hydroxydopamine hemi-parkinsonian rat model. Phytother. Res. 2009, 23, 132–135. [Google Scholar] [CrossRef] [PubMed]
  25. Zhu, G.; Wang, X.; Chen, Y.; Yang, S.; Cheng, H.; Wang, N.; Li, Q. Puerarin protects dopaminergic neurons against 6-hydroxydopamine neurotoxicity via inhibiting apoptosis and upregulating glial cell line-derived neurotrophic factor in a rat model of parkinson’s disease. Planta Med. 2010, 76, 1820–1826. [Google Scholar] [CrossRef] [PubMed]
  26. Yu, X.; He, G.R.; Sun, L.; Lan, X.; Shi, L.L.; Xuan, Z.H.; Du, G.H. Assessment of the treatment effect of baicalein on a model of parkinsonian tremor and elucidation of the mechanism. Life Sci. 2012, 91, 5–13. [Google Scholar] [CrossRef] [PubMed]
  27. Datla, K.P.; Christidou, M.; Widmer, W.W.; Rooprai, H.K.; Dexter, D.T. Tissue distribution and neuroprotective effects of citrus flavonoid tangeretin in a rat model of parkinson’s disease. Neuroreport 2001, 12, 3871–3875. [Google Scholar] [CrossRef] [PubMed]
  28. Leaver, K.R.; Allbutt, H.N.; Creber, N.J.; Kassiou, M.; Henderson, J.M. Oral pre-treatment with epigallocatechin gallate in 6-ohda lesioned rats produces subtle symptomatic relief but not neuroprotection. Brain Res. Bull. 2009, 80, 397–402. [Google Scholar] [CrossRef] [PubMed]
  29. Roghani, M.; Niknam, A.; Jalali-Nadoushan, M.R.; Kiasalari, Z.; Khalili, M.; Baluchnejadmojarad, T. Oral pelargonidin exerts dose-dependent neuroprotection in 6-hydroxydopamine rat model of hemi-parkinsonism. Brain Res. Bull. 2010, 82, 279–283. [Google Scholar] [CrossRef] [PubMed]
  30. Jin, F.; Wu, Q.; Lu, Y.F.; Gong, Q.H.; Shi, J.S. Neuroprotective effect of resveratrol on 6-ohda-induced parkinson’s disease in rats. Eur. J. Pharmacol. 2008, 600, 78–82. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, Y.; Xu, H.; Fu, Q.; Ma, R.; Xiang, J. Protective effect of resveratrol derived from polygonum cuspidatum and its liposomal form on nigral cells in parkinsonian rats. J. Neurol. Sci. 2011, 304, 29–34. [Google Scholar] [CrossRef] [PubMed]
  32. Khan, M.M.; Ahmad, A.; Ishrat, T.; Khan, M.B.; Hoda, M.N.; Khuwaja, G.; Raza, S.S.; Khan, A.; Javed, H.; Vaibhav, K.; et al. Resveratrol attenuates 6-hydroxydopamine-induced oxidative damage and dopamine depletion in rat model of parkinson’s disease. Brain Res. 2010, 1328, 139–151. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, Y.; Zhang, D.Q.; Liao, Z.; Wang, B.; Gong, S.; Wang, C.; Zhang, M.Z.; Wang, G.H.; Cai, H.; Liao, F.F.; et al. Anti-oxidant polydatin (piceid) protects against substantia nigral motor degeneration in multiple rodent models of parkinson’s disease. Mol. Neurodegener. 2015, 10, 4. [Google Scholar] [CrossRef] [PubMed]
  34. Du, X.X.; Xu, H.M.; Jiang, H.; Song, N.; Wang, J.; Xie, J.X. Curcumin protects nigral dopaminergic neurons by iron-chelation in the 6-hydroxydopamine rat model of parkinson’s disease. Neurosci. Bull. 2012, 28, 253–258. [Google Scholar] [CrossRef] [PubMed]
  35. Tripanichkul, W.; Jaroensuppaperch, E.O. Ameliorating effects of curcumin on 6-ohda-induced dopaminergic denervation, glial response, and sod1 reduction in the striatum of hemiparkinsonian mice. Eur. Rev. Med. Pharmacol. Sci. 2013, 17, 1360–1368. [Google Scholar] [PubMed]
  36. Wu, C.R.; Tsai, C.W.; Chang, S.W.; Lin, C.Y.; Huang, L.C. Carnosic acid protects against 6-hydroxydopamine-induced neurotoxicity in in vivo and in vitro model of parkinson’s disease: Involvement of antioxidative enzymes induction. Chem. Biol. Interact. 2015, 225, 40–46. [Google Scholar] [CrossRef] [PubMed]
  37. Zare, K.; Eidi, A.; Roghani, M.; Rohani, A.H. The neuroprotective potential of sinapic acid in the 6-hydroxydopamine-induced hemi-parkinsonian rat. Metab. Brain Dis. 2015, 30, 205–213. [Google Scholar] [CrossRef] [PubMed]
  38. Farbood, Y.; Sarkaki, A.; Dolatshahi, M.; Taqhi Mansouri, S.M.; Khodadadi, A. Ellagic acid protects the brain against 6-hydroxydopamine induced neuroinflammation in a rat model of parkinson’s disease. Basic Clin. Neurosci. 2015, 6, 83–89. [Google Scholar] [PubMed]
  39. Lastres-Becker, I.; Molina-Holgado, F.; Ramos, J.A.; Mechoulam, R.; Fernández-Ruiz, J. Cannabinoids provide neuroprotection against 6-hydroxydopamine toxicity in vivo and in vitro: Relevance to parkinson’s disease. Neurobiol. Dis. 2005, 19, 96–107. [Google Scholar] [CrossRef] [PubMed]
  40. Levites, Y.; Weinreb, O.; Maor, G.; Youdim, M.B.; Mandel, S. Green tea polyphenol (-)-epigallocatechin-3-gallate prevents n-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced dopaminergic neurodegeneration. J. Neurochem. 2001, 78, 1073–1082. [Google Scholar] [CrossRef] [PubMed]
  41. Mandel, S.; Maor, G.; Youdim, M.B. Iron and alpha-synuclein in the substantia nigra of mptp-treated mice: Effect of neuroprotective drugs r-apomorphine and green tea polyphenol (-)-epigallocatechin-3-gallate. J. Mol. Neurosci. 2004, 24, 401–416. [Google Scholar] [CrossRef]
  42. Choi, J.Y.; Park, C.S.; Kim, D.J.; Cho, M.H.; Jin, B.K.; Pie, J.E.; Chung, W.G. Prevention of nitric oxide-mediated 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced parkinson’s disease in mice by tea phenolic epigallocatechin 3-gallate. Neurotoxicology 2002, 23, 367–374. [Google Scholar] [CrossRef]
  43. Xu, Q.; Langley, M.; Kanthasamy, A.G.; Reddy, M.B. Epigallocatechin gallate has a neurorescue effect in a mouse model of parkinson disease. J. Nutr. 2017, 147, 1926–1931. [Google Scholar] [CrossRef] [PubMed]
  44. Reznichenko, L.; Kalfon, L.; Amit, T.; Youdim, M.B.; Mandel, S.A. Low dosage of rasagiline and epigallocatechin gallate synergistically restored the nigrostriatal axis in mptp-induced parkinsonism. Neurodegener. Dis. 2010, 7, 219–231. [Google Scholar] [CrossRef] [PubMed]
  45. Kim, J.S.; Kim, J.M.; O, J.J.; Jeon, B.S. Inhibition of inducible nitric oxide synthase expression and cell death by (-)-epigallocatechin-3-gallate, a green tea catechin, in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of parkinson’s disease. J. Clin. Neurosci. 2010, 17, 1165–1168. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, T.; Zhu, M.; Liang, Z. (-)-epigallocatechin-3-gallate modulates peripheral immunity in the mptp-induced mouse model of parkinson’s disease. Mol. Med. Rep. 2018, 17, 4883–4888. [Google Scholar] [CrossRef] [PubMed]
  47. Kim, H.G.; Ju, M.S.; Ha, S.K.; Lee, H.; Kim, S.Y.; Oh, M.S. Acacetin protects dopaminergic cells against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced neuroinflammation in vitro and in vivo. Biol. Pharm. Bull. 2012, 35, 1287–1294. [Google Scholar] [CrossRef] [PubMed]
  48. Xue, X.; Liu, H.; Qi, L.; Li, X.; Guo, C.; Gong, D.; Qu, H. Baicalein ameliorated the upregulation of striatal glutamatergic transmission in the mice model of parkinson’s disease. Brain Res. Bull. 2014, 103, 54–59. [Google Scholar] [CrossRef] [PubMed]
  49. Lee, E.; Park, H.R.; Ji, S.T.; Lee, Y.; Lee, J. Baicalein attenuates astroglial activation in the 1-methyl-4-phenyl-1,2,3,4-tetrahydropyridine-induced parkinson’s disease model by downregulating the activations of nuclear factor-κb, erk, and jnk. J. Neurosci. Res. 2014, 92, 130–139. [Google Scholar] [CrossRef] [PubMed]
  50. Jang, S.W.; Liu, X.; Yepes, M.; Shepherd, K.R.; Miller, G.W.; Liu, Y.; Wilson, W.D.; Xiao, G.; Blanchi, B.; Sun, Y.E.; et al. A selective trkb agonist with potent neurotrophic activities by 7,8-dihydroxyflavone. Proc. Natl. Acad. Sci. USA 2010, 107, 2687–2692. [Google Scholar] [CrossRef] [PubMed]
  51. Takano, K.; Tabata, Y.; Kitao, Y.; Murakami, R.; Suzuki, H.; Yamada, M.; Iinuma, M.; Yoneda, Y.; Ogawa, S.; Hori, O. Methoxyflavones protect cells against endoplasmic reticulum stress and neurotoxin. Am. J. Physiol. Cell Physiol. 2007, 292, C353–C361. [Google Scholar] [CrossRef] [PubMed]
  52. Jeong, K.H.; Jeon, M.T.; Kim, H.D.; Jung, U.J.; Jang, M.C.; Chu, J.W.; Yang, S.J.; Choi, I.Y.; Choi, M.S.; Kim, S.R. Nobiletin protects dopaminergic neurons in the 1-methyl-4-phenylpyridinium-treated rat model of parkinson’s disease. J. Med. Food 2015, 18, 409–414. [Google Scholar] [CrossRef] [PubMed]
  53. Yabuki, Y.; Ohizumi, Y.; Yokosuka, A.; Mimaki, Y.; Fukunaga, K. Nobiletin treatment improves motor and cognitive deficits seen in mptp-induced parkinson model mice. Neuroscience 2014, 259, 126–141. [Google Scholar] [CrossRef] [PubMed]
  54. Leem, E.; Nam, J.H.; Jeon, M.T.; Shin, W.H.; Won, S.Y.; Park, S.J.; Choi, M.S.; Jin, B.K.; Jung, U.J.; Kim, S.R. Naringin protects the nigrostriatal dopaminergic projection through induction of gdnf in a neurotoxin model of parkinson’s disease. J. Nutr. Biochem. 2014, 25, 801–806. [Google Scholar] [CrossRef] [PubMed]
  55. Li, S.; Pu, X.P. Neuroprotective effect of kaempferol against a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced mouse model of parkinson’s disease. Biol. Pharm. Bull. 2011, 34, 1291–1296. [Google Scholar] [CrossRef] [PubMed]
  56. Lv, C.; Hong, T.; Yang, Z.; Zhang, Y.; Wang, L.; Dong, M.; Zhao, J.; Mu, J.; Meng, Y. Effect of quercetin in the 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine-induced mouse model of parkinson’s disease. Evid. Based Complement. Alternat. Med. 2012, 2012, 928643. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, Z.T.; Cao, X.B.; Xiong, N.; Wang, H.C.; Huang, J.S.; Sun, S.G.; Wang, T. Morin exerts neuroprotective actions in parkinson disease models in vitro and in vivo. Acta Pharmacol. Sin. 2010, 31, 900–906. [Google Scholar] [CrossRef] [PubMed]
  58. Lee, Y.; Park, H.R.; Chun, H.J.; Lee, J. Silibinin prevents dopaminergic neuronal loss in a mouse model of parkinson’s disease via mitochondrial stabilization. J. Neurosci. Res. 2015, 93, 755–765. [Google Scholar] [CrossRef] [PubMed]
  59. Jung, U.J.; Jeon, M.T.; Choi, M.S.; Kim, S.R. Silibinin attenuates mpp⁺-induced neurotoxicity in the substantia nigra in vivo. J. Med. Food 2014, 17, 599–605. [Google Scholar] [CrossRef] [PubMed]
  60. Geed, M.; Garabadu, D.; Ahmad, A.; Krishnamurthy, S. Silibinin pretreatment attenuates biochemical and behavioral changes induced by intrastriatal mpp+ injection in rats. Pharmacol. Biochem. Behav. 2014, 117, 92–103. [Google Scholar] [CrossRef] [PubMed]
  61. Liu, L.X.; Chen, W.F.; Xie, J.X.; Wong, M.S. Neuroprotective effects of genistein on dopaminergic neurons in the mice model of parkinson’s disease. Neurosci. Res. 2008, 60, 156–161. [Google Scholar] [CrossRef] [PubMed]
  62. Kumar, H.; Kim, I.S.; More, S.V.; Kim, B.W.; Bahk, Y.Y.; Choi, D.K. Gastrodin protects apoptotic dopaminergic neurons in a toxin-induced parkinson’s disease model. Evid. Based Complement. Alternat. Med. 2013, 2013, 514095. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, S.; Jing, H.; Yang, H.; Liu, Z.; Guo, H.; Chai, L.; Hu, L. Tanshinone i selectively suppresses pro-inflammatory genes expression in activated microglia and prevents nigrostriatal dopaminergic neurodegeneration in a mouse model of parkinson’s disease. J. Ethnopharmacol. 2015, 164, 247–255. [Google Scholar] [CrossRef] [PubMed]
  64. Nagarajan, S.; Chellappan, D.R.; Chinnaswamy, P.; Thulasingam, S. Ferulic acid pretreatment mitigates mptp-induced motor impairment and histopathological alterations in c57bl/6 mice. Pharm. Biol. 2015, 53, 1591–1601. [Google Scholar] [CrossRef] [PubMed]
  65. Xu, C.L.; Qu, R.; Zhang, J.; Li, L.F.; Ma, S.P. Neuroprotective effects of madecassoside in early stage of parkinson’s disease induced by mptp in rats. Fitoterapia 2013, 90, 112–118. [Google Scholar] [CrossRef] [PubMed]
  66. Vajragupta, O.; Boonchoong, P.; Watanabe, H.; Tohda, M.; Kummasud, N.; Sumanont, Y. Manganese complexes of curcumin and its derivatives: Evaluation for the radical scavenging ability and neuroprotective activity. Free Radic. Biol. Med. 2003, 35, 1632–1644. [Google Scholar] [CrossRef] [PubMed]
  67. Rajeswari, A.; Sabesan, M. Inhibition of monoamine oxidase-b by the polyphenolic compound, curcumin and its metabolite tetrahydrocurcumin, in a model of parkinson’s disease induced by mptp neurodegeneration in mice. Inflammopharmacology 2008, 16, 96–99. [Google Scholar] [CrossRef] [PubMed]
  68. Ojha, R.P.; Rastogi, M.; Devi, B.P.; Agrawal, A.; Dubey, G.P. Neuroprotective effect of curcuminoids against inflammation-mediated dopaminergic neurodegeneration in the mptp model of parkinson’s disease. J. Neuroimmune Pharmacol. 2012, 7, 609–618. [Google Scholar] [CrossRef] [PubMed]
  69. Chen, X.C.; Zhou, Y.C.; Chen, Y.; Zhu, Y.G.; Fang, F.; Chen, L.M. Ginsenoside rg1 reduces mptp-induced substantia nigra neuron loss by suppressing oxidative stress. Acta Pharmacol. Sin. 2005, 26, 56–62. [Google Scholar] [CrossRef] [PubMed]
  70. Anandhan, A.; Tamilselvam, K.; Radhiga, T.; Rao, S.; Essa, M.M.; Manivasagam, T. Theaflavin, a black tea polyphenol, protects nigral dopaminergic neurons against chronic mptp/probenecid induced parkinson’s disease. Brain Res. 2012, 1433, 104–113. [Google Scholar] [CrossRef] [PubMed]
  71. Anandhan, A.; Janakiraman, U.; Manivasagam, T. Theaflavin ameliorates behavioral deficits, biochemical indices and monoamine transporters expression against subacute 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (mptp)-induced mouse model of parkinson’s disease. Neuroscience 2012, 218, 257–267. [Google Scholar] [CrossRef] [PubMed]
  72. Anandhan, A.; Essa, M.M.; Manivasagam, T. Therapeutic attenuation of neuroinflammation and apoptosis by black tea theaflavin in chronic mptp/probenecid model of parkinson’s disease. Neurotox. Res. 2013, 23, 166–173. [Google Scholar] [CrossRef] [PubMed]
  73. Patil, S.P.; Jain, P.D.; Sancheti, J.S.; Ghumatkar, P.J.; Tambe, R.; Sathaye, S. Neuroprotective and neurotrophic effects of apigenin and luteolin in mptp induced parkinsonism in mice. Neuropharmacology 2014, 86, 192–202. [Google Scholar] [CrossRef] [PubMed]
  74. Rekha, K.R.; Selvakumar, G.P.; Sivakamasundari, R.I. Effects of syringic acid on chronic mptp/probenecid induced motor dysfunction, dopaminergic markers expression and neuroinflammation in c57bl/6 mice. Biomed. Aging Pathol. 2014, 4, 95–104. [Google Scholar] [CrossRef]
  75. Karuppagounder, S.S.; Madathil, S.K.; Pandey, M.; Haobam, R.; Rajamma, U.; Mohanakumar, K.P. Quercetin up-regulates mitochondrial complex-i activity to protect against programmed cell death in rotenone model of parkinson’s disease in rats. Neuroscience 2013, 236, 136–148. [Google Scholar] [CrossRef] [PubMed]
  76. Cui, Q.; Li, X.; Zhu, H. Curcumin ameliorates dopaminergic neuronal oxidative damage via activation of the akt/nrf2 pathway. Mol. Med. Rep. 2016, 13, 1381–1388. [Google Scholar] [CrossRef] [PubMed]
  77. Ojha, S.; Javed, H.; Azimullah, S.; Abul Khair, S.B.; Haque, M.E. Neuroprotective potential of ferulic acid in the rotenone model of parkinson’s disease. Drug Des. Dev. Ther. 2015, 9, 5499–5510. [Google Scholar]
  78. Sonia Angeline, M.; Sarkar, A.; Anand, K.; Ambasta, R.K.; Kumar, P. Sesamol and naringenin reverse the effect of rotenone-induced pd rat model. Neuroscience 2013, 254, 379–394. [Google Scholar] [CrossRef] [PubMed]
  79. Al-Amri, J.S.; Hagras, M.M.; Mohamed I, M. Effect of epigallocatechin-3-gallate on inflammatory mediators release in lps-induced parkinson’s disease in rats. Indian J. Exp. Biol. 2013, 51, 357–362. [Google Scholar] [PubMed]
  80. Mansouri, Z.; Sabetkasaei, M.; Moradi, F.; Masoudnia, F.; Ataie, A. Curcumin has neuroprotection effect on homocysteine rat model of parkinson. J. Mol. Neurosci. 2012, 47, 234–242. [Google Scholar] [CrossRef] [PubMed]
  81. Grünblatt, E.; Mandel, S.; Youdim, M.B. Neuroprotective strategies in parkinson’s disease using the models of 6-hydroxydopamine and mptp. Ann. N. Y. Acad. Sci. 2000, 899, 262–273. [Google Scholar] [CrossRef] [PubMed]
  82. Shahpiri, Z.; Bahramsoltani, R.; Hosein Farzaei, M.; Farzaei, F.; Rahimi, R. Phytochemicals as future drugs for parkinson’s disease: A comprehensive review. Rev. Neurosci. 2016, 27, 651–668. [Google Scholar] [CrossRef] [PubMed]
  83. Xia, N.; Fang, F.; Zhang, P.; Cui, J.; Tep-Cullison, C.; Hamerley, T.; Lee, H.J.; Palmer, T.; Bothner, B.; Lee, J.H.; et al. A knockin reporter allows purification and characterization of mda neurons from heterogeneous populations. Cell Rep. 2017, 18, 2533–2546. [Google Scholar] [CrossRef] [PubMed]
  84. Kim, H.D.; Jeong, K.H.; Jung, U.J.; Kim, S.R. Myricitrin ameliorates 6-hydroxydopamine-induced dopaminergic neuronal loss in the substantia nigra of mouse brain. J. Med. Food 2016, 19, 374–382. [Google Scholar] [CrossRef] [PubMed]
  85. Jung, U.J.; Kim, S.R. Beneficial effects of flavonoids against parkinson’s disease. J. Med. Food 2018. [Google Scholar] [CrossRef] [PubMed]
  86. He, Y.; Thong, P.S.; Lee, T.; Leong, S.K.; Mao, B.Y.; Dong, F.; Watt, F. Dopaminergic cell death precedes iron elevation in mptp-injected monkeys. Free Radic. Biol. Med. 2003, 35, 540–547. [Google Scholar] [CrossRef]
  87. Salazar, J.; Mena, N.; Hunot, S.; Prigent, A.; Alvarez-Fischer, D.; Arredondo, M.; Duyckaerts, C.; Sazdovitch, V.; Zhao, L.; Garrick, L.M.; et al. Divalent metal transporter 1 (dmt1) contributes to neurodegeneration in animal models of parkinson’s disease. Proc. Natl. Acad. Sci. USA 2008, 105, 18578–18583. [Google Scholar] [CrossRef] [PubMed]
  88. Lv, Z.; Jiang, H.; Xu, H.; Song, N.; Xie, J. Increased iron levels correlate with the selective nigral dopaminergic neuron degeneration in parkinson’s disease. J. Neural Transm (Vienna) 2011, 118, 361–369. [Google Scholar] [CrossRef] [PubMed]
  89. Lau, Y.S.; Patki, G.; Das-Panja, K.; Le, W.D.; Ahmad, S.O. Neuroprotective effects and mechanisms of exercise in a chronic mouse model of parkinson’s disease with moderate neurodegeneration. Eur. J. Neurosci. 2011, 33, 1264–1274. [Google Scholar] [CrossRef] [PubMed]
  90. Zaitone, S.A.; Abo-Elmatty, D.M.; Shaalan, A.A. Acetyl-l-carnitine and α-lipoic acid affect rotenone-induced damage in nigral dopaminergic neurons of rat brain, implication for parkinson’s disease therapy. Pharmacol. Biochem. Behav. 2012, 100, 347–360. [Google Scholar] [CrossRef] [PubMed]
  91. Jiang, P.; Dickson, D.W. Parkinson’s disease: Experimental models and reality. Acta Neuropathol. 2018, 135, 13–32. [Google Scholar] [CrossRef] [PubMed]
  92. Rein, M.J.; Renouf, M.; Cruz-Hernandez, C.; Actis-Goretta, L.; Thakkar, S.K.; da Silva Pinto, M. Bioavailability of bioactive food compounds: A challenging journey to bioefficacy. Br. J. Clin. Pharmacol. 2013, 75, 588–602. [Google Scholar] [CrossRef] [PubMed]
  93. Wahlqvist, M.L. Food structure is critical for optimal health. Food Funct. 2016, 7, 1245–1250. [Google Scholar] [CrossRef] [PubMed]
  94. Wu, W.R.; Zhu, X.Z. Involvement of monoamine oxidase inhibition in neuroprotective and neurorestorative effects of ginkgo biloba extract against mptp-induced nigrostriatal dopaminergic toxicity in c57 mice. Life Sci. 1999, 65, 157–164. [Google Scholar] [CrossRef]
  95. Datla, K.P.; Bennett, R.D.; Zbarsky, V.; Ke, B.; Liang, Y.F.; Higa, T.; Bahorun, T.; Aruoma, O.I.; Dexter, D.T. The antioxidant drink effective microorganism-x (em-x) pre-treatment attenuates the loss of nigrostriatal dopaminergic neurons in 6-hydroxydopamine-lesion rat model of parkinson’s disease. J. Pharm. Pharmacol. 2004, 56, 649–654. [Google Scholar] [CrossRef] [PubMed]
  96. Ahlemeyer, B.; Krieglstein, J. Neuroprotective effects of ginkgo biloba extract. Cell. Mol. Life Sci. 2003, 60, 1779–1792. [Google Scholar] [CrossRef] [PubMed]
  97. Kim, M.S.; Lee, J.I.; Lee, W.Y.; Kim, S.E. Neuroprotective effect of ginkgo biloba l. Extract in a rat model of parkinson’s disease. Phytother. Res. 2004, 18, 663–666. [Google Scholar] [CrossRef] [PubMed]
  98. Yang, S.F.; Wu, Q.; Sun, A.S.; Huang, X.N.; Shi, J.S. Protective effect and mechanism of ginkgo biloba leaf extracts for parkinson disease induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. Acta Pharmacol. Sin. 2001, 22, 1089–1093. [Google Scholar] [PubMed]
  99. Ahmad, M.; Saleem, S.; Ahmad, A.S.; Yousuf, S.; Ansari, M.A.; Khan, M.B.; Ishrat, T.; Chaturvedi, R.K.; Agrawal, A.K.; Islam, F. Ginkgo biloba affords dose-dependent protection against 6-hydroxydopamine-induced parkinsonism in rats: Neurobehavioural, neurochemical and immunohistochemical evidences. J. Neurochem. 2005, 93, 94–104. [Google Scholar] [CrossRef] [PubMed]
  100. Guo, S.; Yan, J.; Yang, T.; Yang, X.; Bezard, E.; Zhao, B. Protective effects of green tea polyphenols in the 6-ohda rat model of parkinson’s disease through inhibition of ros-no pathway. Biol. Psychiatry 2007, 62, 1353–1362. [Google Scholar] [CrossRef] [PubMed]
  101. Chen, M.; Wang, T.; Yue, F.; Li, X.; Wang, P.; Li, Y.; Chan, P.; Yu, S. Tea polyphenols alleviate motor impairments, dopaminergic neuronal injury, and cerebral α-synuclein aggregation in mptp-intoxicated parkinsonian monkeys. Neuroscience 2015, 286, 383–392. [Google Scholar] [CrossRef] [PubMed]
  102. Pérez-H, J.; Carrillo-S, C.; García, E.; Ruiz-Mar, G.; Pérez-Tamayo, R.; Chavarría, A. Neuroprotective effect of silymarin in a mptp mouse model of parkinson’s disease. Toxicology 2014, 319, 38–43. [Google Scholar] [CrossRef] [PubMed]
  103. Singhal, N.K.; Srivastava, G.; Patel, D.K.; Jain, S.K.; Singh, M.P. Melatonin or silymarin reduces maneb- and paraquat-induced parkinson’s disease phenotype in the mouse. J. Pineal Res. 2011, 50, 97–109. [Google Scholar] [CrossRef] [PubMed]
  104. Li, X.; Matsumoto, K.; Murakami, Y.; Tezuka, Y.; Wu, Y.; Kadota, S. Neuroprotective effects of polygonum multiflorum on nigrostriatal dopaminergic degeneration induced by paraquat and maneb in mice. Pharmacol. Biochem. Behav. 2005, 82, 345–352. [Google Scholar] [CrossRef] [PubMed]
  105. Datla, K.P.; Zbarsky, V.; Rai, D.; Parkar, S.; Osakabe, N.; Aruoma, O.I.; Dexter, D.T. Short-term supplementation with plant extracts rich in flavonoids protect nigrostriatal dopaminergic neurons in a rat model of parkinson’s disease. J. Am. Coll. Nutr. 2007, 26, 341–349. [Google Scholar] [CrossRef] [PubMed]
  106. Moreira, E.L.; Rial, D.; Aguiar, A.S.; Figueiredo, C.P.; Siqueira, J.M.; DalBó, S.; Horst, H.; de Oliveira, J.; Mancini, G.; dos Santos, T.S.; et al. Proanthocyanidin-rich fraction from croton celtidifolius baill confers neuroprotection in the intranasal 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine rat model of parkinson’s disease. J. Neural Transm (Vienna) 2010, 117, 1337–1351. [Google Scholar] [CrossRef] [PubMed]
  107. Choi, J.G.; Park, G.; Kim, H.G.; Oh, D.S.; Kim, H.; Oh, M.S. In vitro and in vivo neuroprotective effects of walnut (juglandis semen) in models of parkinson’s disease. Int. J. Mol. Sci. 2016, 17. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, Z.; Li, G.; Szeto, S.S.W.; Chong, C.M.; Quan, Q.; Huang, C.; Cui, W.; Guo, B.; Wang, Y.; Han, Y.; et al. Examining the neuroprotective effects of protocatechuic acid and chrysin on in vitro and in vivo models of parkinson disease. Free Radic. Biol. Med. 2015, 84, 331–343. [Google Scholar] [CrossRef] [PubMed]
  109. Gao, X.; Cassidy, A.; Schwarzschild, M.A.; Rimm, E.B.; Ascherio, A. Habitual intake of dietary flavonoids and risk of parkinson disease. Neurology 2012, 78, 1138–1145. [Google Scholar] [CrossRef] [PubMed]
  110. Chan, D.K.; Mellick, G.D.; Hung, W.T.; Woo, J. Genetic and environmental risk factors and their interactions for parkinson’s disease in a chinese population. J. Clin Neurosci. 2003, 10, 313–315. [Google Scholar] [CrossRef]
  111. Checkoway, H.; Powers, K.; Smith-Weller, T.; Franklin, G.M.; Longstreth, W.T.; Swanson, P.D. Parkinson’s disease risks associated with cigarette smoking, alcohol consumption, and caffeine intake. Am. J. Epidemiol. 2002, 155, 732–738. [Google Scholar] [CrossRef] [PubMed]
  112. Tan, E.K.; Tan, C.; Fook-Chong, S.M.; Lum, S.Y.; Chai, A.; Chung, H.; Shen, H.; Zhao, Y.; Teoh, M.L.; Yih, Y.; et al. Dose-dependent protective effect of coffee, tea, and smoking in parkinson’s disease: A study in ethnic chinese. J. Neurol. Sci. 2003, 216, 163–167. [Google Scholar] [CrossRef] [PubMed]
  113. Tan, L.C.; Koh, W.P.; Yuan, J.M.; Wang, R.; Au, W.L.; Tan, J.H.; Tan, E.K.; Yu, M.C. Differential effects of black versus green tea on risk of parkinson’s disease in the singapore chinese health study. Am. J. Epidemiol. 2008, 167, 553–560. [Google Scholar] [CrossRef] [PubMed]
  114. Kandinov, B.; Giladi, N.; Korczyn, A.D. Smoking and tea consumption delay onset of parkinson’s disease. Parkinsonism Relat. Disord. 2009, 15, 41–46. [Google Scholar] [CrossRef] [PubMed]
  115. Mak, J.C. Potential role of green tea catechins in various disease therapies: Progress and promise. Clin. Exp. Pharmacol. Physiol. 2012, 39, 265–273. [Google Scholar] [CrossRef] [PubMed]
  116. Li, F.J.; Ji, H.F.; Shen, L. A meta-analysis of tea drinking and risk of parkinson’s disease. Sci. World J. 2012, 2012, 923464. [Google Scholar] [CrossRef] [PubMed]
  117. Dutta, D.; Mohanakumar, K.P. Tea and parkinson’s disease: Constituents of tea synergize with antiparkinsonian drugs to provide better therapeutic benefits. Neurochem. Int. 2015, 89, 181–190. [Google Scholar] [CrossRef] [PubMed]
  118. Youdim, K.A.; Dobbie, M.S.; Kuhnle, G.; Proteggente, A.R.; Abbott, N.J.; Rice-Evans, C. Interaction between flavonoids and the blood-brain barrier: In vitro studies. J. Neurochem. 2003, 85, 180–192. [Google Scholar] [CrossRef] [PubMed]
  119. Youdim, K.A.; Qaiser, M.Z.; Begley, D.J.; Rice-Evans, C.A.; Abbott, N.J. Flavonoid permeability across an in situ model of the blood-brain barrier. Free Radic. Biol. Med. 2004, 36, 592–604. [Google Scholar] [CrossRef] [PubMed]
  120. de Boer, V.C.; Dihal, A.A.; van der Woude, H.; Arts, I.C.; Wolffram, S.; Alink, G.M.; Rietjens, I.M.; Keijer, J.; Hollman, P.C. Tissue distribution of quercetin in rats and pigs. J. Nutr. 2005, 135, 1718–1725. [Google Scholar] [CrossRef] [PubMed]
  121. Peng, H.W.; Cheng, F.C.; Huang, Y.T.; Chen, C.F.; Tsai, T.H. Determination of naringenin and its glucuronide conjugate in rat plasma and brain tissue by high-performance liquid chromatography. J. Chromatogr. B Biomed. Sci. Appl. 1998, 714, 369–374. [Google Scholar] [CrossRef]
  122. Tsai, T.H.; Chen, Y.F. Determination of unbound hesperetin in rat blood and brain by microdialysis coupled to microbore liquid chromatography. J. Food Drug Anal. 2000, 8, 331–336. [Google Scholar]
  123. Suganuma, M.; Okabe, S.; Oniyama, M.; Tada, Y.; Ito, H.; Fujiki, H. Wide distribution of (3h)(-)-epigallocatechin gallate, a cancer preventive tea polyphenol, in mouse tissue. Carcinogenesis 1998, 19, 1771–1776. [Google Scholar] [CrossRef] [PubMed]
  124. Abd El Mohsen, M.M.; Kuhnle, G.; Rechner, A.R.; Schroeter, H.; Rose, S.; Jenner, P.; Rice-Evans, C.A. Uptake and metabolism of epicatechin and its access to the brain after oral ingestion. Free Radic. Biol. Med. 2002, 33, 1693–1702. [Google Scholar] [CrossRef]
  125. Wang, J.; Ferruzzi, M.G.; Ho, L.; Blount, J.; Janle, E.M.; Gong, B.; Pan, Y.; Gowda, G.A.; Raftery, D.; Arrieta-Cruz, I.; et al. Brain-targeted proanthocyanidin metabolites for alzheimer’s disease treatment. J. Neurosci. 2012, 32, 5144–5150. [Google Scholar] [CrossRef] [PubMed]
  126. Talavéra, S.; Felgines, C.; Texier, O.; Besson, C.; Gil-Izquierdo, A.; Lamaison, J.L.; Rémésy, C. Anthocyanin metabolism in rats and their distribution to digestive area, kidney, and brain. J. Agric. Food Chem. 2005, 53, 3902–3908. [Google Scholar] [CrossRef] [PubMed]
  127. Andres-Lacueva, C.; Shukitt-Hale, B.; Galli, R.L.; Jauregui, O.; Lamuela-Raventos, R.M.; Joseph, J.A. Anthocyanins in aged blueberry-fed rats are found centrally and may enhance memory. Nutr. Neurosci. 2005, 8, 111–120. [Google Scholar] [CrossRef] [PubMed]
  128. Kalt, W.; Blumberg, J.B.; McDonald, J.E.; Vinqvist-Tymchuk, M.R.; Fillmore, S.A.; Graf, B.A.; O’Leary, J.M.; Milbury, P.E. Identification of anthocyanins in the liver, eye, and brain of blueberry-fed pigs. J. Agric. Food Chem. 2008, 56, 705–712. [Google Scholar] [CrossRef] [PubMed]
  129. Zhao, W.; Wang, J.; Bi, W.; Ferruzzi, M.; Yemul, S.; Freire, D.; Mazzola, P.; Ho, L.; Dubner, L.; Pasinetti, G.M. Novel application of brain-targeting polyphenol compounds in sleep deprivation-induced cognitive dysfunction. Neurochem. Int. 2015, 89, 191–197. [Google Scholar] [CrossRef] [PubMed]
  130. Chen, T.Y.; Kritchevsky, J.; Hargett, K.; Feller, K.; Klobusnik, R.; Song, B.J.; Cooper, B.; Jouni, Z.; Ferruzzi, M.G.; Janle, E.M. Plasma bioavailability and regional brain distribution of polyphenols from apple/grape seed and bilberry extracts in a young swine model. Mol. Nutr. Food Res. 2015, 59, 2432–2447. [Google Scholar] [CrossRef] [PubMed]
  131. Passamonti, S.; Vrhovsek, U.; Vanzo, A.; Mattivi, F. Fast access of some grape pigments to the brain. J. Agric. Food Chem. 2005, 53, 7029–7034. [Google Scholar] [CrossRef] [PubMed]
  132. El Mohsen, M.A.; Marks, J.; Kuhnle, G.; Moore, K.; Debnam, E.; Kaila Srai, S.; Rice-Evans, C.; Spencer, J.P. Absorption, tissue distribution and excretion of pelargonidin and its metabolites following oral administration to rats. Br. J. Nutr. 2006, 95, 51–58. [Google Scholar] [CrossRef] [PubMed]
  133. Koutsos, A.; Tuohy, K.M.; Lovegrove, J.A. Apples and cardiovascular health—Is the gut microbiota a core consideration? Nutrients 2015, 7, 3959–3998. [Google Scholar] [CrossRef] [PubMed]
  134. Cardona, F.; Andrés-Lacueva, C.; Tulipani, S.; Tinahones, F.J.; Queipo-Ortuño, M.I. Benefits of polyphenols on gut microbiota and implications in human health. J. Nutr. Biochem. 2013, 24, 1415–1422. [Google Scholar] [CrossRef] [PubMed]
  135. Espín, J.C.; González-Sarrías, A.; Tomás-Barberán, F.A. The gut microbiota: A key factor in the therapeutic effects of (poly)phenols. Biochem. Pharmacol. 2017, 139, 82–93. [Google Scholar] [CrossRef] [PubMed]
  136. Requena, T.; Martínez-Cuesta, M.C.; Peláez, C. Diet and microbiota linked in health and disease. Food Funct. 2018, 9, 688–704. [Google Scholar] [CrossRef] [PubMed]
  137. Selma, M.V.; Beltrán, D.; Luna, M.C.; Romo-Vaquero, M.; García-Villalba, R.; Mira, A.; Espín, J.C.; Tomás-Barberán, F.A. Isolation of human intestinal bacteria capable of producing the bioactive metabolite isourolithin a from ellagic acid. Front. Microbiol. 2017, 8, 1521. [Google Scholar] [CrossRef] [PubMed]
  138. Yuan, T.; Ma, H.; Liu, W.; Niesen, D.B.; Shah, N.; Crews, R.; Rose, K.N.; Vattem, D.A.; Seeram, N.P. Pomegranate’s neuroprotective effects against alzheimer’s disease are mediated by urolithins, its ellagitannin-gut microbial derived metabolites. ACS Chem. Neurosci. 2016, 7, 26–33. [Google Scholar] [CrossRef] [PubMed]
  139. Verzelloni, E.; Pellacani, C.; Tagliazucchi, D.; Tagliaferri, S.; Calani, L.; Costa, L.G.; Brighenti, F.; Borges, G.; Crozier, A.; Conte, A.; et al. Antiglycative and neuroprotective activity of colon-derived polyphenol catabolites. Mol. Nutr. Food Res. 2011, 55 (Suppl. 1), S35–S43. [Google Scholar] [CrossRef] [PubMed]
  140. Liu, W.; Ma, H.; Frost, L.; Yuan, T.; Dain, J.A.; Seeram, N.P. Pomegranate phenolics inhibit formation of advanced glycation endproducts by scavenging reactive carbonyl species. Food Funct. 2014, 5, 2996–3004. [Google Scholar] [CrossRef] [PubMed]
  141. Bookheimer, S.Y.; Renner, B.A.; Ekstrom, A.; Li, Z.; Henning, S.M.; Brown, J.A.; Jones, M.; Moody, T.; Small, G.W. Pomegranate juice augments memory and fmri activity in middle-aged and older adults with mild memory complaints. Evid. Based Complement. Alternat. Med. 2013, 2013, 946298. [Google Scholar] [CrossRef] [PubMed]
  142. Kerr, J.S.; Adriaanse, B.A.; Greig, N.H.; Mattson, M.P.; Cader, M.Z.; Bohr, V.A.; Fang, E.F. Mitophagy and alzheimer’s disease: Cellular and molecular mechanisms. Trends Neurosci. 2017, 40, 151–166. [Google Scholar] [CrossRef] [PubMed]
  143. Ryu, D.; Mouchiroud, L.; Andreux, P.A.; Katsyuba, E.; Moullan, N.; Nicolet-Dit-Félix, A.A.; Williams, E.G.; Jha, P.; Lo Sasso, G.; Huzard, D.; et al. Urolithin a induces mitophagy and prolongs lifespan in c. Elegans and increases muscle function in rodents. Nat. Med. 2016, 22, 879–888. [Google Scholar] [CrossRef] [PubMed]
  144. Gasperotti, M.; Passamonti, S.; Tramer, F.; Masuero, D.; Guella, G.; Mattivi, F.; Vrhovsek, U. Fate of microbial metabolites of dietary polyphenols in rats: Is the brain their target destination? ACS Chem. Neurosci. 2015, 6, 1341–1352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Seeram, N.P.; Zhang, Y.J.; Sartipipour, M.; Henning, S.M.; Lee, R.P.; Harris, D.M.; Moro, A.; Heber, D. Pharmacokinetics and tissue disposition of urolithin a, an ellagitannin-derived metabolite, in mice. FASEB J. 2007, 21, A1081. [Google Scholar]
  146. DaSilva, N.A.; Nahar, P.P.; Ma, H.; Eid, A.; Wei, Z.; Meschwitz, S.; Zawia, N.H.; Slitt, A.L.; Seeram, N.P. Pomegranate ellagitannin-gut microbial-derived metabolites, urolithins, inhibit neuroinflammation in vitro. Nutr. Neurosci. 2017, 1–11. [Google Scholar] [CrossRef] [PubMed]
  147. Thakur, P.; Nehru, B. Inhibition of neuroinflammation and mitochondrial dysfunctions by carbenoxolone in the rotenone model of parkinson’s disease. Mol. Neurobiol. 2015, 51, 209–219. [Google Scholar] [CrossRef] [PubMed]
  148. Dhanasekaran, D.N.; Reddy, E.P. Jnk signaling in apoptosis. Oncogene 2008, 27, 6245–6251. [Google Scholar] [CrossRef] [PubMed]
Table 1. Summary of the neuroprotective effects of polyphenols in PD models.
Table 1. Summary of the neuroprotective effects of polyphenols in PD models.
Compound, Route and DosageSpecies (Sex)NeurotoxinNeuroprotective EffectsReferences
Naringenin (50 mg/kg; p.o.) 4-day pre-treatmentSprague- Dawley rats (male)6-OHDA↑TH-positive cells
↑DA, DOPAC, and HVA
Zbarsky et al. 2005 [14]
Naringenin (70 mg/kg; p.o.) 4-day pre-treatmentC57BL/6 mice (male)6-OHDA↑rotational behavior
↑TH-positive neurons
↑DA, DOPAC, and HVA
↑protein expression of NRf2, HO-1, GCLC, GCLM, HO-1
↑GSH
↓ROS
↑protein expression of JNK and p38
Lou et al. 2014 [15]
Naringin (80 mg/kg; i.p.) 1-day pre-treatment and 6-day post-treatmentC57BL/6 mice (male)6-OHDA↑TH-positive neurons
↓Iba1-positive microglia
↓IL-1β positive microglia
↓6-OHDA—inducedp-4E-BP1
Kim et al. 2016 [16]
Hesperidin (50 mg/kg; p.o.) 28-day post-treatmentaged C57BL/6 mice (female)6-OHDA↓time spent immobile (antidepressant-like activity)
↑spatial learning and memory skill
↑DA, DOPAC and HVA
↑GSH level and TRAP
↑CAT and GPx activity
↓ROS levels
↓6-OHDA—induced GR activity
Antunes et al. 2014 [17]
Hesperetin (50 mg/kg; p.o.) 1-week post-treatmentWistar rats (male)6-OHDA↑rotational behavior
↑motor balance and coordination
↓MDA
↑GSH
↑CAT
↑Bcl-2
↓GFAP expression and DNA fragmentation
Kiasalari et al. 2016 [18]
Quercetin (50 mg/kg; i.g.) 4-day pre-treatmentSprague-Dawley rats (male)6-OHDANo effectsZbarsky et al. 2005 [14]
Quercetin (50–200 mg/kg; i.p.) once a day for one week before and for one week after the 6-OHDA-infusion (100 mg/kg; i.p.) twice a day for one week before and for one week after the 6-OHDA-infusionWistar rats (male)6-OHDANo effectsKääriäinen et al. 2008 [19]
Quercetin (30 mg/kg; i.p.) 14-day post-treatmentSprague-Dawley rats (male)6-OHDA↑DA
↑neuronal density of striatum
↑Nissl-stained neurons
↓PCs
↓LPO
↑GSH
Haleagrahara et al. 2011 [20]
Rutin (25 mg/kg; p.o.) 3-week pre-treatmentWistar rats (male)6-OHDA↑motor co-ordination
↑working performance
↑DA and DOPAC
↓TNF-α and IL-1β
↓NO
↓TBARS, H2O2, and PCs
↑GSH
↑GPx, GR, SOD and CAT activity
Khan et al. 2012 [21]
Troxerutin (150 mg/kg; p.o.) 1-week pre-treatmentWistar rats (male)6-OHDA↑motor function
↑numbers of TH-positive
↑Nissl positive neurons
↑DA and DOPAC
↓MDA, ROS, NO2
↓GFAP
↓TBARS, H2O2, and PCs
↓DNA fragmentation
↑GPx, GR, SOD and CAT activity
Baluchnejadmojarad et al. 2017 [22]
Myricitrin (60 mg/kg; i.p.) 1-day pre-treatment and 6-day post-treatmentC57BL/6 mice (male)6-OHDA↑motor function
↑TH-positive neurons
↓Iba1-positive microglia
↓TNF-α-positive microglia
↑TH activity
↑p-4E-BP1 and p-TH levels
↑mTORC1
Kim et al. 2016 [16]
Myricetin (5 µl of solution 0.5 mg/mL injected into lateral cerebral ventricle) 7-day post-treatmentWistar rats (female)6-OHDA↑TH-positive neurons
↓iron-staining cells
↑DA, DOPAC and HVA
↑TH and GAPDH mRNA
Ma et al. 2007 [23]
Genistein (10 mg/kg; i.p.) a single dose 1 h before surgerySprague–Dawley rats (male)6-OHDA↑rotational behavior
↑Nissl-stained neurons
Baluchnejadmojarad et al. 2009 [24]
Puerarin (0.12 mg/kg; i.p.) 10-day co-treatment tSprague–Dawley rats (male)6-OHDA↑TH-positive cells
↑DA, DOPAC and HVA
↓degeneration of DAargic neurons
↓apoptosis (TUNEL assay)
↓Bax protein expression
↑GDNF protein expression
Zhu et al. 2010 [25]
Baicalein (200 mg/kg; i.g.) 1-week pre-treatmentSprague–Dawley rats (male)6-OHDA↓muscle tremor
↑GABA level
↓ GLu level
↓COI mRNA expression
↓GABA-T protein expression
↑GS protein expression
Yu et al. 2012 [26]
Tangeretin (20 mg/kg; p.o.) 4-day pre-treatmentSprague-Dawley rats (male)6-OHDA↑TH-positive cells
↑DA
Datla et al. 2001 [27]
EGCG (1 and 2 mg/kg; p.o.) 14-day pre-treatmentSprague–Dawley rats (female)6-OHDA↓postural abnormalities
↑ability to cross a narrow beam
Leaver et al. 2009 [28]
Pelargonidin (10 and/or 20 mg/kg; p.o.) 1 day before and on the day of surgery.Wistar rats (male)6-OHDA↑rotational behavior
↑SN neurons
↓TBARS
Roghani et al. 2010 [29]
RES (10, 20 and 40 mg/kg; i.g.) 10-week post-treatmentSprague–Dawley rats (male)6-OHDA↑rotational behavior
↓COX-2 and TNF-α mRNA
↓COX-2 protein
Jin et al. 2008 [30]
RES (20 mg/kg; i.g.) 14-day post-treatment treatmentWistar rats (male)6-OHDA↑rotational behavior
↑total nigral cells and DA neurons
↑total antioxidant capacity
↓number of apoptotic nigral cells
↓ROS
Wang et al. 2011 [31]
RES (20 mg/kg; i.p.) 15-day pre-treatmentWistar rats (male)6-OHDA↑rotational behaviour, motor coordination
↑TH-positive cells
↑DA, DOPAC
↓TBARS, PCs
↓PLA2, COX-2 protein
↑activity of GSH, SOD, CAT, GPx, GR
↑Na+/K+-ATPase activity
Khan et al. 2010 [32]
Piceid (50 mg/kg; p.o.), 14-day co-treatmentSprague-Dawley rats (male)6-OHDA↓motor defects
↓MDA
↑SOD activity
Chen et al. 2015 [33]
CUR (200 mg/kg; i.g.) twice a day, 24-day pre-treatmentWistar rats (female)6-OHDA↑TH-positive cells
↑DA, DOPAC and HVA
↓iron-stained cells
Du et al. 2012 [34]
CUR (200 mg/kg; i.p.) 7-day post-treatmentICR mice (male)6-OHDA↑TH-positive cells
↓GFAP, Iba1 protein
↑SOD1 level
Tripanichkul et al. 2013 [35]
CA (20 mg/kg; p.o.) 3 times per week, 3-week pre-treatmentWistar rats (male)6-OHDA↑locomotor time and distance traveled
↓TBARS
↑GSH
↑GCLC, GCLM, GR, and SOD protein
↓phosphorylated JNK and p38 protein (↓activation)
↑Bcl-2/Bax
↑cleaved caspase-3/caspase-3 and cleaved PARP/PARP
Wu et al. 2015 [36]
SA (20 mg/kg; p.o.) 2-day pre-treatmentWistar rats (male)6-OHDA↑rotational behavior
↑Nissl-stained, TH-positive and total SN neurons
↓iron-stained cells
↓MDA and nitrite level
Zare et al. 2015 [37]
Ellagic acid (50 mg/kg; i.g) 10-day post-treatmentWistar rats (male)6-OHDA↓motor deficiencies
↓IL-1β and TNF-α protein
Farbood et al. 2015 [38]
Delta 9-tetrahydrocannabinol (3 mg/kg; i.p.) 2-week post-treatmentSprague-Dawley rats (male)6-OHDA↑DA
↑TH mRNA and activity
Lastres-Becker et al. 2005 [39]
EGCG (2 and 10 mg/kg; p.o.) 10-day pre-treatmentC57-BL mice (male)MPTP↑DA
↑TH level and activity
↓MPTP-induced activity of SOD and CAT
Levites et al. 2001 [40]
EGCG (2 mg/kg; p.o.) 10-day pre-treatmentC57-BL mice (male)MPTP↑DA
↑TH level and activity
↓α-synuclein
↓Bcl-2, Bax protein
↑PKCα protein
Mandel et al. 2004 [41]
EGCG (25 mg/kg; p.o.) 1-day pre-treatment and 5-day co-treatmentC57B6 mice (male)MPTP↑TH-positive neurons
↑DA, DOPAC and HVA
↑TH activity
↓nNOS mRNAs, nNOS level and activity
Choi et al. 2002 [42]
EGCG (25 mg/kg; p.o.) 7-day post-treatmentC57 black mice (male)MPTP↑rotational behaviour
↑DA and DOPAC
↓PCs
↑ferroportin protein
Xu et al. 2017 [43]
EGCG (10 mg/kg; i.g.) 14-day post-treatmentC57/BL6 mice (male)MPTP↑DA, DOPAC and HVA
↑Ras expression
Reznichenko et al. 2010 [44]
EGCG (10 , 50 mg/kg; i.p.) 10-day pre-treatment and 4-day co-treatmentC57Bl/6 mice (male)MPTP↑TH-positive neurons
↓iNOS expression
Kim et al. 2010 [45]
EGCG (25 , 50 mg/kg/day; i.g.) 1-day pre-treatment and 20-day post-treatmentC57BL/6 mice (male)MPTP↓motor coordination
↑TH-positive neurons
↓CD3+CD4+/CD3+CD8+ T cells
↓TNF-α and IL-6 protein in plasma
Zhou et al. 2018 [46]
Acacetin (10 mg/kg; p.o) 3-day co-treatmentC57BL/6 mice (male)MPTP↓movement impairment
↑TH-positive neurons
↓damage in DAergic cells
↑DA
↓iNOS mRNA and COX-2 mRNA
Kim et al. 2012 [47]
Baicalein (10 mg/kg; i.g.) 5-day co-treatmentC57BL/6 mice (male)MPTP↓motor dysfunction
↓MPTP-induced glutamatergic transmission, presynaptic glutamate release and upregulation of synaptic GluR1 subunit
Xue et al. 2014 [48]
Baicalein (1 and 10 mg/kg; i.g.) 7-day pre-treatmentC57BL/6 mice (male)MPTP↓motor dysfunction
↑TH-positive neurons
↓GFAP, Iba1 protein
↓phosphorylated ERK and JNK-
↓activation
Lee et al. 2014 [49]
7,8-dihydroxyflavone (5, 20, 40, and 100 mg/kg; i.p) 7-day pre-and 7-day co-treatmentB57/BL mice (male)MPTP↑TH-positive neurons
↑TrkB activity
↓caspase-3 protein
Jang et al. 2010 [50]
Tangeretin (10 mg/kg; i.p.) 4-day pre-treatmentC57BL/6 mice (male)MPTP↑TH-positive neurons
↑DA
↑GRP-78 protein
Takano et al. 2007 [51]
Nobiletin (10 mg/kg; i.p.) 1-day pre-treatment and 6-day-post-treatmentSprague Dawley (SD) rats (female)MPTP↑TH-positive neurons
↑DA
↓Iba1 protein and IL-1β
↑GDNF protein
Jeong et al. 2015 [52]
Nobiletin (50 mg/kg; i.p.) 14-day post-treatmentC57BL/6 mice (male)MPTP↓motor and cognitive impairment
↑Ca2+/calmodulin-dependent protein kinase II (CaMKII)
↑DARPP-32, dopamine- and cAMP-regulated phosphoprotein-32
Yabuki et al. 2014 [53]
Naringin (8 and 800 mg/kg; i.p) 1-day pre-and 6-day post-treatmentSprague Dawley (SD) rats (female)MPTP↑TH-positive neurons
↑GDNF protein
↑mTORC1 activity
↓TNF-α protein
Leem et al. 2014 [54]
Kaempferol (25, 50 and 100 mg/kg; p.o) 14-day pre-treatmentC57BL/6 mice (male)MPTP↑Motor behavioral
↑TH-positive neurons
↑DA and DOPAC
↓MDA
↑activity of SOD and GPx
Li and Pu 2011 [55]
Quercetin (100, 200 mg/kg; p.o.) 10-day pre-treatment and 4-day co-treatmentC57BL/6 mice (male)MPTP↑motor balance and coordination
↑DA
↑GPx, SOD activity
↑Na+, K+-ATPase
↓4-HNE
↑AChE activity
Lv et al. 2012 [56]
Morin (5, 20, 40, and 100 mg/kg; i.p) 5-day pre-and 5-day co-treatmentB57/BL mice (male)MPTP↓cataleptic time (bar test) and
↑ number of steps (drag test)
↑DA
↑TH-positive neurons
Zhang et al. 2010 [57]
Silibinin (1 or 10 mg/kg; i.p.) 5-day co-treatmentC57BL/6 mice (male)MPTP↓Motor dysfunction
↓DAergic neuron damage
Lee et al. 2015 [58]
Silibinin (50, 100 mg/kg; i.p.) 5-day co-treatmentSprague Dawley (female)MPTP↑TH-positive neurons
↓IL-1β, TNF-α and iNOS protein
Jung et al. 2014 [59]
Silibinin (100, 200 mg/kg; i.g.) 7-day co-treatmentalbino rats of Charles– Foster strain (male)MPTP↑spatial memory and locomotor activity
↓mitochondrial complex-I and IV activity
↑mitochondrial complex-II and V activity
↑mitochondrial membrane potential
↓NO, MDA
↑SOD activity
↓caspase-3 and NFκB protein
Geed et al. 2014 [60]
Genistein (10 mg/kg; i.p.) 3-day pre-treatment and 5-day-co-treatmentC57BL/6 mice (male)MPTP↑TH-positive neurons
↑DA, DOPAC and HVA
↑TH and DAT mRNA
↓Bcl-2 mRNA
Liu et al. 2008 [61]
Gastrodin (10, 30 and 60 mg/kg; p.o.) 15-day co-treatmentC57BL/6 mice (male)MPTP↓bradykinesia and motor impairment
↑TH protein
↓GFAP protein
↑Bcl-2 protein
↓Bax and caspase-3 protein
↓PARP cleavage
Kumar et al. 2013 [62]
Tanshinone I (5 and 10 mg/kg; i.g.) 7-day pre- and co-treatmentC57BL/6 mice (male)MPTP↓Motor dysfunction
↑DA, DOPAC and HVA
↑TH-positive neurons
↓Iba1-positive activated microglia cells
↓TNF-α and IL-10 protein
Wang et al. 2015 [63]
Ferulic acid (40 mg/kg; i.g.) 3-day pre-treatment and 4-day co-treatmentC57BL/6 mice (male)MPTP↑motor balance and coordination
↓anxiety
↓degeneration of DAargic neurons
↓Bax/Bcl-2 ratio
Nagarajan et al. 2015 [64]
Madecassoside (15, 30 and 60 mg/kg; i.g) 7-day pre-treatment and 14-day co-treatmentWistar rats m (male)MPTP↑limb coordination and limb placing
↑DA, DOPAC and HVA
↓MDA
↑GSH
↑Bcl-2/Bax ratio
↑BDNF protein
Xu et al. 2013 [65]
Curcumin (50 mg/kg; i.p.) 3-times at time points 1, 3, and 7 h post first MPTP injectionICR mice (male)MPTP↑DA
↑density of DA neurons
Vajragupta et al. 2003 [66]
Curcumin (80 mg/kg; i.p.) 7-day co-treatmentSwiss albino mice (male)MPTP↑DA and DOPAC
↓MAO B activity
Rajeswari et al. 2008 [67]
Curcumin (150 mg/kg; p.o.) 1-week pre-treatmentC57BL/6 mice (male)MPTP↑motor performance (rotarod test)
↓steep reduction intotal ambulation time (open field test)
↑DA, DOPAC and HVA
↓GFAP over-expression
↓IL-6 and TNF-α protein
↓iNOS protein and NO content
Ojha et al. 2012 [68]
Ginsenoside Rg1 (5, 10 and 20 mg/kg; i.p.) 3-day pre-treatment and 5-day co-treatmentC57-BL mice (male)MPTP↑numbers of TH-positive and Nissl positive neurons in SN
↓TUNEL positive neurons in SN
↑GSH
↓MPTP-induced SOD activity
↓MPTP-induced phosphor—JNK and phospho-c-Jun level
Chen et al. 2005 [69]
Piceid (100 and 200 mg/kg; p.o.) 7-day co-treatment and 7-day post-treatmentC57BL/6 mice (male)MPTP↑locomotor activity
↑TH-positive neurons
↑SOD activity
↓MDA level
↑p-Akt expression
↓activated caspase 3 expression
Chen et al. 2015 [33]
TF (10 mg/kg; i.g.) 35-day co-treatmentC57BL/6 mice (male)MPTP/p↑Locomotor activity
↑TH- and DAT-positive neurons
↓caspase-3, 8, and 9 protein
Anandhan et al. 2012 [70]
TF (10 mg/kg; i.g.) 3-day pre-treatment and 4-day co-treatmentC57BL/6 mice (male)MPTP↑Locomotor activity
↑DAT-positive neurons
↓TBARS, ↑ GSH
↑SOD, CAT and GPx activity
↓MAO-B activity
Anandhan et al. 2012b [71]
TF (10 mg/kg; i.g.) 35-day co-treatmentC57BL/6 mice (male)MPTP/p↓Akinesia, catalepsy
↓IL-1β, TNF-α, IL-6 as well as MPTP-induced IL-4 and IL-10
↓GFAP and COX-2 protein
↑Bcl-2 protein
↓Bax protein
Anandhan et al. 2013 [72]
Apigenin (5, 10 and 20 mg/kg; p.o.), luteolin (10 and 20 mg/kg; p.o.) 5-day pre-treatment and 21-day co-treatmentSwiss-albino mice (male)MPTP/p↑Locomotor activity
↑TH-positive neurons
↓MDA, ↑GSH
↑SOD and CAT activity
↓GFAP and TNF-α protein
↑BDNF protein
↓MAO-B activity
Patil et al. 2014 [73]
SA (20 mg/kg; i.g.) 35-day co-treatmentC57BL/6 mice (male)MPTP/p↓impairment of motor coordination
↑TH- and DAT-positive neurons
↑TH, DAT and VMAT2 protein
↑DA, DOPAC and HVA
↓TBARS, ↑ GSH
↑SOD, CAT and GPx activity
↓IL-1β, TNF-α protein
↓COX-2 protein
Rekha et al. 2014 [74]
Quercetin (25, 50, 75 mg/kg; i.p.) 4-day post-treatmentSprague Dawley rats (male)ROT↑rotational behavior
↑TH-positive neurons
↓TUNEL positive neurons
↑DA
↑complex-I activity
↑GSH, GSSG
↓ROT-induced activity of SOD and CAT
Karuppagounder et al. 2013 [75]
CUR (20 mg/kg; i.g.) 50-day co-treatmentLewis rats (male)ROT↓postural impairment
↑motor coordination
↑TH activity
↓MDA, ↑ GSH
↑HO-1 and NQO1 protein
↑Akt/Nrf2 phosphorylation (activation)
Cui et al. 2016 [76]
Piceid (80 mg/kg; i.g.) 5-week co-treatmentSprague Dawley rats (male)ROT↓postural impairment
↑motor coordination
↑TH-positive neurons
↑ATP
↑Trx, GSH
↑SOD activity
Chen et al. 2015 [33]
FA (80 mg/kg; i.g.) 4-week co-treatmentWistar rats (male)ROT↑TH-positive neurons
↑SOD and CAT activity
↓MDA, ↑ GSH
↓GFAP and Iba1 protein
↓IL-1β, TNF-α, IL-6 protein
↓COX-2 and iNOS protein
Ojha et al. 2015 [77]
Sesamol (15 mg/kg; p.o.) naringenin(10 mg/kg; p.o.) 10-day post-treatmentWistar rats (male)ROT↑motor coordination
↑body weight
↓degenerated neurons
↑TH and ubiquitin expression
↑parkin, CHIP, and DJ1 expression
↓caspase-9 and 3 protein
↑Hsp70 and Hsp90 protein
Sonia Angeline et al. 2013 [78]
EGCG (10 mg/kg; i.p.) 24 h before LPS and continues for 7 daysSprague Dawley rats (male)LPS↑TH-positive neurons
↑DA
↓TNF-α and nitrite
AL-amri et al. 2013 [79]
CUR (50 mg/kg; i.p.) 10-day treatment beginning 5 days prior to HcyLewis rats (male)Hcy↑Locomotor Activity
↑Nissl-stained neurons
↑Bcl-2 protein and ↓ Bax protein
↓Bax/Bcl-2
Mansouri et al. 2012 [80]
↑ = increase, ↓ = decrease.

Share and Cite

MDPI and ACS Style

Kujawska, M.; Jodynis-Liebert, J. Polyphenols in Parkinson’s Disease: A Systematic Review of In Vivo Studies. Nutrients 2018, 10, 642. https://doi.org/10.3390/nu10050642

AMA Style

Kujawska M, Jodynis-Liebert J. Polyphenols in Parkinson’s Disease: A Systematic Review of In Vivo Studies. Nutrients. 2018; 10(5):642. https://doi.org/10.3390/nu10050642

Chicago/Turabian Style

Kujawska, Małgorzata, and Jadwiga Jodynis-Liebert. 2018. "Polyphenols in Parkinson’s Disease: A Systematic Review of In Vivo Studies" Nutrients 10, no. 5: 642. https://doi.org/10.3390/nu10050642

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop