Open access peer-reviewed chapter

Polyphenols as Potential Therapeutic Drugs in Neurodegeneration

Written By

Patrizia Polverino de Laureto, Luana Palazzi and Laura Acquasaliente

Submitted: 08 May 2019 Reviewed: 06 September 2019 Published: 23 October 2019

DOI: 10.5772/intechopen.89575

From the Edited Volume

Neuroprotection - New Approaches and Prospects

Edited by Matilde Otero-Losada, Francisco Capani and Santiago Perez Lloret

Chapter metrics overview

1,110 Chapter Downloads

View Full Metrics

Abstract

Several therapeutic approaches have been suggested so far for the treatment of neurodegenerative diseases, but to date, there are no approved therapies. The available ones are only symptomatic; they are employed to mitigate the disease manifestations and to improve the patient life quality. These diseases are characterized by the accumulation and aggregation of misfolded proteins in the nervous system, with different specific hallmarks. The onset mechanisms are not completely elucidated. Some promising approaches are focused on the inhibition of the amyloid aggregation of the proteins involved in the etiopathology of the disease, such as Aβ peptide, Tau, and α-synuclein, or on the increase of their clearance in order to avoid their aberrant accumulation. Here, we summarize traditional and new therapeutic approaches proposed for Alzheimer’s and Parkinson’s diseases and the recent technologies for brain delivery.

Keywords

  • Alzheimer’s disease
  • Parkinson’s disease
  • amyloidosis
  • protein fibrils and oligomers
  • polyphenols
  • brain delivery technologies

1. Introduction

Alzheimer’s disease (AD) and Parkinson’s disease (PD) are the most common neurodegenerative disorders. They are multifactorial, progressive, age-related, and influenced by genetic and environmental factors. Despite being public health problems and widely studied, there are no effective treatments. The therapies in use at the moment are only symptomatic and focused to ameliorate patients’ life quality. Moreover, there are no diagnostic methods for the early detection of these diseases that, especially at the onset, share some pathological hallmarks. There are specific proteins associated with the diseases, but it is still unclear when and how they lose their functionality and become toxic. Several pathways of cellular dysfunction have been described to explain the toxicity associated with the disease, but the pathological role of proteins involved still remains controversial. Currently, the most promising therapeutic approaches are focused on personalized treatments and targeted drugs.

Here, we summarize some relevant features of the new proposed therapies for AD and PD. In the last decade, renewed interest rises toward alternative pharmacological treatments and products of natural origin, especially those associated with the Mediterranean diet, such as polyphenols. The unexpected benefits and the wide-range properties of polyphenols suggest deepening the study of these molecules for a more comprehensive understanding of their mechanism of action in order to use them in effective therapies.

Advertisement

2. Molecular aspects of Alzheimer’s and Parkinson’s diseases

2.1 Brief introduction to AD and PD

AD is characterized by the gradual decline in the cognitive function, memory loss, and behavior changes [1]. Typical features of the disease are a synaptic deficit in the neocortex and the limbic system, neuronal loss, white matter loss, astrogliosis, microglial cell proliferation, and oxidative stress [2]. The major areas of the human brain affected by AD are schematically represented in Figure 1. The pathological hallmarks of AD are the presence of intracellular flame-shaped neurofibrillary tangles and extracellular plaques in the brain. The tangles are especially present in the perinuclear cytoplasm and are prevalently formed by the Tau protein, in a hyperphosphorylated form. The plaques derive from the progressive accumulation of amyloid β-peptide (Aβ) in a filamentous form [3]. The neuritic plaques have a diameter ranging from 10 to more than 120 μm [2]. The methods used for the diagnosis of the pathology have been standardized. They refer to the density and the grade of compactness of the neuritis amyloid plaques and neurofibrillary tangles [4]. AD aggregates can be classified into positive and negative lesions as a function of their localization and level of progression [5]. Typical positive lesions are represented by amyloid plaques and neurofibrillary tangles, neuropil threads, and dystrophic neurites, essentially formed by hyperphosphorylated Tau [6]. The negative lesions provide loss of neurons and neuropil threads [7].

Figure 1.

Affected brain regions in AD and PD. Cross-section of human brain showing the principal districts affected by AD (green) and PD (blue). AD typically involves parts of the brain involved in memory, like hippocampus and ventricles, and the cerebral cortex responsible for language. In PD nerve cells of the motor cortex and in part of the basal ganglia (composed by substantia nigra, putamen, caudate nucleus, globus pallidus, and locus coeruleus) degenerate. As a result, the basal ganglia cannot control muscle movement as it normally does, leading to tremor, bradykinesia, and hypokinesia.

Clinically, PD typically manifests with motor symptoms, such as bradykinesia, rigidity, tremor at rest, and instability. Since there is no definitive test for the diagnosis of PD, the appearance of these clinical manifestations is important for the early treatment of the disease [8]. PD is characterized by the loss of dopaminergic neurons in the Substantia nigra pars compacta (Figure 1) and by the deposition of intraneuronal proteinaceous aggregates, mainly composed by α -synuclein (Syn), named Lewy bodies and Lewy neurites [9]. Syn was also found in the pathological inclusions of Lewy body variant of both AD and multiple system atrophy. Furthermore, Syn inclusions characterize other neurodegenerative diseases, defined as α-synucleinopathies, including Down’s syndrome, progressive autonomic failure, and familial and sporadic AD [10]. In a very recent study, Shahmoradian and coll. have reported that Lewy bodies are not only formed by Syn deposit but also by clusters of lipid vesicles [11]. These important findings further correlate Syn-lipid interaction with neurodegeneration [12, 13].

AD and PD are generally sporadic and occur in individuals between ages 60 and 70, but the ~20% of patients have a genetically linked familial form. The onset of these forms occurs earlier, and it is associated with mutations in several genes [14]. The main mutations are listed in Table 1. The proteins involved in such neurodegenerative diseases, Aβ, tau, and Syn, are completely distinct in terms of structure and putative functions, most of which are not completely clarified. However, the formation of aggregated structures is a common feature among these macromolecules. Fibrils, which originate from the association of monomeric forms of the proteins, pass through intermediate species such as oligomers (Figure 2). Generally, they can cross the membrane and spread throughout the brain. Several evidences suggest that oligomers are the species responsible for the cytotoxicity. There are many proofs in support of this hypothesis, but unfortunately, due to the extreme heterogeneity in oligomer structures and their transient nature, a conclusive view has not been obtained yet [31, 32, 33]. The atomic structure of fibrils has been studied by several biophysical techniques. A quite accepted hypothesis agrees with the presence of a common molecular organization independent from the original structure of the involved protein: repetitive β-sheet units parallel to the fibril axis with their strands perpendicular to it [34, 35]. Amyloid fibrils can self-assemble in vitro from many structurally different proteins and peptides, not necessarily involved in diseases. It has been postulated that the cross-β structure represents a generic conformation, which represents another folding state for proteins [36, 37]. In addition to these characteristics, there are also some common aspects in the onset of the diseases. Several studies suggest possible interplays and synergistic activities between the involved proteins. Clinton et al. [38] provided evidence that Aβ, tau, and Syn could interact in vivo to promote their self-aggregation, thus accelerating the cognitive dysfunction [38]. High levels of Syn were found in patients suffering from AD [39]. Aβ stimulates Syn fibril formation in the transgenic mouse model through a seed mechanism [40]. In another study, Syn seems to inhibit the deposition of Aβ into the amyloid plaques [41].

Disease Mutated protein Phenotype Notes Refs
AD APP Abnormal production of Aβ www.molgen.ua.ac.be/ADMutations [15]
ApoE Increase of the density of Ab plaques
High risk of AD, late onset of AD and Down syndrome
www.molgen.ua.ac.be/ADMutations [16]
Presenilin1 Increased the Aβ42/Aβ40, and reduced γ-secretase activity >200 mutations [17, 18]
Presenilin2 Increased the Aβ42/Aβ40, and reduced γ-secretase activity Rare, <40 mutations [19]
PD Syn Familiar and early onset PD A53T; A30P, E46K, G51D, H50Q , gene duplication and triplication [20, 21, 22, 23, 24, 25]
Leucine-rich repeat kinase 2 (LRRK2) Autosomal dominant PD; mid-to-late onset and slow progress >20 mutations [26, 27]
E3 ubiquitin ligase Parkin Early-onset PD and parkinsonism >150 mutations, deletions, insertions [28]
PINK1 Sporadic early-onset Parkinsonism >60 mutations [29]
DJ-1 Autosomal recessive PD >10 mutation, deletions [30]

Table 1.

Main mutations involved in familiar forms of AD and PD.

Figure 2.

Scheme of the aggregation process of amyloid proteins. The formation of fibrils occurs through a nucleation-dependent pathway starting from the monomeric form of the protein and leading to fibril elongation through intermediates (oligomers and protofibrils). The formation of the nucleus is the rate limiting step, and at this stage, the protein has acquired an aggregation-prone conformation. Fibrils are composed of a β-sheet structure in which hydrogen bonding occurs along the length of the fibril, and the β-strands run perpendicular to the fibril axis.

2.2 Key proteins in neurodegeneration

2.2.1 Aβ peptide

The Aβ peptide was found in the amyloid plaques in 1984 [3]. Aβ represents a group of peptides constituted by 37–49 residues (Figure 3A), derived from the proteolytic processing of the amyloid precursor protein (APP) [42, 43] (Figure 4). APP is a single membrane-spanning domain protein, containing a large extracellular glycosylated N-terminus and a shorter cytoplasmic C-terminus. The enzymatic processes responsible for the release of Aβ from APP are to date well elucidated [2]. Specifically, APP undergoes several proteolytic cleavages. The processing by α-secretase results in the release of the large fragment sAPPα in the lumen, and the C-terminal fragment (CTF83) remains in the membrane. Two membrane endoproteases β- and γ-secretase sequentially hydrolyze APP. Firstly, APP releases sAPPβ by the action of β-secretase in the extracellular space. A fragment of 99 amino acids, CTFβ, remains bound to the membrane. CTFβ is successively and rapidly processed by γ-secretase generating Aβ. A precise cleavage site was not defined; therefore, Aβ is characterized by heterogeneity at the C-terminal and the peptide can end at position 40 (Aβ40) with a high frequency of occurrence (~80–90%) or at position 42 (Aβ42, ~5–10%). It is well established that Aβ42 generally generates fibrils more quickly than Aβ40 [44]. The production of Aβ is a normal metabolic event; in fact, these species are found in the cerebrospinal fluid and the plasma in healthy subjects [45]. Their abnormal accumulation, deriving from an imbalance between the production and clearance of these peptides, is associated with the pathogenesis of AD. Monomer, oligomer, and fibril forms of Aβ are differently involved in the onset of AD. The most common hypothesis is the Aβ-amyloid cascade [46]. The overproduction or the reduced clearance of Aβ leads to the deposition of fibrillar Aβ in the brain, determining synaptic and neuronal toxicity and thus neurodegeneration. There are many evidences in support of the so-called Aβ-amyloid oligomer hypothesis [31]. The proteolytic degradation of Aβ is a major route of clearance. Neprilysin (NEP) is considered one of the most important endopeptidase for the control of cerebral Aβ levels [47, 48] and for the degradation of some vasoactive peptides including natriuretic peptides and neuropeptides. Aβ clearance is mediated by other proteolytic enzymes such as apolipoprotein E (apoE) [49] and by autophagy [50]. Reduced activity of the clearance enzymes, which could be caused by aging, can contribute to AD development by promoting Aβ accumulation.

Figure 3.

Sequence and structural domain organization for Aβ (A), tau (B), and Syn (C). For Aβ, the residues 12–24 and 30–40 involved in the formation of a cross-β fibril structure are highlighted and connected by dashed lines. In (B), the longest isoform (441 residues) of tau is shown, where N indicates the possible N-terminal insertion defining other isoform, PRR, the proline-rich region, target of phosphorylation (P), and MTBR, the microtubule binding region that can contain three or four repeats (R), and other phosphorylations (P) occur at the C-terminal. In the case of Syn (C), the N- and C-terminals and NAC domains are shown, as well as the position of the mutations responsible for familiar form of PD. Residues 1–95 form the lipid-binding region.

Figure 4.

Scheme of metabolism of APP and accumulation of the Aβ peptide. Aβ1–40/42 peptides are released from APP by the action of two membrane endoprotease β- and γ-secretases. Firstly, APP releases sAPPβ by the action of β-secretase in the extracellular space, and a fragment of 99 amino acids, CTFβ, remains bound to the membrane. CTFβ is successively and rapidly processed by γ-secretase generating Aβ peptides. Under physiological conditions, Aβ1–40/42 are degraded by enzymatic clearance processes. The proteolytic pathway mediated by α-secretase is also shown.

The secondary and tertiary structure of Aβ in solution has been studied by several biophysical techniques. These conformational studies are difficult for the protein high tendency to aggregate in solution. However, Aβ seems to populate distinct states in solution and to adopt a collapsed-coil structure, as deduced by NMR studies [51, 52]. Aβ preferentially binds to negatively charged lipids and acquires α-helical structure in the presence of membranes, membrane-like systems, and fluorinated alcohols [53, 54]. In the presence of phospholipids, Aβ undergoes conformational transition and forms β-sheets [55, 56]. Oligomeric Aβ binds to membranes with high affinity. Upon interaction, a membrane damage can occur as causative of the cellular toxicity [57]. It seems that especially oligomeric Aβ can disrupt the membrane bilayer by a detergent mechanism [58].

2.2.2 Tau

Tau is a neuronal protein associated with the microtubules [59]. Six Tau isoforms, which differ only in their primary structure, were detected in the human brain and central nervous system (Figure 3B), while in the peripheral nervous system other Tau isoforms were also found [60]. The longest isoform contains 441 residues and the shortest 352 residues [61]. Depending on the isoform, the N-terminal can contain 0, 1, or 2 inserts (N). The protein appears largely post-translational modified, especially in terms of phosphorylation (P). Other modifications are acetylation, deamidation, methylation, glycosylation, or ubiquitination [59]. Tau proteins are also subjected to proteolytic degradation that seems to be correlated with AD [62]. The region PRR (proline-rich region) contains the main sites of phosphorylation. Although all the post-translational modifications seem to contribute to the physiological and pathological properties of Tau, the signaling cascades and the effect on protein kinases and phosphatases are not completely clarified yet. The region 244–369 (microtubule binding region, MTBR) is responsible for the binding to the microtubule and contains three or four repeats (R1-R4). Physiologically, Tau stabilizes the microtubule through MTBR, and such binding is modulated by the coordinated actions of kinases and phosphatases. Structurally, Tau belongs to the intrinsically disordered proteins, lacking a well-defined secondary and tertiary structure [59] and can interact with several other proteins. Upon aggregation, Tau can form dimers, oligomers, and larger polymers. In such aggregates, cysteine residues may play an important role [63]. Similarly, to other proteins involved in neurodegeneration, the oligomeric forms have a cytotoxic effect and might be involved in the Tau-related pathogeneses [64]. In neurofibrillary tangles, Tau forms the so-called paired helical filaments (PHFs) and straight filaments (SFs) [65, 66]. In PHF, Tau is ∼three to four-fold more hyperphosphorylated than in the normal brain. The Tau filaments exhibit the typical cross-β structure found in other types of fibrils [67].

2.2.3 α-Synuclein (Syn)

Syn is a small protein (14.4 kDa) mainly expressed in pre-synaptic nerve terminals of the central nervous system and very abundant in erythrocytes and platelets [68]. Despite the intensive investigation and the discovery that the protein plays a central role in synaptic transmission and vesicle recycling [69], the complete Syn biological function remains still elusive. Syn may control the neurotransmitter release, promoting the formation and assembly of the SNARE complex [70, 71]. Syn structure could be divided into three main domains: N-, central, and C-terminals (Figure 3C). The N-terminal region (amino acids 1–60) contains seven imperfect repeats, with a hexameric consensus motif (KTKGEV). All the known missense mutations of Syn, responsible for the familiar forms of PD, are located in this region (Table 1). The central hydrophobic domain (amino acids 61–95) is known as the non-amyloid-β component of AD amyloid plaques (NAC). It is responsible for Syn amyloid aggregation [72]. N-terminal and NAC domains together (amino acids 1–95) mediate the interaction of Syn with lipids, membranes, and fatty acids [73]. The C-terminal domain (amino acids 96–140) is an acidic, negatively charged, highly soluble, and disordered tail, target of post-translational modifications. This region plays a series of important roles, modulates Syn binding to membrane and metals, Syn aggregation and its protein-protein interaction properties. The deletion of this domain increases the aggregation rate of Syn in vitro and in cells [74].

Syn is the prototype of the natively unfolded proteins, but adopts a stable secondary structure as a function of the environment [75]. Multiple studies have demonstrated that Syn is more compact than expected for a random coil due to long-range interactions between the C-terminal tail and the NAC domain as well as electrostatic interactions between the N terminus and the C terminus [76]. Syn is supposed to populate different conformers in solution and can undergo conformational transition as a function of the environment and/or upon binding. The extreme Syn conformational flexibility is responsible for its multifunctional properties, its capability to adopt different conformations, and to interact with different systems and other proteins [77]. For example, the interaction of Syn with negatively charged membranes, vesicles, bilayers, and lipids in general has important physiological consequences [78, 79], corroborating the hypothesis that Syn functions are correlated with lipids [80].

Advertisement

3. Overview of recent therapeutic approaches in Alzheimer’s and Parkinson’s diseases

3.1 Traditional ongoing therapies

Current pharmacological therapies (Table 2) for neurodegenerative diseases focus to ameliorate the life conditions of patients and are generally only palliative. Since in many cases, the aberrant deposition of the protein strongly contributes to the toxicity associated with the diseases, some treatments are currently thought to target such specific proteins (i.e., Syn and Aβ) in order to restore their correct physiological levels in vivo. Given the complexity in the onset and progression of these diseases, treatments should be customized and tailored to the individual needs of the patients.

Table 2.

Current available drugs for the treatment of AD and PD.

In the case of AD, a therapy based on the use of cholinesterase inhibitors (ChEIs) and the N-methyl-d-aspartate (NMDA) antagonist is currently available and Food and Drug Administration (FDA)-approved. In particular, three ChEIs are used: donepezil, rivastigmine, and galantamine [81]. The aim is to increase the levels of acetylcholine, a neurotransmitter responsible for memory and cognitive function, by reducing its enzymatic breakdown. Another class is represented by NMDA receptor antagonists, such as memantine, a noncompetitive antagonist, capable to block the effects of the excitatory neurotransmitter glutamate [82]. There are a series of molecules under study referred to as “disease-modifying” drugs. They should interfere with key steps in AD development, including the deposition of Aβ plaques and neurofibrillary tangle formation, inflammation, oxidative damage, iron deregulation, and cholesterol metabolism. Many drugs are proposed for their ability to alleviate behavioral symptoms of AD. A few examples include antidepressants, such as escitalopram and mirtazapine, anticonvulsants, that is, carbamazepine and levetiracetam, mood stabilizers, and stimulants, such as methylphenidate [83].The treatments for PD are still based on dopaminergic drugs, such as levodopa, the precursor of dopamine [84]. Long-term use of levodopa determines the development of motor problems. In association with levodopa, a decarboxylase inhibitor is administered to prevent some side effects. PD therapy involves the use of dopamine agonists, such as ropinirole or rotigotine, monoamino oxidase B inhibitors, such as rasagiline and selegiline, and catechol-O-methyltransferase (COMT) inhibitors, which can reduce the metabolism of endogenous dopamine.

3.2 New generation therapies

Novel experimental approaches are under investigation and the most promising have as a target the protein involved in the diseases. The stages of intervention could be at the level of the protein synthesis or clearance and at the level of protein aggregation or propagation of the toxic species or their precursors (Figure 5).

  1. Control of the protein concentration in vivo. To reduce the production of Aβ, Tau, and Syn, the RNA interference approach is to date quite attractive [85, 86, 87]. It is based on the idea to inhibit specific protein expression by activating a sequence-specific RNA degradation process. This technology results useful to study gene function, investigate the mechanism of the disease, and validate drug targets. Of course, the suppression of the target protein might have negative implications, due to the alteration of its physiological equilibrium. Additionally, the transcription of the gene can be reduced. Clenbuterol was shown to be efficient in reducing Syn expression by 35% in neuroblastoma cell lines [88]. Some AD therapies based on the modulation of AD gene expression are proposed on the basis of the important progresses made in the understanding of the transcriptional regulation of some enzymes such as beta-secretase 1 (BACE1), apolipoprotein E (apoE), APP amyloid precursor protein (APP), and presenilin (PSEN) promoters [89]. Alternatively, to reduce the level of the active protein in vivo, its clearance can be enhanced. This can be obtained by increasing the intracellular degradation via autophagy or via the ubiquitin system. This topic is excellently reviewed by Boland et al. [90].

  2. Protein aggregation inhibitors. An attractive approach would be the use of small molecules able to bind the monomeric form of the protein preventing its assembly into potentially toxic aggregates. Unfortunately, it remains still unclear which conformation of these proteins must be targeted, since all of them are natively unfolded, and multiple and concurrent events contribute to their conversion in oligomers and fibrils [91]. In this ambit, the use of polyphenols is quite promising, and, as described below, these compounds exhibit in some cases the ability to disaggregate preformed oligomers and fibrils [92].

Figure 5.

New generation therapies in AD and PD. Potential levels of intervention to counteract the abnormal accumulation of the amyloidogenic proteins and restore their physiological concentration, which results from a balance between the rates of synthesis, clearance, aggregation, and propagation.

Advertisement

4. Effect of polyphenol compounds in neurodegeneration

4.1 Natural polyphenol products

Polyphenols are natural compounds, generally secondary metabolites, produced by plants and found mainly in fruits, vegetables, and cereals and in their derivatives. Some of them are synthetized during the normal development of the plant while others are produced in response to stress stimuli [93, 94]. They exert their function acting during the phase of development, reproduction, nutrition, growth, and communication with other plants, as well as in plant defense mechanisms like resistance to microbial pathogens, herbivore, insects, and protection to UV-light radiation [95]. More than 8.000 polyphenols have been identified in different plant species. They all derive from common precursors like phenylalanine and shikimic acid [96]. Often, they are linked with a sugar through the hydroxyl moiety, directly to the aromatic ring or conjugated with other compounds [97]. Polyphenols are characterized by a minimal hydroxyphenyl structure, and despite the multitude of existing polyphenols, they are grouped into different classes according to the number of phenol rings. The main groups are phenolic acids, flavonoids, stilbenes, and lignans [98] (Figure 6).

Figure 6.

Scheme of the main polyphenols and their chemical structures. Polyphenols are grouped into four principal classes: stilbenes, lignans, phenolic acids, and flavonoids. The last one is organized into six subclasses: anthocyanins, flavonols, flavanols, flavanones, chalcones, and others.

4.2 Potential therapeutic applications of polyphenols

Several epidemiological studies have been reported concerning the potentiality of polyphenols compounds in disease treatment and prevention [99, 100]. Polyphenols exert a positive role in cardiovascular disease [101, 102, 103], diabetes [104, 105], cancer [106, 107], aging, and neurodegeneration [108, 109]. One of the main activities of polyphenol resides is their antioxidant properties. Indeed, they are capable to protect cells and macromolecules from oxidative damage which in turn leads to degenerative age-associated diseases [110, 111]. Nevertheless, polyphenol function is also bound to its action on enzymes, immune defense, inflammation, cell signaling, and other pathways critical for the onset of the disease [112]. All these properties make the polyphenols potential drugs for preventing and treating neurodegenerative diseases, in particular AD and PD. Actually, these compounds have shown to be effective in epidemiological, in vitro, and pre-clinical studies, but not in the early phase of the disease.

4.3 Polyphenols in Alzheimer’s and Parkinson’s disease

The effects of polyphenols on AD and PD can be divided into two main categories: the effects on nonamyloidogenic pathways (i.e., anti-oxidation pathway, interaction with cell signaling events, and interactions with enzymes) and the effects on amyloidogenic pathways. Below, the main beneficial effects shown by polyphenols on AD and PD are analyzed.

  1. Effects on memory. One of the hallmarks of AD is the memory impairment. This can be due to deficiency of factors, such as the brain-derived neurotrophic factor (BDNF) and the accumulation of formaldehyde. Polyphenols have been shown to improve the long-term memory by increasing BDNF concentration in vivo and decreasing the accumulation of formaldehyde [113, 114, 115].

  2. Effects on inflammation pathway. Inflammation plays an important role in the development of neurodegeneration. It is demonstrated that there is a correlation between the microglia activation and the neuroinflammatory response [116117]. Upon microglia activation, the transcription factor NF-kB (nuclear factor k-light-chain-enhancer of activated B cells) moves from cytoplasm to nucleus, inducing the expression of interleukins (i.e., IL-1β, IL-6, IL-12, and IL-23), other factors (i.e., TNF-α and iNOS), and cyclooxygenase 2 (COX-2). In this scenario, polyphenols can interact with certain types of kinases (including the mitogen-activated protein (MAP) kinase) preventing the activation of proinflammatory mediators [118, 119]. Polyphenol compounds are able to protect cells from inflammation by acting on reactive oxygen species (ROS), decreasing the secretion of prostaglandin E2 [120, 121, 122, 123] and increasing the amount of the regulatory enzyme sirtuin1 over sirtuin2, unbalanced after accumulation of Aβ [20]. Cell and PD-mouse model studies demonstrated that these compounds decrease the expression of NF-kB and other inflammatory factors [124, 125, 126].

  3. Effects on oxidative pathway, cell death and mitochondrial dysfunction. In neurodegeneration, there is an uncontrolled production of free radicals and ROS that are not detoxified by the dedicated systems [127]. This leads to macromolecule damage and progressively to cell death [128]. Polyphenols lower the amount of ROS, increase the expression of enzymes, like glutathione, dedicated to scavenger the free radicals and prevent the disruption of mitochondrial membranes [129]. In addition, these compounds seem to prevent the lipid peroxidation [130]. These effects indirectly influence the fibrillation process of Syn, affected by some byproducts of lipid oxidation and peroxidation [131], as demonstrated in PD-animal model studies [132]. Moreover, polyphenols inhibit the cell death by acting on proteins involved in the apoptosis mechanism like Bcl/Bax, caspase 3, and protein kinases and by decreasing the accumulation of Aβ fibrils that exert cytotoxic effects [133, 134]. Another important scenario affected by polyphenols is the mitochondrial dysfunction (MD) that becomes increasingly important in the onset of PD [135]. Different factors play a pivotal role in MD: the presence of neurotoxin, Complex 1 deficiency (involved in mitochondrial electron transport), and penetration of mitochondrial membrane by amyloid aggregates [136, 137]. Polyphenol compounds exert their activity restoring membrane potential, increasing the expression and activity of the Complex 1 and scavenging the ROS, free radicals, and metals [138, 139, 140, 141].

  4. Effects on acetylcholinesterase activity. Nearly 30 years ago, dysfunction in the cholinergic system was found correlated with AD and cognitive impairment [142]. This dysfunction can be originated by a reduction in acetylcholine synthesis, reduced levels of choline acetyltransferase, reduced choline uptake, or cholinergic neurons degeneration [143]. The use of acetylcholinesterase inhibitors to restore the cholinergic pathway has proved to alleviate the cognitive dysfunction in neurodegenerative diseases [144]. Polyphenol compounds have shown to inhibit acetylcholinesterase, improving memory, learning, and cognitive functions [145].

  5. Effects on Aβ formation. Polyphenol compounds act on the enzyme responsible for Aβ formation, decreasing the cleavage of APP into the peptide. They interact with and inhibit β-secretase [146]. In addition, they are able to restore the normal levels of γ-secretase, another enzyme involved in APP processing [147].

  6. Effects on the amyloidogenic pathways. Polyphenols can act on Aβ monomer preventing its fibrillation, through the stabilization of the monomer and/or to the formation of an off-pathway oligomer. This can be due to the interaction of polyphenols with metal ions that promote the Aβ aggregation or to the noncovalent interaction with the peptide [148]. They are also able to disaggregate oligomers and fibrils, interacting with the β-sheet structure. This has been confirmed by in vivo studies where polyphenol intake reduces the amyloid deposit in the mouse brain [149, 150]. Polyphenols exert their anti-amyloidogenic action by interfering also with the aggregation of Tau [151, 152, 153], inhibiting Tau phosphorylation in vitro [154] and in vivo [155]. Several polyphenols have been tested for their anti-fibrillogenic properties in vitro and in PD-animal models. Their main activity regards the interaction with Syn monomers leading to protein stabilization and fibrillation prevention [92]. Another factor concerns the formation of not toxic off-pathway oligomers that do not form fibrils nor interact with the membrane [156, 157]. Some polyphenols are also able to interact with oligomeric and fibrillar species, leading to their destabilization [65, 92]. The major effect of polyphenols is due to the noncovalent interaction with the Syn C-terminal domain. In addition, these compounds can chemically modify the lysine residues, present mainly in the N-terminal region, through Michael addition and Schiff-base formation [158]. This reduces the conformational plasticity of Syn and its tendency to be converted into fibrils. Moreover, structure-activity relationship studies indicate that the differences in polyphenols activities reside in the number and position of OH groups in the phenyl ring [159].

Advertisement

5. Polyphenols as a drug in the brain delivery system

5.1 Blood-brain barrier and neurodegeneration

The human brain comprises more than 600 km of blood vessels that guarantee oxygen, energy metabolites, and nutrients to brain cells and remove carbon dioxide and toxic metabolic products from the brain to the systemic circulation. A highly selective semipermeable border, called blood-brain barrier (BBB), separates the circulating blood from the central nervous system (CNS), regulating CNS homeostasis. Brain microvascular endothelia cells, neurons, astrocyte, pericytes, tight junctions, and basal membrane constitute tight brain capillaries in the BBB [160, 161]. It follows that BBB does not have fenestrations or other physical fissures for diffusion of small molecules. In fact, ions, solutes, and hormones can pass the BBB by passive diffusion through the paracellular pathway between adjacent cells. Hydrophilic biomolecules (i.e., proteins and peptides) can cross the BBB within specific and saturable receptor-mediated transport mechanisms [162]. The components of BBB constantly adapt in response to various physiological and pathological modifications into the brain [163, 164]. Loss of BBB integrity is correlated with vascular permeability increase, cerebral blood flow impairs, and hemodynamic response alteration [165]. In neurodegenerative disorders, endothelia degeneration leads to loss of tight junctions [166, 167], brain capillary leakages [168, 169], pericyte degeneration [170], endothelial cell remodeling [164], cellular infiltration [171, 172], and aberrant angiogenesis [173, 174]. All these BBB disruptions let different blood proteins (i.e., fibrinogen, plasminogen, and thrombin), water, and electrolytes to accumulate in different zones of CNS, enhancing the on progress of PD and AD [165]. Consequently, to project effective drugs for neurodegeneration, it is necessary to understand in detail BBB pathological aberrations.

Due to their safeness and tolerance [175, 176, 177], polyphenols are currently studied as neuroprotectors. It is important to point out that for exerting their action, polyphenols must accumulate in the brain in an active form and in sufficient concentration. The limiting step is choosing the right administration route. In most of the clinical studies, the oral administration is the preferred way, but recently the nasal delivery is taken into consideration for the easiness to bypass the BBB [178], the increased bioavailability, the decreased metabolism, and peripheral side effects [179, 180]. The major problem of oral administration relies on poor absorbance of the modified form of polyphenols (i.e., glycosides and ester polymers) in the upper portion of the gut leading to the passage in the colon in which polyphenols are converted by gut-microbiota in the aglycone form or other substances able to be better absorbed [181, 182]. Once absorbed, they can be further modified by enzymes and eliminated [183, 184] or adsorbed to plasmatic proteins (i.e., albumin) and then accumulated in different districts [185].

5.2 Nanotechnology-based delivery system: An innovative strategy

Nanotechnology is a new branch of science involving the formulation, synthesis, and characterization of small particles, with diameters ranging from 1 to 1000 nm [186], which become key players in innovative drug delivery and cell targeting. Recent studies suggest that nanoparticle-based delivery systems represent innovative and promising approaches to improve drug solubility, prevent acid-degradation, minimize toxic side effects, and increase blood availability [187188]. Considering the low bioavailability of polyphenols, different strategies have been developed in order to enhance their chemical stability, solubility, and cell-membrane permeability. These goals have been achieved by adding chemical agents to preserve the structure [189], enzyme inhibitors to contrast biotransformation [190], and lipids or proteins to increase the solubility [191]. Recently, nanoparticle-mediated delivery system is emerged as the most promising approach. Using biodegradable and biocompatible polymers, polyphenols can be encapsulated in different nanostructures and then possibly administrated via intravenous, transdermal, nasal, and oral route. As describe above, this aspect is fundamental in neurological diseases, in which polyphenols must cross the BBB, with the opportune grade of lipophilicity [162, 192, 193] and reach the brain tissue in sufficient quantities for therapeutic use. These new delivery systems are represented by nanospheres, nanocapsules, nanoemulsions, solid lipid nanoparticles, cyclodextrins, liposomes, and micelles (Figure 7).

Figure 7.

Schematic representation of nanosized delivery systems for polyphenols. Nanoparticles can enhance polyphenol bioavailability, enhancing their adsorption across intestinal epithelium, increasing their concentration in the bloodstream, and improving their ability to cross the blood-brain barrier.

Nanospheres (10–200 nm) [194] are homogeneous solid matrix particles characterized by a hydrophobic portion in the inner part and hydrophilic chains anchored on the surface. In nanospheres, the drug is dissolved, entrapped, encapsulated, or attached to the matrix of the polymer, so protected from chemical and enzymatic degradation. Various kinds of polymers are used to prepare nanospheres: polylactic acid (PLA), poly-glycolic acid (PGA), poly-lactic-co-glycolic acid (PLGA), polyethylene glycol (PEG), poly ε-caprolactone (PCL), and chitosan (CS) [195, 196].

Nanocapsules (10–1000 nm) have a similar chemical composition but comprise an oily or aqueous core, which is surrounded by a thin polymer membrane [197198]. The cavity can contain the drug in liquid or solid form. Furthermore, the medication can be carried on nanovector surface or absorbed in the polymeric membrane [198, 199, 200].

Nanoemulsions are oil-in-water or water-in-oil emulsions stabilized by one or more surfactants (i.e., phosphatidylcholine, sodium deoxycholate, sorbitan monolaurate, poloxamers, sodium dodecyl sulfate, and poly(ethylene glycol)) delivered in droplets of small dimensions (100–300 nm) [191]. The strategy allows having a higher surface area and a long-term chemical and physical stability [201, 202]. Nanoemulsions represent an innovative formulation to deliver polyphenols directly into the brain through the intranasal route. In fact, mucoadesive polymers, such as CS, can be added to slow down nasal clearance [191].

Solid lipid nanoparticles (50–1000 nm) [194] are composed of high melting point lipid, organized in a solid core, coated by aqueous surfactants (i.e., sphingomyelins, bile salts, and sterols) [198]. Even though these nanoparticles present high biocompatibility, bioavailability and physical stability, the common undesirable disadvantages are particle growth, arbitrary gelation tendency, and unpredicted dynamic of polymorphic transitions [198].

Cyclodextrins (1–2 nm) [194] are a group of structurally related natural products formed from the bacterial digestion of cellulose. Cyclodextrins are cyclic oligosaccharides consisting of (α-1,4)-linked α-D-glucopyranose units with a lipophilic central cavity and a hydrophilic outer surface [203]. The hydroxyl functions are orientated to the exterior, while the central cavity is wrinkled by the skeletal carbons and ethereal oxygens of the glucose residues. Natural cyclodextrins are classified by the number of glucopyranose units in α-(six units), β-(seven units), and γ-(eight units) [204]. Recently, cyclodextrins containing from 9 to 13 glucopyranose units have been reported. These carriers are useful for increasing the solubility and the stability of poorly water-soluble drugs. Moreover, cyclodextrins can be derivatized with hydroxypropyl, methyl, and sulfobutyl-ether additives [203]. So, drugs can be allocated into the cavity via van der Waals forces, hydrophobic interactions, or hydrogen bonds [205].

Liposomes (30–2000 nm) [194] are phospholipid vesicles containing one or more concentric lipid bilayers enclosing an aqueous space. Liposomes can assemble spontaneously by hydration of lipid-derivate powder (i.e., cholesterol, glycolipids, sphingolipids, long chain fatty acids, and membrane proteins) in aqueous buffer [195]. Due to their ability to capture hydrophilic and lipophilic substances, in the aqueous space or into the lipid bilayer membrane, respectively, they can protect drugs from early inactivation, degradation, and loss [206].

Micelles (5–100 nm) are colloidal dispersions, consisting of amphiphilic copolymers (i.e., PEG, PLGA, and PCL) that assemble naturally in water at a specific concentration and temperature [207]. When polymer concentration is greater than the critical micelle concentration, micelles start to be assembled: hydrophobic fragments of amphiphilic reagents form the core, whereas hydrophilic portion form the shells [208]. Micelles are characterized by high stability, biocompatibility, and ability to keep in solution poorly soluble drugs.

5.3 Nanotechnology as an innovative delivery system of polyphenols

The use of biodegradable and biocompatible polymers allows rationalizing the design of innovative nanostructures able to encapsulate polyphenols that can cross the BBB, improving the limitations associated with conventional administrations. In this scenario, curcumin is the most studied drug candidate, due to the prominent results obtained in the animal model of neurodegenerative diseases [209, 210, 211]. In fact, the efficacy of curcumin is so far limited by the poor aqueous solubility, low adsorption in the gastrointestinal tract, and rapid metabolism. Nanosphere of PGLA containing curcumin can be the right strategy for crossing BBB. Recent studies indicated how curcumin-PGLA nanoparticles can interfere with Aβ aggregation and improve the brain self-repair mechanism, increasing the neural stem cell proliferation and neuronal differentiation [212]. In the same way, liposomes loaded with curcumin can efficiently inhibit the in vitro formation of Aβ fibrils and deposition in the brain [213]. Curcumin-solid lipid nanoparticles seem to be effective for MD and central oxidative stress [214]. In addition, curcumin and piperine co-loaded glycerol mono-oleate nanoparticles can interfere with Syn aggregation, reducing oxidative damage and apoptosis [215]. Curcumin was also taken in consideration for intranasal delivery to the central nervous system by nanoemulsions. In the presence of CS, nanoemulsions of curcumin (added in the oil phase) can effectively cross the mucosa without showing cytotoxicity [209].

Another good candidate is resveratrol. It is known for its ability to induce the degradation of APP and to remove Aβ [216]. But, due to its rapid and extensive metabolism, resveratrol is subjected to a person-to-person bioavailability. PEG-PCL and PGLA nanoparticles loaded with resveratrol let a controlled release profile of the drug, essential for prolonging its plasmatic level and the antioxidant activity [217, 218]. A promising approach is the oil-in-water nanoemulsion [219]. Adding Vitamin E and other surfactants, this formulation can reach the brain via the nasal route, with encouraging efficacy [220]. Furthermore, the co-encapsulation of curcumin and resveratrol (1:1 weight ratio) in mucoadhesive nanoemulsions protects the active substances from degradation and preserves their antioxidant properties. Notably, in vivo quantification in animal brain indicated an increase of the amount of the two polyphenols after 6 hours [221]. Unfortunately, these systems have not yet reached clinical trials, but the results accumulated so far encourage new original therapeutic approaches.

Advertisement

Acknowledgments

This project was supported by Progetti di Ateneo-University of Padova 2017-N. C93C1800002600 and by MIUR-PNRA (Programma Nazionale Ricerche in Antartide) (PNRA16_00068). We thank Samuele Cesaro and Ferdinando Polverino de Laureto for the elaboration of the images.

References

  1. 1. Katzman R. Alzheimer’s disease. The New England Journal of Medicine. 1986;314(15):964-973
  2. 2. Selkoe DJ. Alzheimer’s disease: Genes, proteins, and therapy. Physiological Reviews. 2001;81(2):741-766
  3. 3. Glenner GG, Wong CW. Alzheimer’s disease: Initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochemical and Biophysical Research Communications. 1984;120(3):885-890
  4. 4. Lee VM, Goedert M, Trojanowski JQ. Neurodegenerative tauopathies. Annual Review of Neuroscience. 2001;24:1121-1159
  5. 5. Serrano-Pozo A, Frosch MP, Masliah E, Hyman BT. Neuropathological alterations in Alzheimer disease. Cold Spring Harbor Perspectives in Medicine. 2011;1(1):a006189
  6. 6. Trojanowski JQ , Lee VM. “Fatal attractions” of proteins. A comprehensive hypothetical mechanism underlying Alzheimer’s disease and other neurodegenerative disorders. Annals of the New York Academy of Sciences. 2000;924:62-67
  7. 7. Masliah E, Miller A, Terry RD. The synaptic organization of the neocortex in Alzheimer’s disease. Medical Hypotheses. 1993;41(4):334-340
  8. 8. Jankovic J. Parkinson’s disease: Clinical features and diagnosis. Journal of Neurology, Neurosurgery, and Psychiatry. 2008;79(4):368-376
  9. 9. Spillantini MG, Schmidt ML, Lee VM, Trojanowski JQ , Jakes R, Goedert M. Alpha-synuclein in Lewy bodies. Nature. 1997;388(6645):839-840
  10. 10. Goedert M. Alpha-synuclein and neurodegenerative diseases. Nature Reviews. Neuroscience. 2001;2(7):492-501
  11. 11. Shahmoradian SH, Lewis AJ, Genoud C, Hench J, Moors TE, Navarro PP, et al. Lewy pathology in Parkinson’s disease consists of crowded organelles and lipid membranes. Nature Neuroscience. 2019;22(7):1099-1109
  12. 12. Ruiperez V, Darios F, Davletov B. Alpha-synuclein, lipids and Parkinson’s disease. Progress in Lipid Research. 2010;49(4):420-428
  13. 13. Fecchio C, Palazzi L, Polverino de Laureto P. Alpha-synuclein and polyunsaturated fatty acids: Molecular basis of the interaction and implication in neurodegeneration. Molecules. 2018;23(7):1531-1551
  14. 14. Bekris LM, Yu CE, Bird TD, Tsuang DW. Genetics of Alzheimer disease. Journal of Geriatric Psychiatry and Neurology. 2010;23(4):213-227
  15. 15. Goate A, Chartier-Harlin MC, Mullan M, Brown J, Crawford F, Fidani L, et al. Segregation of a missense mutation in the amyloid precursor protein gene with familial Alzheimer’s disease. Nature. 1991;349(6311):704-706
  16. 16. Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, et al. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science. 1993;261(5123):921-923
  17. 17. Citron M, Westaway D, Xia W, Carlson G, Diehl T, Levesque G, et al. Mutant presenilins of Alzheimer’s disease increase production of 42-residue amyloid beta-protein in both transfected cells and transgenic mice. Nature Medicine. 1997;3(1):67-72
  18. 18. Kelleher RJ 3rd, Shen J. Presenilin-1 mutations and Alzheimer’s disease. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(4):629-631
  19. 19. Scheuner D, Eckman C, Jensen M, Song X, Citron M, Suzuki N, et al. Secreted amyloid beta-protein similar to that in the senile plaques of Alzheimer’s disease is increased in vivo by the presenilin 1 and 2 and APP mutations linked to familial Alzheimer’s disease. Nature Medicine. 1996;2(8):864-870
  20. 20. Porquet D, Casadesus G, Bayod S, Vicente A, Canudas AM, Vilaplana J, et al. Dietary resveratrol prevents Alzheimer’s markers and increases life span in SAMP8. Age (Dordrecht, Netherlands). 2013;35(5):1851-1865
  21. 21. Kruger R, Kuhn W, Muller T, Woitalla D, Graeber M, Kosel S, et al. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease. Nature Genetics. 1998;18(2):106-108
  22. 22. Lesage S, Anheim M, Letournel F, Bousset L, Honore A, Rozas N, et al. G51D alpha-synuclein mutation causes a novel parkinsonian-pyramidal syndrome. Annals of Neurology. 2013;73(4):459-471
  23. 23. Polymeropoulos MH, Higgins JJ, Golbe LI, Johnson WG, Ide SE, Di Iorio G, et al. Mapping of a gene for Parkinson’s disease to chromosome 4q21-q23. Science. 1996;274(5290):1197-1199
  24. 24. Singleton AB, Farrer M, Johnson J, Singleton A, Hague S, Kachergus J, et al. Alpha-Synuclein locus triplication causes Parkinson’s disease. Science. 2003;302(5646):841
  25. 25. Zarranz JJ, Alegre J, Gomez-Esteban JC, Lezcano E, Ros R, Ampuero I, et al. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Annals of Neurology. 2004;55(2):164-173
  26. 26. Nuytemans K, Theuns J, Cruts M, Van Broeckhoven C. Genetic etiology of Parkinson disease associated with mutations in the SNCA, PARK2, PINK1, PARK7, and LRRK2 genes: A mutation update. Human Mutation. 2010;31(7):763-780
  27. 27. Kumari U, Tan EK. LRRK2 in Parkinson’s disease: Genetic and clinical studies from patients. The FEBS Journal. 2009;276(22):6455-6463
  28. 28. Dawson TM, Dawson VL. The role of parkin in familial and sporadic Parkinson’s disease. Movement Disorders. 2010;25(Suppl 1):S32-S39
  29. 29. Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey K, Gispert S, et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1. Science. 2004;304(5674):1158-1160
  30. 30. Cookson MR. Parkinsonism due to mutations in PINK1, parkin, and DJ-1 and oxidative stress and mitochondrial pathways. Cold Spring Harbor Perspectives in Medicine. 2012;2(9):a009415
  31. 31. Cline EN, Bicca MA, Viola KL, Klein WL. The amyloid-beta oligomer hypothesis: Beginning of the third decade. Journal of Alzheimer’s Disease. 2018;64(s1):S567-s610
  32. 32. Serpell LC. Alzheimer’s amyloid fibrils: Structure and assembly. Biochimica et Biophysica Acta. 2000;1502(1):16-30
  33. 33. Petkova AT, Ishii Y, Balbach JJ, Antzutkin ON, Leapman RD, Delaglio F, et al. A structural model for Alzheimer’s beta -amyloid fibrils based on experimental constraints from solid state NMR. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(26):16742-16747
  34. 34. Nelson R, Sawaya MR, Balbirnie M, Madsen AO, Riekel C, Grothe R, et al. Structure of the cross-beta spine of amyloid-like fibrils. Nature. 2005;435(7043):773-778
  35. 35. Makin OS, Atkins E, Sikorski P, Johansson J, Serpell LC. Molecular basis for amyloid fibril formation and stability. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(2):315-320
  36. 36. Fitzpatrick AW, Debelouchina GT, Bayro MJ, Clare DK, Caporini MA, Bajaj VS, et al. Atomic structure and hierarchical assembly of a cross-beta amyloid fibril. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(14):5468-5473
  37. 37. Dobson CM. The structural basis of protein folding and its links with human disease. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 2001;356(1406):133-145
  38. 38. Clinton LK, Blurton-Jones M, Myczek K, Trojanowski JQ , LaFerla FM. Synergistic interactions between Abeta, tau, and alpha-synuclein: Acceleration of neuropathology and cognitive decline. The Journal of Neuroscience. 2010;30(21):7281-7289
  39. 39. Larson ME, Sherman MA, Greimel S, Kuskowski M, Schneider JA, Bennett DA, et al. Soluble alpha-synuclein is a novel modulator of Alzheimer’s disease pathophysiology. The Journal of Neuroscience. 2012;32(30):10253-10266
  40. 40. Masliah E, Rockenstein E, Veinbergs I, Sagara Y, Mallory M, Hashimoto M, et al. Beta-amyloid peptides enhance alpha-synuclein accumulation and neuronal deficits in a transgenic mouse model linking Alzheimer’s disease and Parkinson’s disease. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(21):12245-12250
  41. 41. Bachhuber T, Katzmarski N, McCarter JF, Loreth D, Tahirovic S, Kamp F, et al. Inhibition of amyloid-beta plaque formation by alpha-synuclein. Nature Medicine. 2015;21(7):802-807
  42. 42. Haass C, Selkoe DJ. Cellular processing of beta-amyloid precursor protein and the genesis of amyloid beta-peptide. Cell. 1993;75(6):1039-1042
  43. 43. Nunan J, Small DH. Regulation of APP cleavage by alpha-, beta- and gamma-secretases. FEBS Letters. 2000;483(1):6-10
  44. 44. Kim W, Hecht MH. Sequence determinants of enhanced amyloidogenicity of Alzheimer a{beta}42 peptide relative to a{beta}40. The Journal of Biological Chemistry. 2005;280(41):35069-35076
  45. 45. Shoji M, Golde TE, Ghiso J, Cheung TT, Estus S, Shaffer LM, et al. Production of the Alzheimer amyloid beta protein by normal proteolytic processing. Science. 1992;258(5079):126-129
  46. 46. Selkoe DJ. Amyloid beta protein precursor and the pathogenesis of Alzheimer’s disease. Cell. 1989;58(4):611-612
  47. 47. Iwata N, Tsubuki S, Takaki Y, Shirotani K, Lu B, Gerard NP, et al. Metabolic regulation of brain Abeta by neprilysin. Science. 2001;292(5521):1550-1552
  48. 48. Turner RT 3rd, Koelsch G, Hong L, Castanheira P, Ermolieff J, Ghosh AK, et al. Subsite specificity of memapsin 2 (beta-secretase): Implications for inhibitor design. Biochemistry. 2001;40(34):10001-10006
  49. 49. Kim J, Basak JM, Holtzman DM. The role of apolipoprotein E in Alzheimer’s disease. Neuron. 2009;63(3):287-303
  50. 50. Bendiske J, Bahr BA. Lysosomal activation is a compensatory response against protein accumulation and associated synaptopathogenesis--an approach for slowing Alzheimer disease? Journal of Neuropathology and Experimental Neurology. 2003;62(5):451-463
  51. 51. Zhang S, Iwata K, Lachenmann MJ, Peng JW, Li S, Stimson ER, et al. The Alzheimer’s peptide a beta adopts a collapsed coil structure in water. Journal of Structural Biology. 2000;130(2-3):130-141
  52. 52. Sgourakis NG, Yan Y, McCallum SA, Wang C, Garcia AE. The Alzheimer’s peptides Abeta40 and 42 adopt distinct conformations in water: A combined MD/NMR study. Journal of Molecular Biology. 2007;368(5):1448-1457
  53. 53. D’Ursi AM, Armenante MR, Guerrini R, Salvadori S, Sorrentino G, Picone D. Solution structure of amyloid beta-peptide (25-35) in different media. Journal of Medicinal Chemistry. 2004;47(17):4231-4238
  54. 54. Crescenzi O, Tomaselli S, Guerrini R, Salvadori S, D’Ursi AM, Temussi PA, et al. Solution structure of the Alzheimer amyloid beta-peptide (1-42) in an apolar microenvironment. Similarity with a virus fusion domain. European Journal of Biochemistry. 2002;269(22):5642-5648
  55. 55. McLaurin J, Chakrabartty A. Characterization of the interactions of Alzheimer beta-amyloid peptides with phospholipid membranes. European Journal of Biochemistry. 1997;245(2):355-363
  56. 56. Williams TL, Serpell LC. Membrane and surface interactions of Alzheimer’s Abeta peptide–insights into the mechanism of cytotoxicity. The FEBS Journal. 2011;278(20):3905-3917
  57. 57. Butterfield SM, Lashuel HA. Amyloidogenic protein-membrane interactions: Mechanistic insight from model systems. Angewandte Chemie (International Ed. in English). 2010;49(33):5628-5654
  58. 58. Bode DC, Freeley M, Nield J, Palma M, Viles JH. Amyloid-beta oligomers have a profound detergent-like effect on lipid membrane bilayers, imaged by atomic force and electron microscopy. The Journal of Biological Chemistry. 2019;294(19):7566-7572
  59. 59. Avila J, Jimenez JS, Sayas CL, Bolos M, Zabala JC, Rivas G, et al. Tau Structures. Frontiers in Aging Neuroscience. 2016;8:262
  60. 60. Goedert M, Spillantini MG, Cairns NJ, Crowther RA. Tau proteins of Alzheimer paired helical filaments: Abnormal phosphorylation of all six brain isoforms. Neuron. 1992;8(1):159-168
  61. 61. Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA. Multiple isoforms of human microtubule-associated protein tau: Sequences and localization in neurofibrillary tangles of Alzheimer’s disease. Neuron. 1989;3(4):519-526
  62. 62. Zhang Z, Song M, Liu X, Kang SS, Kwon IS, Duong DM, et al. Cleavage of tau by asparagine endopeptidase mediates the neurofibrillary pathology in Alzheimer’s disease. Nature Medicine. 2014;20(11):1254-1262
  63. 63. Soeda Y, Yoshikawa M, Almeida OF, Sumioka A, Maeda S, Osada H, et al. Toxic tau oligomer formation blocked by capping of cysteine residues with 1,2-dihydroxybenzene groups. Nature Communications. 2015;6:10216
  64. 64. Gerson JE, Sengupta U, Lasagna-Reeves CA, Guerrero-Munoz MJ, Troncoso J, Kayed R. Characterization of tau oligomeric seeds in progressive supranuclear palsy. Acta Neuropathologica Communications. 2014;2:73
  65. 65. Ahsan N, Mishra S, Jain MK, Surolia A, Gupta S. Curcumin Pyrazole and its derivative N-(3-Nitrophenylpyrazole) Curcumin inhibit aggregation, disrupt fibrils and modulate toxicity of wild type and mutant alpha-Synuclein. Scientific Reports. 2015;5:9862
  66. 66. Grundke-Iqbal I, Iqbal K, Quinlan M, Tung YC, Zaidi MS, Wisniewski HM. Microtubule-associated protein tau. A component of Alzheimer paired helical filaments. The Journal of Biological Chemistry. 1986;261(13):6084-6089
  67. 67. Fitzpatrick AWP, Falcon B, He S, Murzin AG, Murshudov G, Garringer HJ, et al. Cryo-EM structures of tau filaments from Alzheimer’s disease. Nature. 2017;547(7662):185-190
  68. 68. Barbour R, Kling K, Anderson JP, Banducci K, Cole T, Diep L, et al. Red blood cells are the major source of alpha-synuclein in blood. Neurodegenerative Diseases. 2008;5(2):55-59
  69. 69. Pineda A, Burre J. Modulating membrane binding of alpha-synuclein as a therapeutic strategy. Proceedings of the National Academy of Sciences of the United States of America. 2017;114(6):1223-1225
  70. 70. Chandra S, Gallardo G, Fernandez-Chacon R, Schluter OM, Sudhof TC. Alpha-synuclein cooperates with CSPalpha in preventing neurodegeneration. Cell. 2005;123(3):383-396
  71. 71. Burre J, Sharma M, Tsetsenis T, Buchman V, Etherton MR, Sudhof TC. Alpha-synuclein promotes SNARE-complex assembly in vivo and in vitro. Science. 2010;329(5999):1663-1667
  72. 72. Ueda K, Fukushima H, Masliah E, Xia Y, Iwai A, Yoshimoto M, et al. Molecular cloning of cDNA encoding an unrecognized component of amyloid in Alzheimer disease. Proceedings of the National Academy of Sciences of the United States of America. 1993;90(23):11282-11286
  73. 73. Eliezer D, Kutluay E, Bussell R Jr, Browne G. Conformational properties of alpha-synuclein in its free and lipid-associated states. Journal of Molecular Biology. 2001;307(4):1061-1073
  74. 74. Hoyer W, Cherny D, Subramaniam V, Jovin TM. Impact of the acidic C-terminal region comprising amino acids 109-140 on alpha-synuclein aggregation in vitro. Biochemistry. 2004;43(51):16233-16242
  75. 75. Uversky VN, Eliezer D. Biophysics of Parkinson’s disease: Structure and aggregation of alpha-synuclein. Current Protein & Peptide Science. 2009;10(5):483-499
  76. 76. Dedmon MM, Lindorff-Larsen K, Christodoulou J, Vendruscolo M, Dobson CM. Mapping long-range interactions in alpha-synuclein using spin-label NMR and ensemble molecular dynamics simulations. Journal of the American Chemical Society. 2005;127(2):476-477
  77. 77. Davidson WS, Jonas A, Clayton DF, George JM. Stabilization of alpha-synuclein secondary structure upon binding to synthetic membranes. The Journal of Biological Chemistry. 1998;273(16):9443-9449
  78. 78. Madine J, Doig AJ, Middleton DA. A study of the regional effects of alpha-synuclein on the organization and stability of phospholipid bilayers. Biochemistry. 2006;45(18):5783-5792
  79. 79. Kamp F, Beyer K. Binding of alpha-synuclein affects the lipid packing in bilayers of small vesicles. The Journal of Biological Chemistry. 2006;281(14):9251-9259
  80. 80. Galvagnion C. The role of lipids interacting with alpha-synuclein in the pathogenesis of Parkinson’s disease. Journal of Parkinson’s Disease. 2017;7(3):433-450
  81. 81. Hansen RA, Gartlehner G, Webb AP, Morgan LC, Moore CG, Jonas DE. Efficacy and safety of donepezil, galantamine, and rivastigmine for the treatment of Alzheimer’s disease: A systematic review and meta-analysis. Clinical Interventions in Aging. 2008;3(2):211-225
  82. 82. van Marum RJ. Update on the use of memantine in Alzheimer’s disease. Neuropsychiatric Disease and Treatment. 2009;5:237-247
  83. 83. Baazaoui N, Iqbal K. A novel therapeutic approach to treat Alzheimer’s disease by neurotrophic support during the period of synaptic compensation. Journal of Alzheimer’s Disease. 2018;62(3):1211-1218
  84. 84. Jankovic J, Aguilar LG. Current approaches to the treatment of Parkinson’s disease. Neuropsychiatric Disease and Treatment. 2008;4(4):743-757
  85. 85. Chen S, Ge X, Chen Y, Lv N, Liu Z, Yuan W. Advances with RNA interference in Alzheimer’s disease research. Drug Design, Development and Therapy. 2013;7:117-125
  86. 86. Sapru MK, Yates JW, Hogan S, Jiang L, Halter J, Bohn MC. Silencing of human alpha-synuclein in vitro and in rat brain using lentiviral-mediated RNAi. Experimental Neurology. 2006;198(2):382-390
  87. 87. Lewis PA. Emerging pathways in genetic Parkinson’s disease. The FEBS Journal. 2008;275(23):5747
  88. 88. Mittal S, Bjornevik K, Im DS, Flierl A, Dong X, Locascio JJ, et al. beta2-Adrenoreceptor is a regulator of the alpha-synuclein gene driving risk of Parkinson’s disease. Science. 2017;357(6354):891-898
  89. 89. Chen XF, Zhang YW, Xu H, Bu G. Transcriptional regulation and its misregulation in Alzheimer’s disease. Molecular Brain. 2013;6:44
  90. 90. Boland B, Yu WH, Corti O, Mollereau B, Henriques A, Bezard E, et al. Promoting the clearance of neurotoxic proteins in neurodegenerative disorders of ageing. Nature Reviews. Drug Discovery. 2018;17(9):660-688
  91. 91. Brundin P, Dave KD, Kordower JH. Therapeutic approaches to target alpha-synuclein pathology. Experimental Neurology. 2017;298(Pt B):225-235
  92. 92. Palazzi L, Bruzzone E, Bisello G, Leri M, Stefani M, Bucciantini M, et al. Oleuropein aglycone stabilizes the monomeric alpha-synuclein and favours the growth of non-toxic aggregates. Scientific Reports. 2018;8(1):8337-8354
  93. 93. Harborne JB. Introduction to Ecological Biochemistry. 2nd ed. New York: Academic Press; 1982
  94. 94. Nicholson R, Hammerschmidt R. Phenolic compounds and their role in disease resistance. Annual Review of Phytopathology. 1992;30:369-389
  95. 95. Lattanzio V, Cardinali A, Ruta C, Morone Fortunato I, Lattanzio VM, Linsalata V, et al. Relationship of secondary metabolism to growth in oregano (Origanum vulgare L.) shoot cultures under nutritional stress. Environmental and Experimental Botany. 2009;65(1):54-62
  96. 96. Andersen ØM, Markham KR. Flavonoids, chemistry, biochemistry and applications. Journal of Natural Products. 2007;70:1-1256
  97. 97. Kondratyuk TP, Pezzuto JM. Natural product polyphenols of relevance to human health. Pharmaceutical Biology. 2004;42:46-63
  98. 98. Spencer JP, Abd El Mohsen MM, Minihane AM, Mathers JC. Biomarkers of the intake of dietary polyphenols: Strengths, limitations and application in nutrition research. The British Journal of Nutrition. 2008;99(1):12-22
  99. 99. Arts IC, Hollman PC. Polyphenols and disease risk in epidemiologic studies. The American Journal of Clinical Nutrition. 2005;81(1 Suppl):317s-325s
  100. 100. Scalbert A, Manach C, Morand C, Remesy C, Jimenez L. Dietary polyphenols and the prevention of diseases. Critical Reviews in Food Science and Nutrition. 2005;45(4):287-306
  101. 101. Nardini M, Natella F, Scaccini C. Role of dietary polyphenols in platelet aggregation. A review of the supplementation studies. Platelets. 2007;18(3):224-243
  102. 102. Renaud S, de Lorgeril M. Wine, alcohol, platelets, and the French paradox for coronary heart disease. Lancet. 1992;339(8808):1523-1526
  103. 103. Dubick MA, Omaye ST. Evidence for grape, wine and tea polyphenols as modulators of atherosclerosis and ischemic heart disease in humans. Journal of Nutraceuticals, Functional & Medical Foods. 2001;3:67-93
  104. 104. Rizvi SI, Zaid MA. Insulin-like effect of (−)epicatechin on erythrocyte membrane acetylcholinesterase activity in type 2 diabetes mellitus. Clinical and Experimental Pharmacology & Physiology. 2001;28(9):776-778
  105. 105. Rizvi SI, Zaid MA, Anis R, Mishra N. Protective role of tea catechins against oxidation-induced damage of type 2 diabetic erythrocytes. Clinical and Experimental Pharmacology & Physiology. 2005;32(1-2):70-75
  106. 106. Yang CS, Landau JM, Huang MT, Newmark HL. Inhibition of carcinogenesis by dietary polyphenolic compounds. Annual Review of Nutrition. 2001;21:381-406
  107. 107. Johnson IT, Williamson G, Musk SR. Anticarcinogenic factors in plant foods: A new class of nutrients? Nutrition Research Reviews. 1994;7(1):175-204
  108. 108. Cao G, Booth SL, Sadowski JA, Prior RL. Increases in human plasma antioxidant capacity after consumption of controlled diets high in fruit and vegetables. The American Journal of Clinical Nutrition. 1998;68(5):1081-1087
  109. 109. Letenneur L, Proust-Lima C, Le Gouge A, Dartigues JF, Barberger-Gateau P. Flavonoid intake and cognitive decline over a 10-year period. American Journal of Epidemiology. 2007;165(12):1364-1371
  110. 110. Luqman S, Rizvi SI. Protection of lipid peroxidation and carbonyl formation in proteins by capsaicin in human erythrocytes subjected to oxidative stress. Phytotherapy Research. 2006;20(4):303-306
  111. 111. Pandey KB, Rizvi SI. Plant polyphenols as dietary antioxidants in human health and disease. Oxidative Medicine and Cellular Longevity. 2009;2(5):270-278
  112. 112. Garcia-Lafuente A, Guillamon E, Villares A, Rostagno MA, Martinez JA. Flavonoids as anti-inflammatory agents: Implications in cancer and cardiovascular disease. Inflammation Research. 2009;58(9):537-552
  113. 113. Zhao YN, Li WF, Li F, Zhang Z, Dai YD, Xu AL, et al. Resveratrol improves learning and memory in normally aged mice through microRNA-CREB pathway. Biochemical and Biophysical Research Communications. 2013;435(4):597-602
  114. 114. Mei Y, Jiang C, Wan Y, Lv J, Jia J, Wang X, et al. Aging-associated formaldehyde-induced norepinephrine deficiency contributes to age-related memory decline. Aging Cell. 2015;14(4):659-668
  115. 115. Moorthi P, Premkumar P, Priyanka R, Jayachandran KS, Anusuyadevi M. Pathological changes in hippocampal neuronal circuits underlie age-associated neurodegeneration and memory loss: Positive clue toward SAD. Neuroscience. 2015;301:90-105
  116. 116. Imamura K, Hishikawa N, Sawada M, Nagatsu T, Yoshida M, Hashizume Y. Distribution of major histocompatibility complex class II-positive microglia and cytokine profile of Parkinson’s disease brains. Acta Neuropathologica. 2003;106(6):518-526
  117. 117. Beach TG, Sue LI, Walker DG, Lue LF, Connor DJ, Caviness JN, et al. Marked microglial reaction in normal aging human substantia nigra: Correlation with extraneuronal neuromelanin pigment deposits. Acta Neuropathologica. 2007;114(4):419-424
  118. 118. Chen X, Yang X, Liu T, Guan M, Feng X, Dong W, et al. Kaempferol regulates MAPKs and NF-kappaB signaling pathways to attenuate LPS-induced acute lung injury in mice. International Immunopharmacology. 2012;14(2):209-216
  119. 119. Choi JS, Islam MN, Ali MY, Kim EJ, Kim YM, Jung HA. Effects of C-glycosylation on anti-diabetic, anti-Alzheimer’s disease and anti-inflammatory potential of apigenin. Food and Chemical Toxicology. 2014;64:27-33
  120. 120. Yao Y, Li J, Niu Y, Yu JQ , Yan L, Miao ZH, et al. Resveratrol inhibits oligomeric Abeta-induced microglial activation via NADPH oxidase. Molecular Medicine Reports. 2015;12(4):6133-6139
  121. 121. Wight RD, Tull CA, Deel MW, Stroope BL, Eubanks AG, Chavis JA, et al. Resveratrol effects on astrocyte function: Relevance to neurodegenerative diseases. Biochemical and Biophysical Research Communications. 2012;426(1):112-115
  122. 122. Kim YA, Lim SY, Rhee SH, Park KY, Kim CH, Choi BT, et al. Resveratrol inhibits inducible nitric oxide synthase and cyclooxygenase-2 expression in beta-amyloid-treated C6 glioma cells. International Journal of Molecular Medicine. 2006;17(6):1069-1075
  123. 123. Cheng X, Wang Q , Li N, Zhao H. Effects of resveratrol on hippocampal astrocytes and expression of TNF-alpha in Alzheimer’s disease model rate. Wei Sheng Yan Jiu. 2015;44(4):610-614
  124. 124. Raza SS, Khan MM, Ahmad A, Ashafaq M, Islam F, Wagner AP, et al. Neuroprotective effect of naringenin is mediated through suppression of NF-kappaB signaling pathway in experimental stroke. Neuroscience. 2013;230:157-171
  125. 125. Jia Z, Babu PV, Si H, Nallasamy P, Zhu H, Zhen W, et al. Genistein inhibits TNF-alpha-induced endothelial inflammation through the protein kinase pathway a and improves vascular inflammation in C57BL/6 mice. International Journal of Cardiology. 2013;168(3):2637-2645
  126. 126. Qureshi AA, Guan XQ , Reis JC, Papasian CJ, Jabre S, Morrison DC, et al. Inhibition of nitric oxide and inflammatory cytokines in LPS-stimulated murine macrophages by resveratrol, a potent proteasome inhibitor. Lipids in Health and Disease. 2012;11:76
  127. 127. Kumar H, Lim HW, More SV, Kim BW, Koppula S, Kim IS, et al. The role of free radicals in the aging brain and Parkinson’s disease: Convergence and parallelism. International Journal of Molecular Sciences. 2012;13(8):10478-10504
  128. 128. Floyd RA, Carney JM. Free radical damage to protein and DNA: Mechanisms involved and relevant observations on brain undergoing oxidative stress. Annals of Neurology. 1992;32(Suppl):S22-S27
  129. 129. Kwon KJ, Kim HJ, Shin CY, Han SH. Melatonin potentiates the neuroprotective properties of resveratrol against beta-amyloid-induced neurodegeneration by modulating AMP-activated protein kinase pathways. Journal of Clinical Neurology. 2010;6(3):127-137
  130. 130. Schinella G, Mosca S, Cienfuegos-Jovellanos E, Ángeles Pasamar M, Muguerza B, Ramón D, et al. Antioxidant properties of polyphenol-rich cocoa products industrially processed. Food Research International. 2010;43(6):1614-1623
  131. 131. Xiang W, Schlachetzki JC, Helling S, Bussmann JC, Berlinghof M, Schaffer TE, et al. Oxidative stress-induced posttranslational modifications of alpha-synuclein: Specific modification of alpha-synuclein by 4-hydroxy-2-nonenal increases dopaminergic toxicity. Molecular and Cellular Neurosciences. 2013;54:71-83
  132. 132. Frei B, Higdon JV. Antioxidant activity of tea polyphenols in vivo: Evidence from animal studies. The Journal of Nutrition. 2003;133(10):3275s-3284s
  133. 133. Bastianetto S, Krantic S, Chabot JG, Quirion R. Possible involvement of programmed cell death pathways in the neuroprotective action of polyphenols. Current Alzheimer Research. 2011;8(5):445-451
  134. 134. Zhang K, Ma Z, Wang J, Xie A, Xie J. Myricetin attenuated MPP(+)-induced cytotoxicity by anti-oxidation and inhibition of MKK4 and JNK activation in MES23.5 cells. Neuropharmacology. 2011;61(1-2):329-335
  135. 135. van Loo G, Saelens X, van Gurp M, MacFarlane M, Martin SJ, Vandenabeele P. The role of mitochondrial factors in apoptosis: A Russian roulette with more than one bullet. Cell Death and Differentiation. 2002;9(10):1031-1042
  136. 136. Lev N, Melamed E, Offen D. Apoptosis and Parkinson’s disease. Progress in Neuro-Psychopharmacology & Biological Psychiatry. 2003;27(2):245-250
  137. 137. Camilleri A, Zarb C, Caruana M, Ostermeier U, Ghio S, Hogen T, et al. Mitochondrial membrane permeabilisation by amyloid aggregates and protection by polyphenols. Biochimica et Biophysica Acta. 2013;1828(11):2532-2543
  138. 138. Kumar A, Prakash A, Dogra S. Naringin alleviates cognitive impairment, mitochondrial dysfunction and oxidative stress induced by D-galactose in mice. Food and Chemical Toxicology. 2010;48(2):626-632
  139. 139. Gao QG, Xie JX, Wong MS, Chen WF. IGF-I receptor signaling pathway is involved in the neuroprotective effect of genistein in the neuroblastoma SK-N-SH cells. European Journal of Pharmacology. 2012;677(1-3):39-46
  140. 140. Jeon SM, Bok SH, Jang MK, Lee MK, Nam KT, Park YB, et al. Antioxidative activity of naringin and lovastatin in high cholesterol-fed rabbits. Life Sciences. 2001;69(24):2855-2866
  141. 141. Renault TT, Teijido O, Antonsson B, Dejean LM, Manon S. Regulation of Bax mitochondrial localization by Bcl-2 and Bcl-x(L): Keep your friends close but your enemies closer. The International Journal of Biochemistry & Cell Biology. 2013;45(1):64-67
  142. 142. Perry EK, Tomlinson BE, Blessed G, Bergmann K, Gibson PH, Perry RH. Correlation of cholinergic abnormalities with senile plaques and mental test scores in senile dementia. British Medical Journal. 1978;2(6150):1457-1459
  143. 143. Fisher A. Cholinergic treatments with emphasis on m1 muscarinic agonists as potential disease-modifying agents for Alzheimer’s disease. Neurotherapeutics. 2008;5(3):433-442
  144. 144. Lahiri DK, Rogers JT, Greig NH, Sambamurti K. Rationale for the development of cholinesterase inhibitors as anti-Alzheimer agents. Current Pharmaceutical Design. 2004;10(25):3111-3119
  145. 145. Zhang L, Cao H, Wen J, Xu M. Green tea polyphenol (−)-epigallocatechin-3-gallate enhances the inhibitory effect of huperzine a on acetylcholinesterase by increasing the affinity with serum albumin. Nutritional Neuroscience. 2009;12(4):142-148
  146. 146. Zaky A, Mohammad B, Moftah M, Kandeel KM, Bassiouny AR. Apurinic/apyrimidinic endonuclease 1 is a key modulator of aluminum-induced neuroinflammation. BMC Neuroscience. 2013;14:26
  147. 147. Ohta K, Mizuno A, Ueda M, Li S, Suzuki Y, Hida Y, et al. Autophagy impairment stimulates PS1 expression and gamma-secretase activity. Autophagy. 2010;6(3):345-352
  148. 148. Toni M, Massimino ML, De Mario A, Angiulli E, Spisni E. Metal dyshomeostasis and their pathological role in prion and prion-like diseases: The basis for a nutritional approach. Frontiers in Neuroscience. 2017;11:3
  149. 149. Yang F, Lim GP, Begum AN, Ubeda OJ, Simmons MR, Ambegaokar SS, et al. Curcumin inhibits formation of amyloid beta oligomers and fibrils, binds plaques, and reduces amyloid in vivo. The Journal of Biological Chemistry. 2005;280(7):5892-5901
  150. 150. Garcia-Alloza M, Borrelli LA, Rozkalne A, Hyman BT, Bacskai BJ. Curcumin labels amyloid pathology in vivo, disrupts existing plaques, and partially restores distorted neurites in an Alzheimer mouse model. Journal of Neurochemistry. 2007;102(4):1095-1104
  151. 151. Cornejo A, Aguilar Sandoval F, Caballero L, Machuca L, Munoz P, Caballero J, et al. Rosmarinic acid prevents fibrillization and diminishes vibrational modes associated to beta sheet in tau protein linked to Alzheimer’s disease. Journal of Enzyme Inhibition and Medicinal Chemistry. 2017;32(1):945-953
  152. 152. Duff K, Kuret J, Congdon EE. Disaggregation of tau as a therapeutic approach to tauopathies. Current Alzheimer Research. 2010;7(3):235-240
  153. 153. Wang J, Santa-Maria I, Ho L, Ksiezak-Reding H, Ono K, Teplow DB, et al. Grape derived polyphenols attenuate tau neuropathology in a mouse model of Alzheimer’s disease. Journal of Alzheimer’s Disease. 2010;22(2):653-661
  154. 154. Gueroux M, Fleau C, Slozeck M, Laguerre M, Pianet I. Epigallocatechin 3-Gallate as an inhibitor of tau phosphorylation and aggregation: A molecular and structural insight. The Journal of Prevention of Alzheimer’s Disease. 2017;4(4):218-225
  155. 155. Rezai-Zadeh K, Shytle D, Sun N, Mori T, Hou H, Jeanniton D, et al. Green tea epigallocatechin-3-gallate (EGCG) modulates amyloid precursor protein cleavage and reduces cerebral amyloidosis in Alzheimer transgenic mice. The Journal of Neuroscience. 2005;25(38):8807-8814
  156. 156. Ehrnhoefer DE, Bieschke J, Boeddrich A, Herbst M, Masino L, Lurz R, et al. EGCG redirects amyloidogenic polypeptides into unstructured, off-pathway oligomers. Nature Structural & Molecular Biology. 2008;15(6):558-566
  157. 157. Yang JE, Rhoo KY, Lee S, Lee JT, Park JH, Bhak G, et al. EGCG-mediated protection of the membrane disruption and cytotoxicity caused by the ‘Active Oligomer’ of alpha-synuclein. Scientific Reports. 2017;7(1):17945
  158. 158. Palhano FL, Lee J, Grimster NP, Kelly JW. Toward the molecular mechanism(s) by which EGCG treatment remodels mature amyloid fibrils. Journal of the American Chemical Society. 2013;135(20):7503-7510
  159. 159. Ardah MT, Paleologou KE, Lv G, Abul Khair SB, Kazim AS, Minhas ST, et al. Structure activity relationship of phenolic acid inhibitors of alpha-synuclein fibril formation and toxicity. Frontiers in Aging Neuroscience. 2014;6:197
  160. 160. Pehlivan SB. Nanotechnology-based drug delivery systems for targeting, imaging and diagnosis of neurodegenerative diseases. Pharmaceutical Research. 2013;30(10):2499-2511
  161. 161. Guerra M, Blazquez JL, Rodriguez EM. Blood-brain barrier and foetal-onset hydrocephalus, with a view on potential novel treatments beyond managing CSF flow. Fluids Barriers CNS. 2017;14(1):19
  162. 162. Ballabh P, Braun A, Nedergaard M. The blood-brain barrier: An overview: Structure, regulation, and clinical implications. Neurobiology of Disease. 2004;16(1):1-13
  163. 163. Banks WA. From blood-brain barrier to blood-brain interface: New opportunities for CNS drug delivery. Nature Reviews. Drug Discovery. 2016;15(4):275-292
  164. 164. Komarova YA, Kruse K, Mehta D, Malik AB. Protein interactions at endothelial junctions and signaling mechanisms regulating endothelial permeability. Circulation Research. 2017;120(1):179-206
  165. 165. Sweeney MD, Sagare AP, Zlokovic BV. Blood-brain barrier breakdown in Alzheimer disease and other neurodegenerative disorders. Nature Reviews. Neurology. 2018;14(3):133-150
  166. 166. Zlokovic BV. Neurovascular pathways to neurodegeneration in Alzheimer’s disease and other disorders. Nature Reviews. Neuroscience. 2011;12(12):723-738
  167. 167. Zhao Z, Nelson AR, Betsholtz C, Zlokovic BV. Establishment and dysfunction of the blood-brain barrier. Cell. 2015;163(5):1064-1078
  168. 168. Hultman K, Strickland S, Norris EH. The APOE varepsilon4/varepsilon4 genotype potentiates vascular fibrin(ogen) deposition in amyloid-laden vessels in the brains of Alzheimer’s disease patients. Journal of Cerebral Blood Flow and Metabolism. 2013;33(8):1251-1258
  169. 169. Pienaar IS, Lee CH, Elson JL, McGuinness L, Gentleman SM, Kalaria RN, et al. Deep-brain stimulation associates with improved microvascular integrity in the subthalamic nucleus in Parkinson’s disease. Neurobiology of Disease. 2015;74:392-405
  170. 170. Sweeney MD, Ayyadurai S, Zlokovic BV. Pericytes of the neurovascular unit: Key functions and signaling pathways. Nature Neuroscience. 2016;19(6):771-783
  171. 171. Cullen KM, Kocsi Z, Stone J. Pericapillary haem-rich deposits: Evidence for microhaemorrhages in aging human cerebral cortex. Journal of Cerebral Blood Flow and Metabolism. 2005;25(12):1656-1667
  172. 172. Gray MT, Woulfe JM. Striatal blood-brain barrier permeability in Parkinson’s disease. Journal of Cerebral Blood Flow and Metabolism. 2015;35(5):747-750
  173. 173. Wada K, Arai H, Takanashi M, Fukae J, Oizumi H, Yasuda T, et al. Expression levels of vascular endothelial growth factor and its receptors in Parkinson’s disease. Neuroreport. 2006;17(7):705-709
  174. 174. Desai Bradaric B, Patel A, Schneider JA, Carvey PM, Hendey B. Evidence for angiogenesis in Parkinson’s disease, incidental Lewy body disease, and progressive supranuclear palsy. Journal of Neural Transmission (Vienna). 2012;119(1):59-71
  175. 175. Boocock DJ, Faust GE, Patel KR, Schinas AM, Brown VA, Ducharme MP, et al. Phase I dose escalation pharmacokinetic study in healthy volunteers of resveratrol, a potential cancer chemopreventive agent. Cancer Epidemiology, Biomarkers & Prevention. 2007;16(6):1246-1252
  176. 176. Almeida L, Vaz-da-Silva M, Falcao A, Soares E, Costa R, Loureiro AI, et al. Pharmacokinetic and safety profile of trans-resveratrol in a rising multiple-dose study in healthy volunteers. Molecular Nutrition & Food Research. 2009;53(Suppl 1):S7-S15
  177. 177. Ringman JM, Frautschy SA, Teng E, Begum AN, Bardens J, Beigi M, et al. Oral curcumin for Alzheimer’s disease: Tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimer’s Research & Therapy. 2012;4(5):43
  178. 178. Crowe TP, Greenlee MHW, Kanthasamy AG, Hsu WH. Mechanism of intranasal drug delivery directly to the brain. Life Sciences. 2018;195:44-52
  179. 179. Hanson LR, Frey WH 2nd. Intranasal delivery bypasses the blood-brain barrier to target therapeutic agents to the central nervous system and treat neurodegenerative disease. BMC Neuroscience. 2008;9(Suppl. 3):S5
  180. 180. Chapman CD, Frey WH 2nd, Craft S, Danielyan L, Hallschmid M, Schioth HB, et al. Intranasal treatment of central nervous system dysfunction in humans. Pharmaceutical Research. 2013;30(10):2475-2484
  181. 181. Aura AM, O’Leary KA, Williamson G, Ojala M, Bailey M, Puupponen-Pimia R, et al. Quercetin derivatives are deconjugated and converted to hydroxyphenylacetic acids but not methylated by human fecal flora in vitro. Journal of Agricultural and Food Chemistry. 2002;50(6):1725-1730
  182. 182. Shortt C, Hasselwander O, Meynier A, Nauta A, Fernandez EN, Putz P, et al. Systematic review of the effects of the intestinal microbiota on selected nutrients and non-nutrients. European Journal of Nutrition. 2018;57(1):25-49
  183. 183. El-Mohsen MA, Bayele H, Kuhnle G, Gibson G, Debnam E, Srai SK, et al. Distribution of [3H]trans-resveratrol in rat tissues following oral administration. The British Journal of Nutrition. 2006;96(1):62-70
  184. 184. Wenzel E, Somoza V. Metabolism and bioavailability of trans-resveratrol. Molecular Nutrition & Food Research. 2005;49(5):472-481
  185. 185. Delmas D, Aires V, Limagne E, Dutartre P, Mazue F, Ghiringhelli F, et al. Transport, stability, and biological activity of resveratrol. Annals of the New York Academy of Sciences. 2011;1215:48-59
  186. 186. Hu B, Liu X, Zhang C, Zeng X. Food macromolecule based nanodelivery systems for enhancing the bioavailability of polyphenols. Journal of Food and Drug Analysis. 2017;25(1):3-15
  187. 187. Aqil F, Munagala R, Jeyabalan J, Vadhanam MV. Bioavailability of phytochemicals and its enhancement by drug delivery systems. Cancer Letters. 2013;334(1):133-141
  188. 188. Irimie AI, Sonea L, Jurj A, Mehterov N, Zimta AA, Budisan L, et al. Future trends and emerging issues for nanodelivery systems in oral and oropharyngeal cancer. International Journal of Nanomedicine. 2017;12:4593-4606
  189. 189. Ader P, Wessmann A, Wolffram S. Bioavailability and metabolism of the flavonol quercetin in the pig. Free Radical Biology & Medicine. 2000;28(7):1056-1067
  190. 190. Brand W, Padilla B, van Bladeren PJ, Williamson G, Rietjens IM. The effect of co-administered flavonoids on the metabolism of hesperetin and the disposition of its metabolites in Caco-2 cell monolayers. Molecular Nutrition & Food Research. 2010;54(6):851-860
  191. 191. Bonferoni MC, Rossi S, Sandri G, Ferrari F, Gavini E, Rassu G, et al. Nanoemulsions for “nose-to-brain” drug delivery. Pharmaceutics. 2019;11(2):1-17
  192. 192. Squillaro T, Schettino C, Sampaolo S, Galderisi U, Di Iorio G, Giordano A, et al. Adult-onset brain tumors and neurodegeneration: Are polyphenols protective? Journal of Cellular Physiology. 2018;233(5):3955-3967
  193. 193. Figueira I, Garcia G, Pimpao RC, Terrasso AP, Costa I, Almeida AF, et al. Polyphenols journey through blood-brain barrier towards neuronal protection. Scientific Reports. 2017;7(1):11456
  194. 194. Squillaro T, Cimini A, Peluso G, Giordano A, Melone MAB. Nano-delivery systems for encapsulation of dietary polyphenols: An experimental approach for neurodegenerative diseases and brain tumors. Biochemical Pharmacology. 2018;154:303-317
  195. 195. Conte R. Polyphenols Nanoencapsulation for therapeutic applications. Journal of Biomolecular Research & Therapeutics. 2019;5:2
  196. 196. Shive MS, Anderson JM. Biodegradation and biocompatibility of PLA and PLGA microspheres. Advanced Drug Delivery Reviews. 1997;28(1):5-24
  197. 197. Anton N, Benoit JP, Saulnier P. Design and production of nanoparticles formulated from nano-emulsion templates-a review. Journal of Controlled Release. 2008;128(3):185-199
  198. 198. Mora-Huertas CE, Fessi H, Elaissari A. Polymer-based nanocapsules for drug delivery. International Journal of Pharmaceutics. 2010;385(1-2):113-142
  199. 199. Radtchenko I, Sukhorukov G, Mohwald H. A novel method for encapsulation of poorly water-soluble drugs: Precipitation in polyelectrolyte multilayer shells. International Journal of Pharmaceutics. 2002;242(1-2):219-223
  200. 200. Khoee S, Yaghoobian M. An investigation into the role of surfactants in controlling particle size of polymeric nanocapsules containing penicillin-G in double emulsion. European Journal of Medicinal Chemistry. 2009;44(6):2392-2399
  201. 201. Comfort C, Garrastazu G, Pozzoli M, Sonvico F. Opportunities and challenges for the nasal administration of nanoemulsions. Current Topics in Medicinal Chemistry. 2015;15(4):356-368
  202. 202. Mahajan HS, Mahajan MS, Nerkar PP, Agrawal A. Nanoemulsion-based intranasal drug delivery system of saquinavir mesylate for brain targeting. Drug Delivery. 2014;21(2):148-154
  203. 203. Tiwari G, Tiwari R, Rai AK. Cyclodextrins in delivery systems: Applications. Journal of Pharmacy & Bioallied Sciences. 2010;2(2):72-79
  204. 204. Szejtli J. Introduction and general overview of cyclodextrin chemistry. Chemical Reviews. 1998;98(5):1743-1754
  205. 205. Stella VJ, He Q. Cyclodextrins. Toxicologic Pathology. 2008;36(1):30-42
  206. 206. Sercombe L, Veerati T, Moheimani F, Wu SY, Sood AK, Hua S. Advances and challenges of liposome assisted drug delivery. Frontiers in Pharmacology. 2015;6:286
  207. 207. Torchilin VP. Structure and design of polymeric surfactant-based drug delivery systems. Journal of Controlled Release. 2001;73(2-3):137-172
  208. 208. Chen Y, Zhang X, Lu J, Huang Y, Li J, Li S. Targeted delivery of curcumin to tumors via PEG-derivatized FTS-based micellar system. The AAPS Journal. 2014;16(3):600-608
  209. 209. Sood S, Jain K, Gowthamarajan K. Optimization of curcumin nanoemulsion for intranasal delivery using design of experiment and its toxicity assessment. Colloids and Surfaces. B, Biointerfaces. 2014;113:330-337
  210. 210. Ono K, Hasegawa K, Naiki H, Yamada M. Curcumin has potent anti-amyloidogenic effects for Alzheimer’s beta-amyloid fibrils in vitro. Journal of Neuroscience Research. 2004;75(6):742-750
  211. 211. Agrawal R, Mishra B, Tyagi E, Nath C, Shukla R. Effect of curcumin on brain insulin receptors and memory functions in STZ (ICV) induced dementia model of rat. Pharmacological Research. 2010;61(3):247-252
  212. 212. Tiwari SK, Agarwal S, Seth B, Yadav A, Nair S, Bhatnagar P, et al. Curcumin-loaded nanoparticles potently induce adult neurogenesis and reverse cognitive deficits in Alzheimer’s disease model via canonical Wnt/beta-catenin pathway. ACS Nano. 2014;8(1):76-103
  213. 213. Taylor M, Moore S, Mourtas S, Niarakis A, Re F, Zona C, et al. Effect of curcumin-associated and lipid ligand-functionalized nanoliposomes on aggregation of the Alzheimer’s Abeta peptide. Nanomedicine. 2011;7(5):541-550
  214. 214. Sandhir R, Yadav A, Mehrotra A, Sunkaria A, Singh A, Sharma S. Curcumin nanoparticles attenuate neurochemical and neurobehavioral deficits in experimental model of Huntington’s disease. Neuromolecular Medicine. 2014;16(1):106-118
  215. 215. Kundu P, Das M, Tripathy K, Sahoo SK. Delivery of dual drug loaded lipid based nanoparticles across the blood-brain barrier impart enhanced neuroprotection in a rotenone induced mouse model of Parkinson’s disease. ACS Chemical Neuroscience. 2016;7(12):1658-1670
  216. 216. Braidy N, Jugder BE, Poljak A, Jayasena T, Mansour H, Nabavi SM, et al. Resveratrol as a potential therapeutic candidate for the treatment and management of Alzheimer’s disease. Current Topics in Medicinal Chemistry. 2016;16(17):1951-1960
  217. 217. Singh G, Pai RS. Optimized PLGA nanoparticle platform for orally dosed trans-resveratrol with enhanced bioavailability potential. Expert Opinion on Drug Delivery. 2014;11(5):647-659
  218. 218. Shao J, Li X, Lu X, Jiang C, Hu Y, Li Q , et al. Enhanced growth inhibition effect of resveratrol incorporated into biodegradable nanoparticles against glioma cells is mediated by the induction of intracellular reactive oxygen species levels. Colloids and Surfaces. B, Biointerfaces. 2009;72(1):40-47
  219. 219. Pangeni R, Sharma S, Mustafa G, Ali J, Baboota S. Vitamin E loaded resveratrol nanoemulsion for brain targeting for the treatment of Parkinson’s disease by reducing oxidative stress. Nanotechnology. 2014;25(48):485102
  220. 220. Chongtham A, Agrawal N. Curcumin modulates cell death and is protective in Huntington’s disease model. Scientific Reports. 2016;6:18736
  221. 221. Nasr M. Development of an optimized hyaluronic acid-based lipidic nanoemulsion co-encapsulating two polyphenols for nose to brain delivery. Drug Delivery. 2016;23(4):1444-1452

Written By

Patrizia Polverino de Laureto, Luana Palazzi and Laura Acquasaliente

Submitted: 08 May 2019 Reviewed: 06 September 2019 Published: 23 October 2019