Next Article in Journal
Protective Effect of Oligonol on Dimethylnitrosamine-Induced Liver Fibrosis in Rats via the JNK/NF-κB and PI3K/Akt/Nrf2 Signaling Pathways
Previous Article in Journal
4-Hydroxynonenal Contributes to Fibroblast Senescence in Skin Photoaging Evoked by UV-A Radiation
Previous Article in Special Issue
Two Distinct Faces of Vitamin C: AA vs. DHA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Beyond Antioxidant Effects: Nature-Based Templates Unveil New Strategies for Neurodegenerative Diseases

by
Andrea Bacci
1,†,
Massimiliano Runfola
1,†,
Simona Sestito
2 and
Simona Rapposelli
1,*
1
Department of Pharmacy, University of Pisa, Via Bonanno 6, 56126 Pisa, Italy
2
Department of Chemistry and Pharmacy, University of Sassari, Via Vienna 2, 07100 Sassari, Italy
*
Author to whom correspondence should be addressed.
These authors (A.B. and M.R.) contributed equally to this work.
Antioxidants 2021, 10(3), 367; https://doi.org/10.3390/antiox10030367
Submission received: 18 January 2021 / Revised: 11 February 2021 / Accepted: 24 February 2021 / Published: 28 February 2021
(This article belongs to the Special Issue Oxidative Stress and Inflammation in the Nervous System)

Abstract

:
The complex network of malfunctioning pathways occurring in the pathogenesis of neurodegenerative diseases (NDDs) represents a huge hurdle in the development of new effective drugs to be used in therapy. In this context, redox reactions act as crucial regulators in the maintenance of neuronal microenvironment homeostasis. Particularly, their imbalance results in the severe compromising of organism’s natural defense systems and subsequently, in the instauration of deleterious OS, that plays a fundamental role in the insurgence and progress of NDDs. Despite the huge efforts in drug discovery programs, the identification process of new therapeutic agents able to counteract the relentless progress of neurodegenerative processes has produced low or no effective therapies. Consequently, a paradigm-shift in the drug discovery approach for these diseases is gradually occurring, paving the way for innovative therapeutical approaches, such as polypharmacology. The aim of this review is to provide an overview of the main pharmacological features of most promising nature-based scaffolds for a possible application in drug discovery, especially for NDDs, highlighting their multifaceted effects against OS and neuronal disorders.

1. Introduction

Neurodegenerative diseases (NDDs) consist of a broad class of pathological conditions characterized by a progressive and irreversible degeneration of the nervous tissue, mainly occurring in the elderly. As life expectancy increases, the lack of efficient treatments able to halt or slow down neurodegenerative processes is turning NDDs into a huge socio-economic challenge for healthcare systems. At the current trends, the global population of those over 60 years of age is forecasted to exceed 2 billion by 2050, with a consequent increase of NDDs, such as Alzheimer’s (AD), Parkinson (PD), Huntington’s Disease (HD), and amyotrophic lateral sclerosis (ALS) [1]. Despite the vast heterogeneity of NDDs clinical phenotypes, high levels of oxidative stress (OS) have been identified as a prominent hallmark in their pathogenesis [2]. Beside lipid peroxidation products (e.g., acrolein) that have been found in high levels in patients affected by NDDs, high levels of other OS biomarkers such as malondialdehyde (MDA) and 4-hydroxynonenal (HME) have been detected in PD, while high levels of 8-hydroxy-2′-deoxyguanosine (8-OHdG) and nitrotyrosine characterize HD; also, ALS tissues have shown high concentrations of of dihydropyrimidinase-related protein 2 (DRP-2), heat shock protein 70, and α-enolase [3].
OS is well-known to play a key role in premature ageing, contributing to the progressive loss of tissue and organ function. At the same time, OS is implicated in several age-related conditions, including cancer, cardiovascular, neurodegenerative, and inflammatory diseases. This phenomenon rises from an imbalance between oxidative and reductive processes occurring during physiological metabolism [4,5]. In pathological conditions reactive oxygen species (ROS) could disrupt cells’ membranes and deeply damage cellular components through a series of peroxidative reactions (e.g., lipid peroxidation) [6]. Moreover, ROS can induce oxidative modifications of proteins and DNA damage. The central nervous system (CNS) is particularly vulnerable to these events due to the large amount of oxygen required for neuronal metabolism and the high concentration of lipids prone to peroxidation that constitute neuronal membranes. Much evidence has also associated cellular ageing to harmful oxidative events induced by metabolic waste products and their accumulation, along with a progressive inefficiency of physiological defence and repair systems (e.g., autophagy) [7]. Furthermore, ageing is associated with gene mutations that lead to cellular malfunctions such as mitochondrial and cellular genome variations, reduced protein biosynthesis, and lipid accumulation [8]. Interestingly, ROS have a prominent role in all these processes. Indeed, further studies performed in animal models reveal the existence of strong interconnection between ROS and metallostasis alterations, protein aggregation, and mitochondrial failures that characterize NDDs. Over the years, the research of new anti-ageing and nature-based agents able to counteract cellular senescence and neurodegeneration, has focused to the identification of small molecules with antioxidant activity; unfortunately, their therapeutic potential seems to be limited, showing that targeting ROS accumulation did not lead to the expected results [9]. Additionally, severe side-effects and poor blood brain barrier permeability are few of the main limits that have undermined their translatability in human model and consequently, their potential therapeutic application [10]. Hence, a novel and paradigm-shifting view on NDDs seems necessary to achieve better treatments of these conditions. Particularly, identification of treatments targeting several pathogenic pathways—i.e., polypharmacology—has emerged as a new pharmaceutical strategy that could propel beneficial effects counteracting the multi-faceted impairment of NDDs. A huge contribute in designing and developing polypharmacological compounds may come from nature. Indeed, a large number of activities—such as antioxidant, chelating, and anti-inflammatory—have been associated to natural compounds [11]. Furthermore, recent studies have proved that many natural compounds are able to regulate autophagy and proteasomal degradation pathways and to curb protein misfolding [12,13]. On this basis, combination of the wide-range benefits provided by natural compounds could represent an attractive therapeutic strategy in order to obtain an effective multi-functional treatment in preventing neurodegeneration. The aim of this review is to provide an overview of the main pharmacological features of most promising natural antioxidants with a potential therapeutic application in the drug discovery process against NDDs.

2. Crucial Role of ROS in Physiological and Pathological Mechanisms

High rates of oxygen metabolism, along with the abundant distribution of redox-active metals and polyunsaturated fatty acids, are some critical points in the neuronal microenvironment [8]. The two unpaired electrons on the O2 have high reactivity to form a group of free radicals called ROS. These species can originate from exogenous sources (irradiations or chemicals), although the main source of endogenous ROS remains the respiratory chain of mitochondria and nicotinamide adenine dinucleotide phosphate (NADPH) oxidase enzyme (NOX). The mitochondrial electron transfer chain (ETC) consists of five complexes able to modulate ROS production [12,13]. Besides NOX enzyme, further enzymatic systems and cellular organelles are involved in ROS production: xanthine oxidases (XO) transfer electrons to O2 generating the superoxide O2•− and H2O2; the endoplasmic reticulum (ER) may produce ROS as by-products; and peroxisomes, containing several oxidases enzymes, could increase the release of H2O2 in the cytosol [14]. Collectively, these mechanisms generate O2•− which does not directly interact with the cellular substrates, but has a crucial role in free radicals production involved in maintaining physiological functions, including proliferation, defense against infectious agents, signal transduction, and gene expression [13].
Physiologically, low and beneficial ROS levels are preserved through the action of several antioxidant enzymes. For example, superoxide dismutase (SOD) inactivates O2•− by conversion into H2O2, that can be then removed by catalases and glutathione peroxidases (GPX), producing H2O and O2. Tonal levels of ROS have a wide range of significant effects in physiological cellular signaling and survival mechanisms. For example, ROS can activate mitogen-activated protein kinases (MAPKs) that constitute a crucial pathway in cardiovascular system [15]. Other important pro-survival transcription factors—such as NF-E2-related factor 2 (Nrf2) and nuclear factor-κB (NF-κB)—are influenced by ROS levels [16]. An imbalance between ROS production and antioxidant defenses contributes to the insurgence of OS, concurring to cellular disfunctions, ageing, and neurodegeneration [17]. Indeed, chronic OS is a well-known pathogenic factor with a prominent role in the etiology of several NDDs. Many peculiar cellular disfunctions and pathological phenomena in neurodegenerative tissues are due to ROS accumulation. Additionally, ROS can promote DNA, RNA, and protein oxidation, lipid peroxidation, mitochondrial failure, and protein aggregation. The overwhelming production or the shortage of ROS levels are both deleterious for cellular homeostasis and play an important role in the malfunctioning process of mitochondria, cells and organisms [14,15]. ROS levels could also affect non-cell autonomous effects, which contribute to the neurodegenerative process mainly through an uncontrolled activation of inflammatory status and immune response. Both neurons’ environmental factors, such as increased OS, either endogenous factors—e.g., protein aggregates accumulation—induce an over-activation of the innate immune cells in the CNS, such as microglia and astrocytes. This uncontrolled inflammatory status results in the production of neurotoxic factors that contribute to amplify the disease states and actively participate in the pathological vicious circle [16].

2.1. ROS and Mitochondrial Dysfunctions

Under normal conditions, ROS are finely regulated, and their concentration kept under control: only about 2% of oxygen consumption in mitochondria is converted to ROS. Recently, Aon et al. proved that this percentage could fluctuate from 0.25% to 11% when the endogenous antioxidant systems are disabled [17]. Basal levels of ROS initiate and coordinate several pathways finely tuned with cellular demands. For example, ROS are required for physiological regulation in cycle progression and their levels are associated with cell differentiation, migration, and proliferation; immune response; and apoptosis [18,19]. Hereby, oxidative homeostasis appears to be fundamental for normal mitochondrial functions. Mitochondria are considered the energy powerhouse of cells. Indeed, they are the main source of cellular ATP, and play important roles in ion homeostasis, metabolic pathways, apoptosis, and in ROS production and consumption [20]. Increased levels of oxidative damage contribute to metabolic stress and cellular injuries, and the mitochondrial DNA (mDNA) represents a critical target for such reactive species. The mitochondrial genome is located close to the inner membrane where the respiratory machinery produces ROS, making mDNA a primary target of ROS and therefore highly exposed to damage, deletion, and mutation [21]. Once mDNA is damaged, mitochondria can raise OS and enter a vicious cycle that led to an increase of ROS production. This phenomenon is called “ROS-induced ROS release” (RIRR) and results in a boosted superoxide production contributing to metabolic OS, genome instability, and cellular lesions [22]. ROS abundance increases genomic instability; several studies confirmed indeed that a marked ROS accumulation is often associated with an increased rate of DNA mutations in mice [14], rats [15], nematodes [23] and humans [24]. Besides mDNA, mitochondrial membranes are highly exposed to ROS-induced damages. The inner membrane is the site where ETC and phosphorylation happen; under physiological conditions its permeability is tightly regulated and allows exclusively the permeation of small neutral molecules. Redox stress may induce the opening of membrane anion channel (IMAC) and mitochondrial permeability transition pore (mPTP), both responsible for mitochondria swelling, thus leading to the collapse of mitochondrial bioenergetic functions and ultimately to cellular death [25]. In this process, ROS are involved in activating IMAC and mPTP; consequently, mitochondria undergo an increase in ROS production through a positive-feedback mechanism [26]. Moreover, high levels of ROS lead to an uncontrolled peroxidation of lipids and phospholipids that constitute cellular membrane, thus altering its biophysical properties, such as fluidity and permeability. These biophysical ROS-induced alterations can also impair the activity of various transporters and respiratory proteins within the inner membrane [27]. Oxidation of thiol moieties of the adenine nucleotide translocator located on mPTP may also promote variations in mitochondrial permeability [28]. Finally, many other enzymes (i.e., nicotinamide adenine dinucleotide dehydrogenase and succinate dehydrogenase) are vulnerable to oxidative alterations of their iron-sulfur centers, and their impairment could induce malfunctions of the ETC [29]. Alterations in membrane permeability promote a change in mitochondrial potential linked to disruption of Ca2+ homeostasis and the consequent overproduction of O2•− [30]. The increased Ca2+ levels could also promote osmotic swelling and the break of mitochondrial membrane, mPTP’s opening, and thus apoptosis [31]. Altogether, high levels of ROS impair cellular homeostasis, leading to harmful effects on biomolecules that ultimately contribute to cellular senescence and, more importantly, to the pathogenesis and progression of neurodegenerative diseases [32,33,34,35].

2.2. Metal Accumulation, ROS Production, and Protein Misfolding

Along with ROS, metals play a significant role in tuning enzymes activity. Alteration in metal homeostasis, namely metallostasis, is an additional key factor in OS insurgence, leading to cellular ageing and NDDs. For decades metals accumulation has been studied as a prominent mechanism in neuro- and cardiotoxicity, generation of free radicals, lipid peroxidation, and protein aggregation [36,37]. Essential biometals—like iron, copper, zinc, manganese, calcium, magnesium—and non-essential metals—like aluminium, lead, and mercury—seem to play a crucial role in the development and progression of neurodegenerative diseases [38,39]. Indeed, multi-valence metals can support ROS cascade determining an enhanced generation of toxic radicals. For example, iron, copper, aluminium, and zinc can take part in Fenton and Haber–Weiss reactions: the first is defined as the reaction between Fe2+ and H2O2 generating Fe3+ and hydroxyl radical, then converted to O2•−; the latter forms hydroxyl radical and hydroxyl anion from reaction of O2•− with H2O2 [22,40,41]. Therefore, an accumulation of metals can generate and support ROS overproduction, overwhelming the endogenous antioxidant defence systems and contributing to OS insurgence [42]. Additionally, metals can exert direct damages to biostructures, such as proteins, interacting with some specific structural regions and destabilizing their native conformation. Multivalent metals can bridge together charged aminoacidic residues of proteins, especially when these residues are prominently exposed like in compromised or misfolded proteins [43]. The correlation between metals and protein aggregation has been widely studied and many evidence suggest that metals accelerate the cytotoxicity of proteins deposition [44,45]. Accordingly, metals accumulation in brain and altered metallostasis are both hallmarks of neurodegenerative disorders. However, it remains unclear if metals accumulation is a leading cause in the onset of NDDs or the result of previous dysfunctions [46]. Nevertheless, among the strategies pursued to limit OS injury and the consequent neurodegeneration, metal chelation appears as an attractive tool to counteract metal-induced cellular damages [47]. To this end, chelating agents have been widely studied and their therapeutic role in CNS disorders and cellular aging continuously investigated [48].

2.3. OS and Protein Misfolding/Accumulation

Another prominent hallmark of many age-related diseases is the impairment of protein homeostasis. Proteostasis dysregulation leads to an accumulation of toxic misfolded and abnormal protein aggregates associated with common NDDs such as AD, PD, ALS, and frontotemporal dementia (FTD). Proteostasis is maintained by proteolytic machineries and their regulators; all of them operate through an extensive monitoring network that makes use of degradation pathways to clear toxic misfolded proteins. During ageing, the efficiency of this network declines over time, leading to a progressive proteostasis imbalance. There are strong evidence indicating that two of these protein degradation systems, the ubiquitin–proteasome system (UPS) and the autophagy–lysosome pathway (ALP) become progressively compromised with ageing [49]. Actually, the impairment of these degradation systems promotes an accumulation of misfolded and abnormal proteins, including amyloid-beta (Aβ) in AD, huntingtin (HTT) in HD, superoxide dismutase-1 (SOD1) and TAR DNA-binding protein 43 (TDP-43) in ALS, and α-synuclein in PD [50]. Proteostasis impairment mostly occurs through irreversible oxidative modifications of different proteins, mainly induced by OS [51]. Beyond the existence of a cross-talk between ALP and UPS, recent biochemical studies suggest a direct interplay of the proteolytic machinery with OS, thus contributing to the aetiology and the progression of neuropathogenic cellular conditions [52]. As a matter of fact, UPS plays a key role in physiological protein turnover, recognizing and degrading insoluble, damaged, and oxidized proteins, preventing their cellular accumulation. OS could aggravate the ageing-related decline of misfolded proteins clearance, entering a self-propagating cycle of ROS-induced protein aggregation and malfunctioning of the proteostasis network [53]. Likewise, autophagy has been proposed as anti-ageing mechanism as it sustains neuronal survival promoting clearance of protein aggregates [54]. Additionally, ALP contributes to proteostasis maintenance driving the degradation of cytotoxic aggregates and preventing their cellular accumulation. Oxidative damage could be reduced by autophagic machinery by removal of unnecessary or damaged organelles, or by limiting the excessive ROS activation in response to neuronal damage [55]. Protein aggregates and ROS-induced dysfunctional mitochondria can propagate damage in surrounding cells, causing further protein aggregations, and spreading oxidative cellular stress, ultimately leading to neurodegenerative processes. Indeed, scavenging of free radicals by antioxidant systems cannot completely control peroxidation and protein aggregation, indicating that autophagy has an essential and effective role in neuronal survival and OS decline [56]. Recent studies revealed that OS can also impair autophagy via direct oxidation of catalytic thiols on several proteins involved in the autophagic cascade, such as autophagy-related (ATG) proteins ATG3 and ATG7 [57]. Therefore, protein accumulation significantly contributes to OS and, on its turn, exacerbates the age-related production and aggregation of misfolded proteins. This synergistic correlation affects both etiology and progression of NDDs, representing either a potential biomarker for diagnosis and a valuable target for the treatment of OS-related proteinopathies [58]. The most relevant physiological and pathological roles addressed to ROS are summarized in Table 1.

3. Nature-Based Compounds against Cellular Aging and Neurodegeneration

Historically, several nature-based compounds from dietary intake—including polyphenols, terpenes, and organosulfur compounds—have shown beneficial effects against OS, cellular ageing, and neurodegeneration mainly associated with their antioxidant activities [59]. Through the years, research in this field has focused on elucidating and exploring the pharmacological profile of natural products against OS and their therapeutic application for NDDs [60]. Beside antioxidant properties, phytonutrients such as quercetin, resveratrol, curcumin, and genistein can propel beneficial effects through activation of different pathways, including autophagy [61,62]. Due to their prominent polypharmacology, natural compounds have been widely employed in drug discovery and their chemical scaffolds frequently used as starting points in the design of new therapeutic agents for several pathologies [63]. More recently, and with the rise of nutraceutics and functional food, a massive number of studies identified polyphenols as potential anti-ageing and neuroprotective molecules that could find application for preventing and/or treating chronic diseases [64,65]. In the next paragraphs, we will discuss promising and attractive properties of selected nature-based scaffolds for a possible application in drug discovery, especially for NDDs, highlighting their multifaceted effects against OS and neuronal disorders.

3.1. Polyamines: Spermidine and Spermine

Historically, polyamines were first discovered by van Leeuwenhoek in 1678 as crystalline substances in seminal plasma, but their structures were characterized only in the 20th century by Rosenheim [66]. Different concentrations of polyamines have been found in fruits, vegetables, and food of animal origin [67]. Spermidine, spermine, and their common diamine precursor putrescine are natural amines distributed widely in all cells, and their biosynthesis, degradation, and membrane transport are finely tuned in mammalians. The first step in polyamines biosynthesis requires the conversion of arginine to L-ornithine catalysed by the ureo-hydrolytic enzyme arginase. Putrescine is then obtained from ornithine through decarboxylation catalyzed by ornithine decarboxylase enzyme (ODC). Finally, spermidine and spermine synthases catalyse the biosynthesis of spermidine and spermine, respectively, adding an aminopropyl group to putrescine or to spermidine (Figure 1) [68].
Polyamines exert a wide range of beneficial effects against cellular aging and OS. Endogenous levels of spermidine decrease with age but remain stable in those older than 90 years of age, suggesting that it may contribute in longevity and anti-aging protection [69]. Spermine and spermidine are well-known for their activity as direct ROS scavengers providing DNA protection from OS [70,71]. Both in vitro and in vivo studies demonstrate that the antioxidant and anti-inflammatory properties of spermidine are associated with decremental cellular levels of nitric oxide synthase (NOS), prostaglandins and cytokines.: such as NOS inducible isoform (iNOS), prostaglandine E2 (PGE2), and pro-inflammatory cytokines, like interleukin-1β (IL-1β) and tumor necrosis factor-α (TNF-α). In addition, spermidine inactivates NF-κB in macrophages influencing pro-inflammatory genes expression [72]. Autophagy has been identified as one of the main mechanisms responsible for anti-ageing effects of spermidine, even though more pathways are affected by this molecule such as cell proliferation and differentiation, lipid metabolism and inflammation [73]. The polycationic nature of polyamines drives their capacity to interact with negatively charged molecules, such as DNA, RNA, proteins, and lipids. For this reason, they are involved in various processes, including DNA stability/repairment and proteins and nucleic acids synthesis [74,75]. Spermidine is also correlated with both hyper- and hypo-acetylation of proteins involved in the autophagic process [76,77]. Spermidine competitively inhibits the acetyltransferase EP300, an autophagy inhibitor that directly acetylates and blocks several ATG complexes and microtubule-associated protein 1A/1B-light chain 3 (LC3), providing a stimulation of autophagic flux in mammalian cells [78]. Autophagy promotion induced by spermidine is also mediated by its inhibitory effect on histone acetyltransferase enzyme P/CAF leading to histone H3 hypoacetylation. Histone acetylation and DNA interaction seem to be involved in the synergistic mechanism by which spermidine, and likely other polyamines, promote autophagy and influence chromatin structure [79]. Another study showed that spermidine also prolongs lifespan and reduces OS in mice models of hepatocarcinoma. These effects are due to the increased acetylation of microtubule-associated protein 1S gene (MAP1S) and consequently autophagy promotion [80]. Moreover, spermine showed a good potential towards clearance of misfolded proteins in prion infected cell cultures [81]. Additionally, it increases acetylation of microtubules, thus enhancing the retrograde transport of autophagic vesicles to lysosomes. Spermine offers also an epigenetic control associated with DNA-methylation: it seems to exert anti-inflammatory protection via DNA methyltransferase (DNMT) activation that regulates methylation of the entire genome as well hypermethylation of Integrin Alpha L (ITGAL) gene. Hypermethylation of ITGAL is correlated with the suppression of inflammatory status [82]. In vivo studies revealed that spermidine exerts a neuroprotective effect in brain injured mice showing reduction of pro-inflammatory cytokines and traumatic brain injury biomarkers. This neuroprotection is given also by a spermidine-driven autophagic activation. In brain samples, after spermidine treatment, Beclin-1 and LC3 markers are upregulated, thus confirming an increase of autophagic activity also in vivo [83]. Finally, polyamines proved to ameliorate cognitive functions. Intrahippocampal co-administration of spermidine and arcaine (an antagonist of polyamine binding site at NMDA receptor) in mice exerts a modulatory effect on memory, presumably by activating NMDA receptors [84].
Despite this plethora of effects and their chemical appeal, polyamines are quietly unexplored as therapeutic tools in designing new multi-target drug candidates for NDDs treatments. One of the first polyamine-based papers on this topic dealt with merging polyamines with the 1-aminoindan scaffold, in order to improve the lipophilicity of the new molecules and then promote the transport at the CNS through biological membranes. Among the new compounds synthesised by Gilad et al., compound 2711 (1, Table 2) compared with natural polyamines, showed to enhance the neuroprotective effect on damaged brain tissue, proving either a potent neuroprotective effect in in vitro/in vivo experimental models of neurotrauma and a good capability to reach the CNS [85]. Accordingly, in 2010 Melchiorre and collaborators reported memoquin as a new compound synthesized following the multi-target ligand design (MTLD) approach [86]. Memoquin (2, Table 2) was obtained by the combination of the 1,4-benzoquinone scaffold (a radical scavenger) with the polyamine structure of caproctamine, an acetylcholinesterase (AChE) inhibitor and muscarinic M2 receptor antagonist. Both in vitro and in vivo assays demonstrated that treatment with memoquin can affect a wide range of pathogenic mechanisms involved in AD, including Aβ aggregation, tau hyperphosphorylation, OS, and AChE and BACE-1 activities [87]. The alkyl-2-methoxy-benzyl fragment of memoquin, which seems to play a key role in AChE inhibitory activity, was also combined with ferulic acid (FA) using polyamine linkers (3, Table 2). New molecules originated with this approach showed a pleiotropic activity with an improved antioxidant profile [88]. Similarly, Zhang and colleagues merged antioxidant isoflavonoid scaffold of genistein with various polyamines, to identify new multifunctional anti-AD agents (4, Table 2). New designed agents were able to inhibit AChE and butyrylcholinesterases (BuChE), and showed chelating properties towards Fe3+, Cu2+, and Zn2+, without inducing cytotoxicity in vitro [89]. Again, Simoni et al. conjugated 3,5-dibenzylidenepiperidin-4-one bioactive motives with spermine to target amyloid aggregation as a promising strategy in AD treatment (5, Table 2). Dicatecholic derivatives showed good inhibition of Aβ42 aggregation, even though they did not show any antioxidant properties in the same in vitro model [90]. Furthermore, in silico studies revealed an additional key functional role of spermine suggesting that it is directly involved in interactions between Aβ42 monomers. The discovery of dicaffeoylspermidine derivatives with antioxidant properties from wolfberry prompted Gao et al. to design dicaffeoylspermidine cyclized derivatives (6, Table 2). These compounds showed significant antioxidant activity in vitro along with an improvement of memory and cognitive functions in fruit flies’ model of senile dementia [91,92]. Taken together, these studies suggest a new and unstudied pharmaceutical potential of polyamines in drug discovery that could be exploited to design new pharmacological agents against NDDs.

3.2. Phenolic Acids

For decades, long-term consumption of polyphenols-rich food has been correlated with beneficial effects in human health and protection against cancers, cardiovascular, and neurodegenerative diseases [93]. Polyphenols are a large class of natural compounds characterized by a chemical scaffold with multiple phenolic functionalities and endowed of several biological activities [94,95]. Among these, phenolic acids, have attracted a growing interest in the pharmaceutical field for their strong antioxidant nature and low toxicity [96,97,98]. Their antioxidant activity is mainly provided by the high reactivity of the phenolic moiety able to interact with free radicals which are stabilized by delocalization. This effect induced by polyphenols results in a marked modification of radical-mediated oxidation processes [99].
Moreover, additional effects have been described in the last years for phenolic acids, like ferulic acid (FA), gallic acid (GA), and caffeic acid (CA) (Figure 2). For example, FA has shown marked neuroprotective effects in a mouse model of cerebral ischemia/reperfusion-induced injury, associated with an increase in SOD and GPH levels, and a consequent reduction of ROS, O2•−, and Ca2+ accumulation. Besides the antioxidant effect, the neuroprotective effect has been also correlated with the downregulation of toll-like receptor 4 (TRL4) and myeloid differentiation primary response 88 (MyD88), both involved in the activation of intracellular NF-κB signalling pathway [100]. Similarly, GA showed a protective role in OS-induced dopaminergic cell lines through two different mechanisms: activation of antioxidant enzymes (i.e., SOD, CAT, GPx, and GR) and regulation of AKT/Keap-1/Nrf2 defence pathway [101]. Polypharmacological profile of phenolic acids is also supported by several studies highlighting their metal chelating properties [102,103], and their ability to positively affect protein aggregation by directly interacting with proteins like Aβ and α-synuclein [104,105]. Additionally, phenolic acids may provide a protective effect regulating autophagy activation. Recent in vitro studies showed that CA is capable of restoring heat shock protein 27 (Hsp27), B-cell lymphoma 2 gene (Bcl-2), and sirtuin 1 (SIRT1) expression, upregulating autophagy, and decreasing mitochondrial ROS by enhancing expression of antioxidant proteins such as GSH, catalase, O-1, NQO-1, and SOD [106]. FA showed protection against vascular dementia by reversing OS state, upregulating LC3-II, and inducing mitophagy in vitro [107,108]. Generally, collected data on phenolic acids underline their cell-protective polypharmacology, suggesting phenolic acids as therapeutic tools for several disorders.
In order to identify new classes of drugs with enhanced neuroprotective effect and following the well-known multi-target designed ligands (MTDL) approach, a growing number of studies report the combination of phenolic acid scaffold with additional pharmacophoric moieties. Particularly, tacrine and its analogues have been widely employed to this end. Tacrine is a well-known cholinesterase inhibitor, the first AChEI approved by FDA for AD. Besides its primary activity, it also induces OS, due to its capability to promote ROS production and glutathione depletion [109]. A tacrine–ferulic acid hybrid namely T6FA (7, Table 3) was evaluated on Aβ-induced cell death in vitro and in in vivo mice model of AD. Results showed that T6FA enhances cognitive impairment, increasing SOD activity, and limiting AChE activity [110]. In 2018, Zhu et al. designed novel tacrine-ferulic acid hybrids protecting the FA’s free phenolic group with different benzyl moieties. Among these, compound bearing the 3,4-dimethyl benzyl scaffold (8, Table 3) showed an interesting pharmacological profile, inhibiting both AChE and BuChE with an EC50 of 37.02 and 101.4 nM, respectively. Moreover, it was able to inhibit Aβ1–42 self-induced aggregation in vitro and to improve cognitive impairment in mice model of AD [111]. Similarly, other AChE inhibitors (e.g., the marketed drugs donepezil and rivastigmine), neurotransmitters, or natural products have been combined with FA, CA, and cinnamic acid with the aim of developing new therapeutic tools against AD. Table 3 summarizes some examples of this approach, including effects described within the corresponding references: serotonin (9), aromatic amides and esters (10), donepezil and N,N-dibenzyl(N-methyl)amine (11 and 12), anilides (13), rivastigmine (14), and diallyl sulfide (15) [112,113,114,115,116,117]. Interestingly, most of the compounds synthesised showed a reduction of Aβ-aggregation, mainly imputable to a direct inhibition of the aggregative process itself or to an indirect effect linked to AChE-inhibition. Coherently, novel ferulic derivatives obtained from combination previously synthetized multitarget ligands pharmacophoric groups showed promising results (16, Table 3) [118]. Particularly, TM10 exhibited excellent BuChE inhibitory activity (IC50 = 8.9 nM) and high selectivity compared with AChE (IC50 = 12.1 μM), along with good antioxidant activity. Moreover, Sang et al. showed that TM10 inhibits Aβ1-42 aggregation and promotes disaggregation of Aβ1-42 fibrils and induces autophagy [118]. All together the positive results obtained following the MTDL approach confirm that the combination of AChEI with phenolic acids could exert neuroprotective effects by interacting with several pathways such as OS, protein-aggregation and autophagy, all aspects which results to be compromised in several NDDs. In particular, FA has been widely employed in multitarget ligand design because of its polypharmacology and low toxicity even if it still presents some drawbacks like low bioavailability and poor BBB permeability. In 2020, Tripathi et al. synthesized a new series of compounds obtained by the combination of FA with 1,3,4-oxadiazole ring in order to improve the interaction of new synthesised molecules with the active pocket of target enzymes (17, Table 3). All compounds proved a remarkable inhibition of AChE, BuChE, and BACE-1. Moreover, selected compounds were able to reduce Aβ aggregation, exert neuroprotective effects on Aβ-induced damage SH-SY5Y cells and exhibit appreciable BBB permeability [119].

3.3. Urolithins

Ellagitannins (ET) represent another class of polyphenols, mainly found in strawberries, walnuts, and pomegranates, with prominent beneficial effects on human health [120]. After dietary intake, ET are slightly absorbed and metabolized by the gut microbiota of humans and other animals [121,122]. First, they are converted into ellagic acid (EA) that retains poor bioavailability, then metabolism in the lower gastrointestinal tract convert them in urolithins. Urolithin A (UA), urolithin B (UB), urolithin C (UC), isourolithin A (iso-UA), and their corresponding phase II conjugate derivatives represent the main metabolites found in tissues and plasma [123]. (Figure 3)
EA pharmacological activities and molecular mechanism are still under investigation, but data collected until now are particularly promising. Interestingly, most of beneficial effects associated with EA are attributed to its active microbial metabolites: urolithins are much better absorbed and have been proposed to be responsible of beneficial effects of ET-rich foods [124]. Coherently, over the last decade, researchers’ interests shifted towards urolithins in order to understand their physiological effects and to explore the mechanisms involved. Structurally, urolithins are dibenzopyran-6-one compounds with various hydroxyl groups substitutions, and UA, the 3,8-dihydroxy derivative, appears to be the most abundant metabolite produced in human. UA has shown non-genotoxic and ADME safety profile in short- and long-term oral exposure in rats [125]. Additionally, it demonstrated anti-mutagenic properties in Caenorhabditis elegans (C. elegans) [126]. Pharmacodynamic studies on pomegranate juice consumption demonstrated that UA can reach micromolar concentrations in humans, without displaying any toxic effects [127]. A prominent interest on neuroprotective effects of urolithins arose upon observation of their scavenger activity against ROS. Recent studies showed that they inhibit intracellular ROS production in vitro, without remarkable cytotoxic effects; interestingly, this effect has been correlated with the number of hydroxyl groups [128]. UA showed neuroprotective effects against H2O2, reducing ROS production, improving mitochondrial activity and reducing ROS-induced lipid peroxidation in murine neuroblastoma cell lines (neuro-2a). Apart from the direct radical scavenging properties, UA seems to exert its antioxidant activity also through the modulation of antioxidant enzymes. UA increased the expression of peroxiredoxins, a family of thiol-dependent peroxidases involved in redox signaling. This correlation may explain the cytoprotection of UA by improving the activity of other antioxidant defence systems, such as SOD, catalase, and glutathione reductase [129]. Moreover, urolithins showed metal-chelating properties linked to the number and position of hydroxyl groups [102,130]. In the last years, urolithins have also earned interest as potential autophagy modulator. Indeed, several studies indicated that UA can promote autophagy in macrophages [131], colorectal cancer cells [132], and microglia human cell lines [133]. Again, UA showed a neuroprotective effect through autophagic activation, repressing ER stress and attenuating neuronal injuries in mice [134]. By contrast, UB cytoprotective effects has been associated with the modulation of the uncanonical p62/Keap1/Nrf2 pathway, resulting in increased levels of downstream antioxidant enzymes [135]. Despite the wide range of effects against OS and the promising results obtained from studies on biotransformation, metabolism, and physiological effects, the use of urolithins as chemical scaffolds for drug design in the field of NDDs is still relatively unexplored. In 2014, Gulcan et al., combined modified 6H-benzo[c]chromen-6-one core of urolithin with rivastigmine (18, Table 4) and donepezil-like scaffolds (19, Table 4), obtaining a small library of derivatives with promising polypharmacological activity. Selected compounds showed micromolar and sub-micromolar IC50 against AChE and BuChE and exhibited comparable activity with donepezil and rivastigmine, in a scopolamine induce passive avoidance test [136]. A few years later, the same research group published new urolithin-based MTDL compounds, obtained from combination of urolithin or tetrahydrourolithin with donepezil-like scaffolds throughout a propylene linker (20, Table 4). Even if these compounds exhibited good anticholinergic activity, they lack the structural requirements to prevent amyloid beta aggregation inhibition, suggesting that the design of new AChEI should not be sufficient to prevent cholinesterase induced Aβ aggregation [137].
Recently, the modulation of the casein kinase system (CK) has emerged as new therapeutic approach for NDDs. In particular, CK2 is a ubiquitous protein kinase that seems to play important roles in neural functions including synaptic transmission, and synaptic plasticity, suggesting a potential critical role also in the progression of AD [138]. Cozza et al. proposed urolithins scaffold as a starting point for designing new and promising CK2 inhibitors. New urolithin derivatives (21, Table 4) showed highly selective sub-micromolar activity against CK2, suggesting that they could led the basis of new approaches against NDDs [139,140]. In 2016, Xie et al. proposed new donepezil-coumarin multi-target hybrids, identifying tetrahydrourolithin derivatives as potential therapeutic tools for AD (22, Table 4). Indeed, they showed good in vitro inhibition of hAChE and hBuChE (IC50 = 1.37 μM and 1.98 μM, respectively), and MAO-B (IC50 = 2.62 μM), and no inhibition towards MAO-A. Moreover, tetrahydrourolithin derivatives showed an interesting in vitro capacity of permeating the BBB, and no cytotoxicity in SH-5YSY and HepG2 cells at concentrations up to 50 μM [141].

3.4. Lipoic Acid

α-Lipoic acid (LA) is a natural disulfide antioxidant compound occurring in vegetables, meat, and fruits. In cells LA is converted to dihydrolipoic acid (DHLA) and plays a prominent role as co-factor for mitochondrial dehydrogenases—i.e., pyruvate dehydrogenase (PDH) and α-ketoglutarate dehydrogenase (KGDH) [142]. Since the 1950s, antioxidant properties of LA have been investigated, identifying in this compound a beneficial micronutrient that could find potential therapeutic application against OS and NDDs [143] (Figure 4).
In vitro evidence observed a direct radical scavenging capacity of LA against OS [144]. However, recent in vivo studies have suggested that LA and DHLA effects are mainly due to the improvement of antioxidant enzymes levels—such as CAT, SOD, GPx and glucose-6-phosphate dehydrogenase (G6PD) [145]. LA restored age-declined GSH levels in rats CNS, sustaining antioxidant defences and protecting against aging alterations [146]. Retention of GSH basal levels and promotion of antioxidant effects were also observed in subarachnoid hemorrhage rat model after LA administration [147]. Additionally, LA is capable of chelating transitional metals, taking part in prooxidant Fenton reactions, and is involved in protein aggregation. Studies employing in vitro model of iron-overload and dietary administration of LA in rats indicated that LA possess antioxidant and neuroprotective effects associated with its iron-chelating properties [148]. Subsequent studies reported good chelating properties both in vitro and in vivo models against additional metals including manganese, aluminium [149,150], and copper [151,152]. In 2019, Bjørklund et al. reviewed the protective chelating properties of LA and other thiol groups containing compounds, towards mercury, cadmium, and lead accumulation [153]. LA has also shown beneficial effects related to inhibition of lipid peroxidation: in patients affected by diabetic neuropathy LA showed a neuroprotective action reducing oxidative stress and lipid peroxidation [154]. In two separate experiments on rats, LA showed antioxidant activity, BBB protection, and a remarkable decrease of lipid peroxidation [147,155].
Since 2000, LA effects on cholinergic system have been deeply studied [156]. LA ameliorated cholinergic deficiency in vascular dementia rats model: levels of ACh and choline acetyltransferase (ChAT) were partially restored along with a decreased AChE activity. Chronic dietary intake of LA in aged Tg2576 mice reduced cognitive deficits induced by amyloid precursors overexpression, recovering spatial learning and memory retention without affecting brain Aβ levels [157]. Due to the multiple activities and the good pharmacokinetics, LA is currently the most studied nature-based compound in clinical trials and suggested as a new treatment option for NDDs including AD, ALS, and multiple sclerosis [158,159]. Based on these promising results, LA has been widely investigated as lead-structure for designing MTDL compounds. One of the first LA-based anti-AD multi-target compound was proposed by Rosini et al. as a combination of LA with the tacrine scaffold (23, Table 5). Lipocrine showed in vitro inhibition of AChE, BuChE, AChE-induced Aβ aggregation and cells protection against ROS [160]. Through the years, the same research group designed and evaluated other LA-multi-target analogues containing substituted tacrine, memoquin, and rivastigmine moieties with promising potential in OS and protein aggregation treatments [161]. Similarly, Table 5 summarized some examples of LA-multitarget ligands designed as neuroprotective and antioxidant molecules for the treatment of AD, thus proving its promising properties for the treatment and/or prevention of neurodegenerative disorders. For each LA-derivative, main pharmacological effects and pharmacophoric portions are indicated: dopamine (24), AChE inhibitors (25), hydrophobic portion of SR3677, a potent Rho-associated kinase 2 (ROCK2) and Aβ inhibitor (26), isosorbide (27), coumarin (28), phenolic acids (29), benzodiazepine (30), melatonin (31), niacin (32), and 3-n-butylphthalide (NBP), an anti-ischemic drug (33) [162,163,164,165,166,167,168,169,170,171].

4. Discussion

In the last 50 years, interest in bioactive natural products has increased considerably. Indeed, extensive studies have been conducted on well-known substances (e.g., polyphenols and phenolic acids), identifying a large number of previously unknown therapeutical effects. Additionally, the impressive advancements in extractive techniques and mass spectrometry allowed the isolation and identification of several previously undetectable active compounds (e.g., urolithins). In parallel, our knowledge on complex uncurable pathologies has widely progressed towards the identification of several mechanisms (e.g., autophagy, proteostasis) involved in their pathogenesis, leading to a paradigm-shift in our conception of several diseases. Beside neoplastic diseases, NDDs have probably received most of the attention of pharmaceutical researchers and drug discovery programs. The identification of a vast tangled network of pathogenic mechanisms responsible of the onset and progression of NDDs has forced the academic world and pharmaceutical industries to re-think the therapeutic approaches to pursue. As discussed above, excessive OS correlates well with several dysfunctionalities observed in these pathologies resulting an intriguing target for new pharmacological agents. However, the development of new drugs able to promote a rebalance of OS still represents a difficult goal to achieve, especially since physiological levels of OS mediators (e.g., ROS) are required for maintaining healthy conditions.
In this context, polypharmacology associated with natural products have gained a prominent role in the drug discovery process for NDDs; intriguingly, the first drug approved by the Chinese regulatory authority in 2019 for AD is a gut-microbiota reprograming agent identified in seaweeds [172]. However, most NDDs (e.g., AD, PD, ALS) still lack a proper efficacious therapy able to halt, reduce or reverse their pathological progression. Interestingly, last years have seen an exceptional increase in clinical trials employing dietary enrichment with natural products (e.g., spermidine) or administration of natural extracts in NDDs patients [73,173]. Particularly, one of the most successful and promising cases is represented by ECGG that is currently in phase 2 as dietary supplement for AD and phase 3 for ALS (source: www.alzforum.org) (accessed on 11 February 2021). However, most of bioactive natural small molecules lack fundamental features required for being successfully delivered as NDDs drugs. Low aqueous solubility, poor PK and PD profiles, as well as high metabolic in vivo conversion, strongly hinder the progress of most natural products through the drug discovery chain. Accordingly, the MTDL approach could help to perfectly merge polypharmacology and natural products benefits, generating new chemical entities with the pharmacological fingerprints of natural products and able to simultaneously target two or more pathological mediators.

5. Conclusions

In the last years, several attempts have been made towards the development of new naturally derived polypharmacological agents and many new multitarget drugs with improved pharmacological features and optimized PK/PD profile have been synthesized. In this review, we summarized a large number of nature-based multitarget drugs developed in the last 10 years. Particularly, we focused on polyphenols and polyamines that share a prominent antioxidant effect. This feature could play a key role in developing new therapies for multifactorial diseases like NDDs, but not only. Indeed, we widely discussed how their therapeutical potential goes beyond the known antioxidant effects, playing a key role in defending the organism from excessive OS and having demonstrated a plethora of beneficial effects, including decreasing Aβ aggregation, promoting autophagy, and reducing metal accumulation. However, few issues have yet to be addressed. To the best of our knowledge, none of the nature-based multitarget molecules presented in this review has progressed through the drug discovery chain to successfully reach the market. The promising activity showed in vitro and even in preclinical studies still required to be validated in human subjects affected by NDDs. Nevertheless, the overview provided by the polypharmacology associated with natural compounds—and their synthetic derivatives—suggests that nature-based templates could represent a valuable tool for the drug design in the effort to realize new valuable drugs for NDDs.

Author Contributions

Conceptualization, A.B., M.R. and S.R.; writing-original draft preparation, A.B. and M.R.; writing—review and editing A.B., M.R., S.S. and S.R., supervision S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, I.P. Age-related neurodegenerative disease research needs aging models. Front. Aging Neurosci. 2015, 7, 168. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, Z.; Zhou, T.; Ziegler, A.C.; Dimitrion, P.; Zuo, L. Oxidative Stress in Neurodegenerative Diseases: From Molecular Mechanisms to Clinical Applications. Oxidative Med. Cell. Longev. 2017, 2017, 1–11. [Google Scholar] [CrossRef] [PubMed]
  3. Sidorova, Y.; Domanskyi, A. Detecting Oxidative Stress Biomarkers in Neurodegenerative Disease Models and Patients. Methods Protoc. 2020, 3, 66. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, X.; Guo, C.; Kong, J. Oxidative stress in neurodegenerative diseases. Neural Regen. Res. 2012, 7, 376–385. [Google Scholar]
  5. Nesi, G.; Sestito, S.; Digiacomo, M.; Rapposelli, S. Oxidative Stress, Mitochondrial Abnormalities and Proteins Deposition: Multitarget Approaches in Alzheimer’s Disease. Curr. Top. Med. Chem. 2017, 17, 3062–3079. [Google Scholar] [PubMed]
  6. Sultana, R.; Perluigi, M.; Butterfield, D.A. Lipid peroxidation triggers neurodegeneration: A redox proteomics view into the Alzheimer disease brain. Free Radic. Biol. Med. 2013, 62, 157–169. [Google Scholar] [CrossRef] [Green Version]
  7. Fujikake, N.; Shin, M.; Shimizu, S. Association Between Autophagy and Neurodegenerative Diseases. Front. Neurosci. 2018, 12, 255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Krisko, A.; Radman, M. Protein damage, ageing and age-related diseases. Open Biol. 2019, 9, 180249. [Google Scholar] [CrossRef] [Green Version]
  9. Kim, G.H.; Kim, J.E.; Rhie, S.J.; Yoon, S. The Role of Oxidative Stress in Neurodegenerative Diseases. Exp. Neurobiol. 2015, 24, 325–340. [Google Scholar] [CrossRef]
  10. Schiavone, S.; Trabace, L. Small Molecules: Therapeutic Application in Neuropsychiatric and Neurodegenerative Disorders. Molecules 2018, 23, 411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Flora, S.J.S. Structural, Chemical and Biological Aspects of Antioxidants for Strategies Against Metal and Metalloid Exposure. Oxidative Med. Cell. Longev. 2009, 2, 191–206. [Google Scholar] [CrossRef]
  12. Zeng, Q.; Siu, W.; Li, L.; Jin, Y.; Liang, S.; Cao, M.; Ma, M.; Wu, Z. Autophagy in Alzheimer’s disease and promising modulatory effects of herbal medicine. Exp. Gerontol. 2019, 119, 100–110. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, S.-R.; Fu, Y.-S.; Tsai, M.-J.; Cheng, H.; Weng, C.-F. Natural Compounds from Herbs that can Potentially Execute as Autophagy Inducers for Cancer Therapy. Int. J. Mol. Sci. 2017, 18, 1412. [Google Scholar] [CrossRef] [Green Version]
  14. Chung, Y.J.; Robert, C.; Gough, S.M.; Rassool, F.V.; Aplan, P.D. Oxidative stress leads to increased mutation frequency in a murine model of myelodysplastic syndrome. Leuk. Res. 2014, 38, 95–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Luceri, C.; Bigagli, E.; Femia, A.P.; Caderni, G.; Giovannelli, L.; Lodovici, M. Aging related changes in circulating reactive oxygen species (ROS) and protein carbonyls are indicative of liver oxidative injury. Toxicol. Rep. 2018, 5, 141–145. [Google Scholar] [CrossRef] [PubMed]
  16. Glass, C.K.; Saijo, K.; Winner, B.; Marchetto, M.C.; Gage, F.H. Mechanisms Underlying Inflammation in Neurodegeneration. Cell 2010, 140, 918–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Aon, M.A.; Stanley, B.A.; Sivakumaran, V.; Kembro, J.M.; O’Rourke, B.; Paolocci, N.; Cortassa, S. Glutathione/thioredoxin systems modulate mitochondrial H2O2 emission: An experimental-computational study. J. Gen. Physiol. 2012, 139, 479–491. [Google Scholar] [CrossRef] [Green Version]
  18. Bae, Y.S.; Oh, H.; Rhee, S.G.; Yoo, Y.D. Regulation of reactive oxygen species generation in cell signaling. Mol. Cells 2011, 32, 491–509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Redza-Dutordoir, M.; Averill-Bates, D.A. Activation of apoptosis signalling pathways by reactive oxygen species. Biochim. Biophys. Acta Bioenerg. 2016, 1863, 2977–2992. [Google Scholar] [CrossRef]
  20. Chandel, N.S. Mitochondria as signaling organelles. BMC Biol. 2014, 12, 34. [Google Scholar] [CrossRef]
  21. Dhillon, V.S.; Fenech, M. Mutations that affect mitochondrial functions and their association with neurodegenerative diseases. Mutat. Res. Rev. Mutat. Res. 2014, 759, 1–13. [Google Scholar] [CrossRef]
  22. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial Reactive Oxygen Species (ROS) and ROS-Induced ROS Release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [Green Version]
  23. Maremonti, E.; Eide, D.M.; Rossbach, L.M.; Lind, O.C.; Salbu, B.; Brede, D.A. In vivo assessment of reactive oxygen species production and oxidative stress effects induced by chronic exposure to gamma radiation in Caenorhabditis elegans. Free Radic. Biol. Med. 2019, 152, 583–596. [Google Scholar] [CrossRef] [PubMed]
  24. Kauppila, J.H.; Stewart, J.B. Mitochondrial DNA: Radically free of free-radical driven mutations. Biochim. Biophys. Acta Bioenerg. 2015, 1847, 1354–1361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Panel, M.; Ghaleh, B.; Morin, D. Mitochondria and aging: A role for the mitochondrial transition pore? Aging Cell 2018, 17, e12793. [Google Scholar] [CrossRef] [PubMed]
  26. Aon, M.A.; Cortassa, S.; Akar, F.G.; O’Rourke, B. Mitochondrial criticality: A new concept at the turning point of life or death. Biochim. Biophys. Acta Mol. Basis Dis. 2006, 1762, 232–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Stewart, V.C.; Heales, S.J. Nitric oxide-induced mitochondrial dysfunction: Implications for neurodegeneration. Free. Radic. Biol. Med. 2003, 34, 287–303. [Google Scholar] [CrossRef]
  28. Halestrap, A.P.; Brenner, C. The adenine nucleotide translocase: A central component of the mitochondrial permeability transition pore and key player in cell death. Curr. Med. Chem. 2003, 10, 1507–1525. [Google Scholar] [CrossRef] [Green Version]
  29. Wei, Y.H.; Lu, C.Y.; Wei, C.Y.; Ma, Y.S.; Lee, H.C. Oxidative stress in human aging and mitochondrial disease-consequences of defective mitochondrial respiration and impaired antioxidant enzyme system. Chin. J. Physiol. 2001, 44, 1–11. [Google Scholar] [PubMed]
  30. Bertero, E.; Maack, C. Calcium Signaling and Reactive Oxygen Species in Mitochondria. Circ. Res. 2018, 122, 1460–1478. [Google Scholar] [CrossRef]
  31. Adam-Vizi, V.; Starkov, A.A. Calcium and Mitochondrial Reactive Oxygen Species Generation: How to Read the Facts. J. Alzheimer’s Dis. 2010, 20, S413–S426. [Google Scholar] [CrossRef] [Green Version]
  32. Tranah, G.J.; Nalls, M.A.; Katzman, S.M.; Yokoyama, J.S.; Lam, E.T.; Zhao, Y.; Mooney, S.; Thomas, F.; Newman, A.B.; Liu, Y.; et al. Mitochondrial DNA Sequence Variation Associated with Dementia and Cognitive Function in the Elderly. J. Alzheimer’s Dis. 2012, 32, 357–372. [Google Scholar] [CrossRef] [Green Version]
  33. Chaturvedi, R.K.; Beal, M.F. Mitochondria targeted therapeutic approaches in Parkinson’s and Huntington’s diseases. Mol. Cell. Neurosci. 2013, 55, 101–114. [Google Scholar] [CrossRef]
  34. Reddy, P.H. Mitochondrial oxidative damage in aging and Alzheimer’s disease: Implications for mitochondrially targeted antioxidant therapeutics. J. Biomed. Biotechnol. 2006, 2006, 31372. [Google Scholar] [CrossRef]
  35. Kawamata, H.; Manfredi, G. Mitochondrial dysfunction and intracellular calcium dysregulation in ALS. Mech. Ageing Dev. 2010, 131, 517–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Dexter, D.T.; Wells, F.R.; Lee, A.J.; Agid, F.; Agid, Y.; Jenner, P.; Marsden, C.D. Increased nigral iron content and alterations in other metal ions occurring in brain in Parkinson’s disease. J. Neurochem. 1989, 52, 1830–1836. [Google Scholar] [CrossRef]
  37. Valko, M.M.H.C.M.; Morris, H.; Cronin, M.T.D. Metals, Toxicity and Oxidative Stress. Curr. Med. Chem. 2005, 12, 1161–1208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Li, Y.; Jiao, Q.; Xu, H.; Du, X.; Shi, L.; Jia, F.; Jiang, H. Biometal Dyshomeostasis and Toxic Metal Accumulations in the Development of Alzheimer’s Disease. Front. Mol. Neurosci. 2017, 10, 339. [Google Scholar] [CrossRef] [PubMed]
  39. Barnham, K.J.; Bush, A.I. Metals in Alzheimer’s and Parkinson’s diseases. Curr. Opin. Chem. Biol. 2008, 12, 222–228. [Google Scholar] [CrossRef] [PubMed]
  40. Jomova, K.; Baros, S.; Valko, M. Redox active metal-induced oxidative stress in biological systems. Transit. Met. Chem. 2012, 37, 127–134. [Google Scholar] [CrossRef]
  41. Zatta, P.; Kiss, T.; Suwalsky, M.; Berthon, G. Aluminium (III) as a promoter of cellular oxidation. Coord. Chem. Rev. 2002, 228, 271–284. [Google Scholar] [CrossRef]
  42. Farina, M.; Avila, D.S.; da Rocha, J.B.T.; Aschner, M. Metals, oxidative stress and neurodegeneration: A focus on iron, manganese and mercury. Neurochem. Int. 2013, 62, 575–594. [Google Scholar] [CrossRef] [Green Version]
  43. Breydo, L.; Uversky, V.N. Role of metal ions in aggregation of intrinsically disordered proteins in neurodegenerative diseases. Metallomics 2011, 3, 1163–1180. [Google Scholar] [CrossRef] [PubMed]
  44. Aschner, M.; Chen, P.; Martinez-Finley, E.J.; Bornhorst, J.; Chakraborty, S. Metal-induced neurodegeneration in C. elegans. Front. Aging Neurosci. 2013, 5, 18. [Google Scholar]
  45. Leal, S.S.; Botelho, H.M.; Gomes, C.M. Metal ions as modulators of protein conformation and misfolding in neurodegeneration. Coord. Chem. Rev. 2012, 256, 2253–2270. [Google Scholar] [CrossRef]
  46. Urrutia, P.J.; Mena, N.P.; Nunez, M.T. The interplay between iron accumulation, mitochondrial dysfunction, and inflammation during the execution step of neurodegenerative disorders. Front. Pharmacol. 2014, 5, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Bolognin, S.; Drago, D.; Messori, L.; Zatta, P. Chelation therapy for neurodegenerative diseases. Med. Res. Rev. 2009, 29, 547–570. [Google Scholar] [CrossRef]
  48. Sales, T.A.; Prandi, I.G.; de Castro, A.A.; Leal, D.H.S.; da Cunha, E.F.F.; Kuca, K.; Ramalho, T.C. Recent Developments in Metal-Based Drugs and Chelating Agents for Neurodegenerative Diseases Treatments. Int. J. Mol. Sci. 2019, 20, 1829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Nedelsky, N.B.; Todd, P.K.; Taylor, J.P. Autophagy and the ubiquitin-proteasome system: Collaborators in neuroprotection. Biochim. Biophys. Acta Mol. Basis Dis. 2008, 1782, 691–699. [Google Scholar] [CrossRef] [Green Version]
  50. Valastyan, J.S.; Lindquist, S. Mechanisms of protein-folding diseases at a glance. Dis. Model. Mech. 2014, 7, 9–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Korovila, I.; Hugo, M.; Castro, J.P.; Weber, D.; Höhn, A.; Grune, T.; Jung, T. Proteostasis, oxidative stress and aging. Redox Biol. 2017, 13, 550–567. [Google Scholar] [CrossRef] [PubMed]
  52. Gregersen, N.; Bross, P. Protein Misfolding and Cellular Stress: An Overview. Adv. Struct. Saf. Stud. 2010, 648, 3–23. [Google Scholar] [CrossRef]
  53. Hipp, M.S.; Park, S.-H.; Hartl, F.U. Proteostasis impairment in protein-misfolding and-aggregation diseases. Trends Cell Biol. 2014, 24, 506–514. [Google Scholar] [CrossRef] [PubMed]
  54. Lim, J.; Yue, Z. Neuronal Aggregates: Formation, Clearance, and Spreading. Dev. Cell 2015, 32, 491–501. [Google Scholar] [CrossRef] [Green Version]
  55. Carloni, S.; Buonocore, G.; Balduini, W. Protective role of autophagy in neonatal hypoxia–ischemia induced brain injury. Neurobiol. Dis. 2008, 32, 329–339. [Google Scholar] [CrossRef]
  56. Giordano, S.; Darley-Usmar, V.; Zhang, J. Autophagy as an essential cellular antioxidant pathway in neurodegenerative disease. Redox Biol. 2014, 2, 82–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Burgoyne, J.R. Oxidative stress impairs autophagy through oxidation of ATG3 and ATG7. Autophagy 2018, 14, 1092–1093. [Google Scholar] [CrossRef] [PubMed]
  58. Kesidou, E.; Lagoudaki, R.; Touloumi, O.; Poulatsidou, K.-N.; Simeonidou, C. Autophagy and neurodegenerative disorders. Neural Regen. Res. 2013, 8, 2275–2283. [Google Scholar] [PubMed]
  59. Almeida, S.; Alves, M.G.; Sousa, M.; Oliveira, P.F.; Silva, B.M. Are Polyphenols Strong Dietary Agents Against Neurotoxicity and Neurodegeneration? Neurotox. Res. 2016, 30, 345–366. [Google Scholar] [CrossRef]
  60. Pohl, F.; Lin, P.K.T. The Potential Use of Plant Natural Products and Plant Extracts with Antioxidant Properties for the Prevention/Treatment of Neurodegenerative Diseases: In Vitro, In Vivo and Clinical Trials. Molecules 2018, 23, 3283. [Google Scholar] [CrossRef] [Green Version]
  61. Pallauf, K.; Rimbach, G. Autophagy, polyphenols and healthy ageing. Ageing Res. Rev. 2013, 12, 237–252. [Google Scholar] [CrossRef]
  62. Hasima, N.; Ozpolat, B. Regulation of autophagy by polyphenolic compounds as a potential therapeutic strategy for cancer. Cell Death Dis. 2014, 5, e1509. [Google Scholar] [CrossRef] [Green Version]
  63. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629–661. [Google Scholar] [CrossRef] [Green Version]
  64. Silva, R.F.M.; Pogačnik, L. Polyphenols from Food and Natural Products: Neuroprotection and Safety. Antioxidants 2020, 9, 61. [Google Scholar] [CrossRef] [Green Version]
  65. Koudoufio, M.; Desjardins, Y.; Feldman, F.; Spahis, S.; Delvin, E.; Levy, E. Insight into Polyphenol and Gut Microbiota Crosstalk: Are Their Metabolites the Key to Understand Protective Effects against Metabolic Disorders? Antioxidants 2020, 9, 982. [Google Scholar] [CrossRef] [PubMed]
  66. Bachrach, U. The early history of polyamine research. Plant Physiol. Biochem. 2010, 48, 490–495. [Google Scholar] [CrossRef] [PubMed]
  67. Muñoz-Esparza, N.C.; Latorre-Moratalla, M.L.; Comas-Basté, O.; Toro-Funes, N.; Veciana-Nogués, M.T.; Vidal-Carou, M.C. Polyamines in Food. Front. Nutr. 2019, 6, 108. [Google Scholar] [CrossRef]
  68. Wallace, H.M. The polyamines: Past, present and future. Essays Biochem. 2009, 46, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Pucciarelli, S.; Moreschini, B.; Micozzi, D.; de Fronzo, G.S.; Carpi, F.M.; Polzonetti, V.; Vincenzetti, S.; Mignini, F.; Napolioni, V. Spermidine and Spermine Are Enriched in Whole Blood of Nona/Centenarians. Rejuvenation Res. 2012, 15, 590–595. [Google Scholar] [CrossRef] [PubMed]
  70. Chattopadhyay, M.K.; Tabor, C.W.; Tabor, H. Polyamine deficiency leads to accumulation of reactive oxygen species in a spe2Delta mutant of Saccharomyces cerevisiae. Yeast 2006, 23, 751–761. [Google Scholar] [CrossRef]
  71. Sava, I.G.; Battaglia, V.; Rossi, C.A.; Salvi, M.; Toninello, A. Free radical scavenging action of the natural polyamine spermine in rat liver mitochondria. Free. Radic. Biol. Med. 2006, 41, 1272–1281. [Google Scholar] [CrossRef]
  72. Jeong, J.-W.; Cha, H.-J.; Han, M.H.; Hwang, S.J.; Lee, D.-S.; Yoo, J.S.; Choi, I.-W.; Kim, S.; Kim, H.-S.; Kim, G.-Y.; et al. Spermidine Protects against Oxidative Stress in Inflammation Models Using Macrophages and Zebrafish. Biomol. Ther. 2018, 26, 146–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ghosh, I.; Sankhe, R.; Mudgal, J.; Arora, D.; Nampoothiri, M. Spermidine, an autophagy inducer, as a therapeutic strategy in neurological disorders. Neuropeptides 2020, 83, 102083. [Google Scholar] [CrossRef] [PubMed]
  74. Molnar, M.M.; Liddell, S.C.; Wadkins, R.M. Effects of Polyamine Binding on the Stability of DNA i-Motif Structures. ACS Omega 2019, 4, 8967–8973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Iacomino, G.; Picariello, G.; d’Agostino, L. DNA and nuclear aggregates of polyamines. Biochim. Biophys. Acta Bioenerg. 2012, 1823, 1745–1755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Zhang, L.; Gong, H.; Sun, Q.; Zhao, R.; Jia, Y. Spermidine-Activated Satellite Cells Are Associated with Hypoacetylation in ACVR2B and Smad3 Binding to Myogenic Genes in Mice. J. Agric. Food Chem. 2018, 66, 540–550. [Google Scholar] [CrossRef] [PubMed]
  77. Eisenberg, T.; Knauer, H.; Schauer, A.; Büttner, S.; Ruckenstuhl, C.; Carmona-Gutierrez, D.; Ring, J.; Schroeder, S.; Magnes, C.; Antonacci, L.; et al. Induction of autophagy by spermidine promotes longevity. Nat. Cell Biol. 2009, 11, 1305–1314. [Google Scholar] [CrossRef] [PubMed]
  78. Pietrocola, F.; Lachkar, S.; Enot, D.P.; Nisosantano, M.; Pedro, J.M.B.-S.; Sica, V.; Izzo, V.; Maiuri, M.C.; Madeo, F.; Marino, G.; et al. Spermidine induces autophagy by inhibiting the acetyltransferase EP300. Cell Death Differ. 2014, 22, 509–516. [Google Scholar] [CrossRef]
  79. Burgio, G.; Corona, D.F.V.; Nicotra, C.M.A.; Carruba, G.; Taibi, G. P/CAF-mediated spermidine acetylation regulates histone acetyltransferase activity. J. Enzym. Inhib. Med. Chem. 2016, 31, 75–82. [Google Scholar] [CrossRef] [Green Version]
  80. Yue, F.; Li, W.; Zou, J.; Jiang, X.; Xu, G.; Huang, H.; Liu, L. Spermidine Prolongs Lifespan and Prevents Liver Fibrosis and Hepatocellular Carcinoma by Activating MAP1S-Mediated Autophagy. Cancer Res. 2017, 77, 2938–2951. [Google Scholar] [CrossRef] [Green Version]
  81. Phadwal, K.; Kurian, D.; Salamat, M.K.F.; Macrae, V.E.; Diack, A.B.; Manson, J.C. Spermine increases acetylation of tubulins and facilitates autophagic degradation of prion aggregates. Sci. Rep. 2018, 8, 1–17. [Google Scholar] [CrossRef] [Green Version]
  82. Soda, K. Spermine and gene methylation: A mechanism of lifespan extension induced by polyamine-rich diet. Amino Acids 2019, 52, 213–224. [Google Scholar] [CrossRef]
  83. Huang, J.; Zhang, H.; Zhang, J.; Yu, H.; Lin, Z.; Cai, Y. Spermidine Exhibits Protective Effects Against Traumatic Brain Injury. Cell. Mol. Neurobiol. 2020, 40, 927–937. [Google Scholar] [CrossRef] [PubMed]
  84. Signor, C.; Mello, C.F.; Porto, G.P.; Ribeiro, D.A.; Rubin, M.A. Spermidine improves fear memory persistence. Eur. J. Pharmacol. 2014, 730, 72–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gilad, G.M.; Gilad, V.H. Novel polyamine derivatives as neuroprotective agents. J. Pharmacol. Exp. Ther. 1999, 291, 39–43. [Google Scholar] [PubMed]
  86. Melchiorre, C.; Bolognesi, M.L.; Minarini, A.; Rosini, M.; Tumiatti, V. Polyamines in Drug Discovery: From the Universal Template Approach to the Multitarget-Directed Ligand Design Strategy. J. Med. Chem. 2010, 53, 5906–5914. [Google Scholar] [CrossRef] [PubMed]
  87. Bolognesi, M.L.; Cavalli, A.; Melchiorre, C. Memoquin: A multi-target-directed ligand as an innovative therapeutic opportunity for Alzheimer’s disease. Neurotherapeutics 2009, 6, 152–162. [Google Scholar] [CrossRef] [Green Version]
  88. Pan, W.; Hu, K.; Bai, P.; Yu, L.; Ma, Q.; Li, T.; Zhang, X.; Chen, C.; Peng, K.; Liu, W.; et al. Design, synthesis and evaluation of novel ferulic acid-memoquin hybrids as potential multifunctional agents for the treatment of Alzheimer’s disease. Bioorg. Med. Chem. Lett. 2016, 26, 2539–2543. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, X.; Wang, J.; Hong, C.; Luo, W.; Wang, C. Design, synthesis and evaluation of genistein-polyamine conjugates as multi-functional anti-Alzheimer agents. Acta Pharm. Sin. B 2015, 5, 67–73. [Google Scholar] [CrossRef] [Green Version]
  90. Simoni, E.; Caporaso, R.; Bergamini, C.; Fiori, J.; Fato, R.; Miszta, P.; Filipek, S.; Caraci, F.; Giuffrida, M.L.; Andrisano, V.; et al. Polyamine Conjugation as a Promising Strategy to Target Amyloid Aggregation in the Framework of Alzheimer’s Disease. ACS Med. Chem. Lett. 2016, 7, 1145–1150. [Google Scholar] [CrossRef] [Green Version]
  91. Zhou, Z.-Q.; Fan, H.-X.; He, R.-R.; Xiao, J.; Tsoi, B.; Lan, K.-H.; Kurihara, H.; So, K.-F.; Yao, X.-S.; Gao, H. Lycibarbarspermidines A–O, New Dicaffeoylspermidine Derivatives from Wolfberry, with Activities against Alzheimer’s Disease and Oxidation. J. Agric. Food Chem. 2016, 64, 2223–2237. [Google Scholar] [CrossRef]
  92. Gao, H.; Yao, X.; He, R.; Chen, G.; Zhou, Z.; Wang, C.; Hu, D.; Fan, H. Dicaffeoyl spermidine cyclized derivatives and use thereof. U.S. Patent No 10,457,702, 29 October 2019. [Google Scholar]
  93. Grosso, G. Effects of Polyphenol-Rich Foods on Human Health. Nutrients 2018, 10, 1089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. de Mello, A.J.M.; Fasolo, D. Chapter 20—Polyphenol Antioxidants from Natural Sources and Contribution to Health Promotion. In Polyphenols in Human Health and Disease; Watson, R.R., Preedy, R.V., Zibadi, S., Eds.; Academic Press: San Diego, CA, USA, 2014; pp. 253–265. [Google Scholar]
  95. Handique, J.; Baruah, J. Polyphenolic compounds: An overview. React. Funct. Polym. 2002, 52, 163–188. [Google Scholar] [CrossRef]
  96. Haque, M.E.; Javed, H.; Azimullah, S.; Khair, S.B.A.; Ojha, S. Neuroprotective potential of ferulic acid in the rotenone model of Parkinson’s disease. Drug Des. Dev. Ther. 2015, 9, 5499–5510. [Google Scholar] [CrossRef] [Green Version]
  97. Magnani, C.; Chiari, B.G.; Isaac, V.L.B.; Correa, M.A.; Salgado, H.R.N. In Vitro Safety Evaluation of Caffeic Acid. Athens J. Health 2014, 1, 181–188. [Google Scholar] [CrossRef]
  98. Zanwar, A.A.; Badole, S.L.; Shende, P.S.; Hegde, M.V.; Bodhankar, S.L. Role of Gallic Acid in Cardiovascular Disorders. In Polyphenols in Human Health and Disease; Elsevier BV: Amsterdam, The Netherlands, 2014; pp. 1045–1047. [Google Scholar]
  99. Shahidi, F.; Janitha, P.K.; Wanasundara, P.D. Phenolic antioxidants. Crit. Rev. Food Sci. Nutr. 1992, 32, 67–103. [Google Scholar] [CrossRef]
  100. Ren, Z.; Zhang, R.; Li, Y.; Li, Y.; Yang, Z.; Yang, H. Ferulic acid exerts neuroprotective effects against cerebral ischemia/reperfusion-induced injury via antioxidant and anti-apoptotic mechanisms in vitro and in vivo. Int. J. Mol. Med. 2017, 40, 1444–1456. [Google Scholar] [CrossRef] [Green Version]
  101. Chandrasekhar, Y.; Kumar, G.P.; Ramya, E.M.; Anilakumar, K.R. Gallic Acid Protects 6-OHDA Induced Neurotoxicity by Attenuating Oxidative Stress in Human Dopaminergic Cell Line. Neurochem. Res. 2018, 43, 1150–1160. [Google Scholar] [CrossRef] [PubMed]
  102. Chan, S.; Kantham, S.; Rao, V.M.; Palanivelu, M.K.; Pham, H.L.; Shaw, P.N.; McGeary, R.P.; Ross, B.P. Metal chelation, radical scavenging and inhibition of Aβ42 fibrillation by food constituents in relation to Alzheimer’s disease. Food Chem. 2016, 199, 185–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Andjelkovic, M.; Vancamp, J.; Demeulenaer, B.; Depaemelaere, G.; Socaciu, C.; Verloo, M.; Verhe, R. Iron-chelation properties of phenolic acids bearing catechol and galloyl groups. Food Chem. 2006, 98, 23–31. [Google Scholar] [CrossRef]
  104. Yu, M.; Chen, X.; Liu, J.; Ma, Q.; Zhuo, Z.; Chen, H.; Zhou, L.; Yang, S.; Zheng, L.; Hou, S.T.; et al. Gallic acid disruption of Abeta1-42 aggregation rescues cognitive decline of APP/PS1 double transgenic mouse. Neurobiol. Dis. 2019, 124, 67–80. [Google Scholar] [CrossRef] [PubMed]
  105. Ardah, M.T.; Paleologou, K.E.; Elv, G.; Khair, S.B.A.; Kazim, A.S.; Minhas, S.T.; Al-Tel, T.H.; Al-Hayani, A.A.; Haque, M.E.; Eeliezer, D.; et al. Structure activity relationship of phenolic acid inhibitors of α-synuclein fibril formation and toxicity. Front. Aging Neurosci. 2014, 6, 197. [Google Scholar] [CrossRef] [PubMed]
  106. Wu, Y.-L.; Chang, J.-C.; Lin, W.-Y.; Li, C.-C.; Hsieh, M.; Chen, H.-W.; Wang, T.-S.; Wu, W.-T.; Liu, C.-S.; Liu, K.-L. Caffeic acid and resveratrol ameliorate cellular damage in cell and Drosophila models of spinocerebellar ataxia type 3 through upregulation of Nrf2 pathway. Free. Radic. Biol. Med. 2018, 115, 309–317. [Google Scholar] [CrossRef]
  107. Askar, M.H.; Hussein, A.M.; Al-Basiony, S.F.; Meseha, R.K.; Metias, E.F.; Salama, M.M.; Antar, A.; El-Sayed, A. Effects of Exercise and Ferulic Acid on Alpha Synuclein and Neuroprotective Heat Shock Protein 70 in An Experimental Model of Parkinsonism Disease. CNS Neurol. Disord. Drug Targets 2019, 18, 156–169. [Google Scholar] [CrossRef]
  108. Chen, J.-L.; Duan, W.-J.; Luo, S.; Li, S.; Ma, X.-H.; Hou, B.-N.; Cheng, S.-Y.; Fang, S.-H.; Wang, Q.; Huang, S.-Q.; et al. Ferulic acid attenuates brain microvascular endothelial cells damage caused by oxygen-glucose deprivation via punctate-mitochondria-dependent mitophagy. Brain Res. 2017, 1666, 17–26. [Google Scholar] [CrossRef] [PubMed]
  109. Osseni, R.A.; Debbasch, C.; Christen, M.O.; Rat, P.; Warnet, J.M. Tacrine-induced Reactive Oxygen Species in a Human Liver Cell Line: The Role of Anethole Dithiolethione as a Scavenger. Toxicol. Vitr. 1999, 13, 683–688. [Google Scholar] [CrossRef]
  110. Pi, R.; Mao, X.; Chao, X.; Cheng, Z.; Liu, M.; Duan, X.; Ye, M.; Chen, X.; Mei, Z.; Han, Y.; et al. Tacrine-6-ferulic acid, a novel multifunctional dimer, inhibits amyloid-beta-mediated Alzheimer’s disease-associated pathogenesis in vitro and in vivo. PLoS ONE 2012, 7, e31921. [Google Scholar] [CrossRef] [Green Version]
  111. Zhu, J.; Yang, H.; Chen, Y.; Lin, H.; Li, Q.; Mo, J.; Bian, Y.; Pei, Y.; Sun, H. Synthesis, pharmacology and molecular docking on multifunctional tacrine-ferulic acid hybrids as cholinesterase inhibitors against Alzheimer’s disease. J. Enzym. Inhib. Med. Chem. 2018, 33, 496–506. [Google Scholar] [CrossRef] [Green Version]
  112. Takahashi, T.; Miyazawa, M. Serotonin derivatives as inhibitors of beta-secretase (BACE 1). Pharmacie 2011, 66, 301–305. [Google Scholar]
  113. Dhiman, P.; Malik, N.; Khatkar, A. Hybrid caffeic acid derivatives as monoamine oxidases inhibitors: Synthesis, radical scavenging activity, molecular docking studies and in silico ADMET analysis. Chem. Central J. 2018, 12, 1–17. [Google Scholar] [CrossRef]
  114. Estrada, M.; Herrera-Arozamena, C.; Pérez, C.; Viña, D.; Romero, A.; Morales-García, J.A.; Pérez-Castillo, A.; Rodríguez-Franco, M.I. New cinnamic—N-benzylpiperidine and cinnamic—N,N-dibenzyl(N-methyl)amine hybrids as Alzheimer-directed multitarget drugs with antioxidant, cholinergic, neuroprotective and neurogenic properties. Eur. J. Med. Chem. 2016, 121, 376–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Benchekroun, M.; Pachón-Angona, I.; Luzet, V.; Martin, H.; Oset-Gasque, M.J.; Marco-Contelles, J.; Ismaili, L. Synthesis, antioxidant and Aβ anti-aggregation properties of new ferulic, caffeic and lipoic acid derivatives obtained by the Ugi four-component reaction. Bioorg. Chem. 2019, 85, 221–228. [Google Scholar] [CrossRef]
  116. Nesi, G.; Chen, Q.; Sestito, S.; Digiacomo, M.; Yang, X.; Wang, S.; Pi, R.; Rapposelli, S. Nature-based molecules combined with rivastigmine: A symbiotic approach for the synthesis of new agents against Alzheimer’s disease. Eur. J. Med. Chem. 2017, 141, 232–239. [Google Scholar] [CrossRef] [PubMed]
  117. Simoni, E.; Serafini, M.M.; Bartolini, M.; Caporaso, R.; Pinto, A.; Necchi, D.; Fiori, J.; Andrisano, V.; Minarini, A.; Rosini, M.; et al. Nature-Inspired Multifunctional Ligands: Focusing on Amyloid-Based Molecular Mechanisms of Alzheimer’s Disease. ChemMedChem 2016, 11, 1309–1317. [Google Scholar] [CrossRef] [PubMed]
  118. Sang, Z.; Wang, K.; Han, X.; Cao, M.; Tan, Z.; Liu, W. Design, Synthesis, and Evaluation of Novel Ferulic Acid Derivatives as Multi-Target-Directed Ligands for the Treatment of Alzheimer’s Disease. ACS Chem. Neurosci. 2019, 10, 1008–1024. [Google Scholar] [CrossRef] [PubMed]
  119. Tripathi, A.; Choubey, P.K.; Sharma, P.; Seth, A.; Saraf, P.; Shrivastava, S.K. Design, synthesis, and biological evaluation of ferulic acid based 1,3,4-oxadiazole hybrids as multifunctional therapeutics for the treatment of Alzheimer’s disease. Bioorg. Chem. 2020, 95, 103506. [Google Scholar] [CrossRef]
  120. Landete, J. Ellagitannins, ellagic acid and their derived metabolites: A review about source, metabolism, functions and health. Food Res. Int. 2011, 44, 1150–1160. [Google Scholar] [CrossRef]
  121. Llorach, R.; Cerdá, B.; Cerón, J.J.; Espín, J.C.; Tomás-Barberán, F.A. Evaluation of the bioavailability and metabolism in the rat of punicalagin, an antioxidant polyphenol from pomegranate juice. Eur. J. Nutr. 2003, 42, 18–28. [Google Scholar] [CrossRef]
  122. Jurgoński, A.; Juśkiewicz, J.; Fotschki, B.; Kołodziejczyk, K.; Milala, J.; Kosmala, M.; Grzelak-Błaszczyk, K.; Markiewicz, L. Metabolism of strawberry mono- and dimeric ellagitannins in rats fed a diet containing fructo-oligosaccharides. Eur. J. Nutr. 2017, 56, 853–864. [Google Scholar] [CrossRef] [Green Version]
  123. Tomás-Barberán, F.A.; García-Villalba, R.; González-Sarrías, A.; Selma, M.V.; Espín, J.C. Ellagic Acid Metabolism by Human Gut Microbiota: Consistent Observation of Three Urolithin Phenotypes in Intervention Trials, Independent of Food Source, Age, and Health Status. J. Agric. Food Chem. 2014, 62, 6535–6538. [Google Scholar] [CrossRef]
  124. Cerdá, B.; Periago, P.M.; Espín, A.J.C.; Tomás-Barberán, F.A. Identification of Urolithin A as a Metabolite Produced by Human Colon Microflora from Ellagic Acid and Related Compounds. J. Agric. Food Chem. 2005, 53, 5571–5576. [Google Scholar] [CrossRef] [PubMed]
  125. Heilman, J.; Andreux, P.; Tran, N.; Rinsch, C.; Blanco-Bose, W. Safety assessment of Urolithin A, a metabolite produced by the human gut microbiota upon dietary intake of plant derived ellagitannins and ellagic acid. Food Chem. Toxicol. 2017, 108, 289–297. [Google Scholar] [CrossRef] [PubMed]
  126. Ryu, D.; Mouchiroud, L.; Andreux, P.A.; Katsyuba, E.; Moullan, N.; Nicolet-Dit-Félix, A.A.; Williams, E.G.; Jha, P.; Sasso, G.L.; 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]
  127. Seeram, N.P.; Zhang, Y.; McKeever, R.; Henning, S.M.; Lee, R.-P.; Suchard, M.A.; Li, Z.; Chen, S.; Thames, G.; Zerlin, A.; et al. Pomegranate Juice and Extracts Provide Similar Levels of Plasma and Urinary Ellagitannin Metabolites in Human Subjects. J. Med. Food 2008, 11, 390–394. [Google Scholar] [CrossRef] [PubMed]
  128. Bialonska, D.; Kasimsetty, S.G.; Khan, S.I.; Ferreira, D. Urolithins, Intestinal Microbial Metabolites of Pomegranate Ellagitannins, Exhibit Potent Antioxidant Activity in a Cell-Based Assay. J. Agric. Food Chem. 2009, 57, 10181–10186. [Google Scholar] [CrossRef]
  129. Cásedas, G.; Les, F.; Choya-Foces, C.; Hugo, M.; López, V. The Metabolite Urolithin-A Ameliorates Oxidative Stress in Neuro-2a Cells, Becoming a Potential Neuroprotective Agent. Antioxidants 2020, 9, 177. [Google Scholar] [CrossRef] [Green Version]
  130. Saha, P.; Yeoh, B.S.; Singh, R.; Chandrasekar, B.; Vemula, P.K.; Haribabu, B.; Vijay-Kumar, M.; Jala, V.R. Gut Microbiota Conversion of Dietary Ellagic Acid into Bioactive Phytoceutical Urolithin a Inhibits Heme Perox-idases. PLoS ONE 2016, 11, e0156811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Boakye, Y.D.; Groyer, L.; Heiss, E.H. An increased autophagic flux contributes to the anti-inflammatory potential of urolithin A in macrophages. Biochim. Biophys. Acta Gen. Subj. 2018, 1862, 61–70. [Google Scholar] [CrossRef]
  132. Zhao, W.; Shi, F.; Guo, Z.; Zhao, J.; Song, X.; Yang, H. Metabolite of ellagitannins, urolithin A induces autophagy and inhibits metastasis in human sw620 colorectal cancer cells. Mol. Carcinog. 2018, 57, 193–200. [Google Scholar] [CrossRef] [Green Version]
  133. Velagapudi, R.; Lepiarz, I.; El-Bakoush, A.; Katola, F.O.; Bhatia, H.; Fiebich, B.L.; Olajide, O.A. Induction of Autophagy and Activation of SIRT-1 Deacetylation Mechanisms Mediate Neuroprotection by the Pomegranate Metabolite Urolithin A in BV2 Microglia and Differentiated 3D Human Neural Progenitor Cells. Mol. Nutr. Food Res. 2019, 63, e1801237. [Google Scholar] [CrossRef] [Green Version]
  134. Ahsan, A.; Zheng, Y.; Wu, X.; Tang, W.; Liu, M.; Ma, S.; Jiang, L.; Hu, W.; Zhang, X.; Chen, Z. Urolithin A-activated autophagy but not mitophagy protects against ischemic neuronal injury by inhibiting ER stress in vitro and in vivo. CNS Neurosci. Ther. 2019, 25, 976–986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zheng, D.; Liu, Z.; Zhou, Y.; Hou, N.; Yan, W.; Qin, Y.; Ye, Q.; Cheng, X.-Y.; Xiao, Q.; Wu, X.; et al. Urolithin B, a gut microbiota metabolite, protects against myocardial ischemia/reperfusion injury via p62/Keap1/Nrf2 signaling pathway. Pharmacol. Res. 2020, 153, 104655. [Google Scholar] [CrossRef]
  136. Gulcan, H.O.; Unlu, S.; Esiringu, I.; Ercetin, T.; Sahin, Y.; Oz, D.; Sahin, M.F. Design, synthesis and biological evaluation of novel 6H-benzo[c]chromen-6-one, and 7,8,9,10-tetrahydro-benzo[c]chromen-6-one derivatives as potential cholinesterase inhibitors. Bioorg. Med. Chem. 2014, 22, 5141–5154. [Google Scholar] [CrossRef]
  137. Norouzbahari, M.; Burgaz, E.V.; Erçetin, T.; Fallah, A.; Foroumadi, A.; Firoozpour, L.; Sahin, M.F.; Gazi, M.; Gulcan, H.O. Design, Synthesis and Characterization of Novel Urolithin Derivatives as Cholinesterase Inhibitor Agents. Lett. Drug Des. Discov. 2018, 15, 1131–1140. [Google Scholar] [CrossRef]
  138. Blanquet, P. Casein kinase 2 as a potentially important enzyme in the nervous system. Prog. Neurobiol. 2000, 60, 211–246. [Google Scholar] [CrossRef]
  139. Perez, D.I.; Gil, C.; Martinez, A. Protein kinases CK1 and CK2 as new targets for neurodegenerative diseases. Med. Res. Rev. 2010, 31, 924–954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Cozza, G.; Gianoncelli, A.; Bonvini, P.; Zorzi, E.; Pasquale, R.; Rosolen, A.; Pinna, L.A.; Meggio, F.; Zagotto, G.; Moro, S. Urolithin as a converging scaffold linking ellagic acid and coumarin analogues: Design of potent protein kinase CK2 inhibitors. ChemMedChem 2011, 6, 2273–2286. [Google Scholar] [CrossRef]
  141. Xie, S.-S.; Lan, J.-S.; Wang, X.; Wang, Z.-M.; Jiang, N.; Li, F.; Wu, J.-J.; Wang, J.; Kong, L.-Y. Design, synthesis and biological evaluation of novel donepezil–coumarin hybrids as multi-target agents for the treatment of Alzheimer’s disease. Bioorg. Med. Chem. 2016, 24, 1528–1539. [Google Scholar] [CrossRef] [PubMed]
  142. Walgren, J.L.; Amani, Z.; McMillan, J.M.; Locher, M.; Buse, M.G. Effect of R(+)alpha-lipoic acid on pyruvate metabolism and fatty acid oxidation in rat hepatocytes. Metabolism 2004, 53, 165–173. [Google Scholar] [CrossRef] [PubMed]
  143. Packer, L.; Witt, E.H.; Tritschler, H.J. Alpha-lipoic acid as a biological antioxidant. Free. Radic. Biol. Med. 1995, 19, 227–250. [Google Scholar] [CrossRef]
  144. Li, Y.; Zhao, Y.; Yu, W.; Jiang, S. Scavenging ability on ROS of alpha-lipoic acid (ALA). Food Chem. 2004, 84, 563–567. [Google Scholar] [CrossRef]
  145. Moraes, T.B.; Zanin, F.; da Rosa, A.; de Oliveira, A.; Coelho, J.; Petrillo, F.; Wajner, M.; Dutra-Filho, C.S. Lipoic acid prevents oxidative stress in vitro and in vivo by an acute hyperphenylalaninemia chemically-induced in rat brain. J. Neurol. Sci. 2010, 292, 89–95. [Google Scholar] [CrossRef]
  146. Suh, J.H.; Wang, H.; Liu, R.-M.; Liu, J.; Hagen, T.M. (R)-alpha-lipoic acid reverses the age-related loss in GSH redox status in post-mitotic tissues: Evidence for increased cysteine requirement for GSH synthesis. Arch. Biochem. Biophys. 2004, 423, 126–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Ersahin, M.; Toklu, H.Z.; Çetinel, Ş.; Yüksel, M.; Erzik, C.; Berkman, M.Z.; Yeğen, B.Ç.; Sener, G.; Yeǧen, B.Ç. Alpha Lipoic Acid Alleviates Oxidative Stress and Preserves Blood Brain Permeability in Rats with Subarachnoid Hemorrhage. Neurochem. Res. 2009, 35, 418–428. [Google Scholar] [CrossRef] [PubMed]
  148. Camiolo, G.; Tibullo, D.; Giallongo, C.; Romano, A.; Parrinello, N.L.; Musumeci, G.; di Rosa, M.; Vicario, N.; Brundo, M.V.; Amenta, F.; et al. α-Lipoic Acid Reduces Iron-induced Toxicity and Oxidative Stress in a Model of Iron Overload. Int. J. Mol. Sci. 2019, 20, 609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Ibrahim, F.; Nomier, M.A.; Sabik, L.M.E.; Shaheen, M.A. Manganese-induced neurotoxicity and the potential protective effects of lipoic acid and Spirulina platensis. Toxicol. Mech. Methods 2020, 30, 497–507. [Google Scholar] [CrossRef] [PubMed]
  150. Al-Otaibi, S.S.; Arafah, M.M.; Sharma, B.; Alhomida, A.S.; Siddiqi, N.J. Synergistic Effect of Quercetin and α-Lipoic Acid on Aluminium Chloride Induced Neurotoxicity in Rats. J. Toxicol. 2018, 2018, 1–8. [Google Scholar] [CrossRef] [Green Version]
  151. Smirnova, J.; Kabin, E.; Järving, I.; Bragina, O.; Tõugu, V.; Plitz, T.; Palumaa, P. Copper(I)-binding properties of de-coppering drugs for the treatment of Wilson disease. α-Lipoic acid as a potential anti-copper agent. Sci. Rep. 2018, 8, 1463. [Google Scholar] [CrossRef]
  152. Khalaf, A.A.; Zaki, A.R.; Galal, M.K.; Ogaly, H.A.; Ibrahim, M.A.; Hassan, A. The potential protective effect of α-lipoic acid against nanocopper particle–induced hepatotoxicity in male rats. Hum. Exp. Toxicol. 2016, 36, 881–891. [Google Scholar] [CrossRef]
  153. Bjørklund, G.; Crisponi, G.; Nurchi, V.M.; Cappai, R.; Djordjevic, A.B.; Aaseth, J. A Review on Coordination Properties of Thiol-Containing Chelating Agents Towards Mercury, Cadmium, and Lead. Molecules 2019, 24, 3247. [Google Scholar] [CrossRef] [Green Version]
  154. Androne, L.; Gavan, N.A.; Veresiu, I.A.; Orasan, R. In vivo effect of lipoic acid on lipid peroxidation in patients with diabetic neuropathy. Vivo 2000, 14, 327–330. [Google Scholar]
  155. Freitas, R. The evaluation of effects of lipoic acid on the lipid peroxidation, nitrite formation and antioxidant enzymes in the hippocampus of rats after pilocarpine-induced seizures. Neurosci. Lett. 2009, 455, 140–144. [Google Scholar] [CrossRef] [PubMed]
  156. Haugaard, N.; Levin, R.M.; Surname, F. Regulation of the activity of choline acetyl transferase by lipoic acid. Mol. Cell. Biochem. 2000, 213, 61–63. [Google Scholar] [CrossRef] [PubMed]
  157. Quinn, J.F.; Bussiere, J.R.; Hammond, R.S.; Montine, T.J.; Henson, E.; Jones, R.E.; Stackman, R.W. Chronic dietary α-lipoic acid reduces deficits in hippocampal memory of aged Tg2576 mice. Neurobiol. Aging 2007, 28, 213–225. [Google Scholar] [CrossRef] [PubMed]
  158. Hager, K.; Kenklies, M.; McAfoose, J.; Engel, J.; Münch, G. Alpha-lipoic acid as a new treatment option for Alzheimer’s disease—A 48 months follow-up analysis. J. Neural. Transm. Suppl. 2007, 72, 189–193. [Google Scholar]
  159. Fava, A.; Pirritano, D.; Plastino, M.; Cristiano, D.; Puccio, G.; Colica, C.; Ermio, C.; de Bartolo, M.; Mauro, G.; Bosco, D. The Effect of Lipoic Acid Therapy on Cognitive Functioning in Patients with Alzheimer’s Disease. J. Neurodegener. Dis. 2013, 2013, 454253. [Google Scholar] [CrossRef]
  160. Rosini, M.; Andrisano, V.; Bartolini, M.; Bolognesi, M.L.; Hrelia, P.; Minarini, A.; Tarozzi, A.; Melchiorre, C. Rational Approach to Discover Multipotent Anti-Alzheimer Drugs. J. Med. Chem. 2005, 48, 360–363. [Google Scholar] [CrossRef]
  161. Rosini, M.; Simoni, E.; Bartolini, M.; Tarozzi, A.; Matera, R.; Milelli, A.; Hrelia, P.; Andrisano, V.; Bolognesi, M.L.; Melchiorre, C. Exploiting the lipoic acid structure in the search for novel multitarget ligands against Alzheimer’s disease. Eur. J. Med. Chem. 2011, 46, 5435–5442. [Google Scholar] [CrossRef] [PubMed]
  162. Koufaki, M.; Kiziridi, C.; Nikoloudaki, F.; Alexis, M.N. Design and synthesis of 1,2-dithiolane derivatives and evaluation of their neuroprotective activity. Bioorg. Med. Chem. Lett. 2007, 17, 4223–4227. [Google Scholar] [CrossRef]
  163. Estrada, M.; Pérez, C.; Soriano, E.; Laurini, E.; Romano, M.; Pricl, S.; Morales-García, A.J.; Pérez-Castillo, A.; Rodríguez-Franco, M.I. New neurogenic lipoic-based hybrids as innovative Alzheimer’s drugs with sigma-1 agonism and beta-secretase inhibition. Future Med. Chem. 2016, 8, 1191–1207. [Google Scholar] [CrossRef]
  164. Tu, Y.-L.; Chen, Q.-H.; Wang, S.-N.; Uri, A.; Yang, X.-H.; Chu, J.-Q.; Chen, J.-K.; Luo, B.-L.; Chen, X.-H.; Wen, S.-J.; et al. Discovery of lipoic acid-4-phenyl-1H-pyrazole hybrids as novel bifunctional ROCK inhibitors with antioxidant activity. RSC Adv. 2016, 6, 58516–58520. [Google Scholar] [CrossRef]
  165. Jones, M.; Wang, J.; Harmon, S.; Kling, B.; Heilmann, J.; Gilmer, J.F. Novel Selective Butyrylcholinesterase Inhibitors Incorporating Antioxidant Functionalities as Potential Bimodal Therapeutics for Alzheimer’s Disease. Molecules 2016, 21, 440. [Google Scholar] [CrossRef] [Green Version]
  166. Jalili-Baleh, L.; Forootanfar, H.; Küçükkılınç, T.T.; Nadri, H.; Abdolahi, Z.; Ameri, A.; Jafari, M.; Ayazgok, B.; Baeeri, M.; Foroumadi, A.; et al. Design, synthesis and evaluation of novel multi-target-directed ligands for treatment of Alzheimer’s disease based on coumarin and lipoic acid scaffolds. Eur. J. Med. Chem. 2018, 152, 600–614. [Google Scholar] [CrossRef]
  167. Pagoni, A.; Marinelli, L.; di Stefano, A.; Ciulla, M.; Turkez, H.; Mardinoglu, A.; Vassiliou, S.; Cacciatore, I. Novel anti-Alzheimer phenol-lipoyl hybrids: Synthesis, physico-chemical characterization, and biological evaluation. Eur. J. Med. Chem. 2020, 186, 111880. [Google Scholar] [CrossRef] [PubMed]
  168. Michalska, P.; Tenti, G.; Satriani, M.; Cores, A.; Ramos, M.T.; García, A.G.; Menéndez, J.C.; León, R. Aza-CGP37157-lipoic hybrids designed as novel Nrf2-inducers and antioxidants exert neuroprotection against oxidative stress and show neuroinflammation inhibitory properties. Drug Dev. Res. 2019, 81, 283–294. [Google Scholar] [CrossRef]
  169. Pachón-Angona, I.; Martin, H.; Chhor, S.; Oset-Gasque, M.J.; Refouvelet, B.; Marco-Contelles, J.; Ismaili, L. Synthesis of new ferulic/lipoic/comenic acid-melatonin hybrids as antioxidants and Nrf2 activators via Ugi reaction. Futur. Med. Chem. 2019, 11, 3097–3108. [Google Scholar] [CrossRef] [PubMed]
  170. Wang, R.; Zhang, L.; Liao, R.; Li, Q.; Pi, R.; Yang, X. N2L, a novel lipoic acid-niacin dimer protects HT22 cells against beta-amyloid peptide-induced damage through attenuating apoptosis. Metab. Brain Dis. 2019, 34, 1761–1770. [Google Scholar] [CrossRef] [PubMed]
  171. Uppakara, K.; Jamornwan, S.; Duan, L.-X.; Yue, K.-R.; Sunrat, C.; Dent, E.W.; Wan, S.-B.; Saengsawang, W. Novel α-Lipoic Acid/3-n-Butylphthalide Conjugate Enhances Protective Effects against Oxidative Stress and 6-OHDA Induced Neuronal Damage. ACS Chem. Neurosci. 2020, 11, 1634–1642. [Google Scholar] [CrossRef]
  172. Syed, Y.Y. Correction to: Sodium Oligomannate: First Approval. Drugs 2020, 80, 445–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Al Zahrani, N.A.; El-Shishtawy, R.M.; Asiri, A.M. Recent developments of gallic acid derivatives and their hybrids in medicinal chemistry: A review. Eur. J. Med. Chem. 2020, 204, 112609. [Google Scholar] [CrossRef]
Figure 1. Biosynthesis of polyamines. Briefly, arginase operates the removal of guanidino group from essential amino acid arginine, producing L-ornithine. Then, decarboxylation of L-ornithine mediated by ODC enzyme generates putrescine that is converted to spermidine by spermidine synthase that adds an aminopropyl group to the molecule. Finally, addition of a second aminopropyl moiety to spermidine generates spermine.
Figure 1. Biosynthesis of polyamines. Briefly, arginase operates the removal of guanidino group from essential amino acid arginine, producing L-ornithine. Then, decarboxylation of L-ornithine mediated by ODC enzyme generates putrescine that is converted to spermidine by spermidine synthase that adds an aminopropyl group to the molecule. Finally, addition of a second aminopropyl moiety to spermidine generates spermine.
Antioxidants 10 00367 g001
Figure 2. Structures of the main phenolic acids.
Figure 2. Structures of the main phenolic acids.
Antioxidants 10 00367 g002
Figure 3. Ellagic acid (EA) and its urolithins metabolites after gut microbiota metabolism. Briefly, after intestinal uptake the hydrolysis and subsequent decarboxylation of EA’s lactone generates urolithin D (UD), that provides UC losing a hydroxyl group. Then, the removal of a phenolic group in 9 or 8 position affords UA or iso-UA, respectively. Finally, UB is produced by removal of the phenolic group in 8 or 9.
Figure 3. Ellagic acid (EA) and its urolithins metabolites after gut microbiota metabolism. Briefly, after intestinal uptake the hydrolysis and subsequent decarboxylation of EA’s lactone generates urolithin D (UD), that provides UC losing a hydroxyl group. Then, the removal of a phenolic group in 9 or 8 position affords UA or iso-UA, respectively. Finally, UB is produced by removal of the phenolic group in 8 or 9.
Antioxidants 10 00367 g003
Figure 4. Lipoic acid (LA) and its reduced form, dihydrolipoic acid (DHLA).
Figure 4. Lipoic acid (LA) and its reduced form, dihydrolipoic acid (DHLA).
Antioxidants 10 00367 g004
Table 1. Main physiological and pathological roles of ROS.
Table 1. Main physiological and pathological roles of ROS.
Physiological Role of ROSRef.Pathological Role of ROSRef.
Signalling between mitochondria and surrounding cells[20]mDNA damage, deletion, and mutation [21]
Regulation of cellular proliferation, differentiation and apoptosis[18,19]Mitochondrial membrane permeability alteration and mitochondrial failure[22,25,27,31]
Induction of MAPKs activation in cardiovascular system.[20]Lipid peroxidation[27]
Influence on pro-survival transcription factors (i.g. Nrf2 and NF-κB).[21]ETC enzymes malfunctions[29]
Adaption and regulation of hypoxia[20]Promotion of inflammation[16]
Regulation of immune functions [19,20]Metallostasis and metal accumulation[40,42]
Induction of autophagy[20]Proteostasis and misfolded proteins clearance impairment[51,53]
Table 2. Polyamines derivatives and their main biological effects.
Table 2. Polyamines derivatives and their main biological effects.
EntryScaffolds CombinationStructureEffectsRef.
11-aminoindan beared with polyamine scaffold Antioxidants 10 00367 i001Neuroprotection against NMDA toxicity and ischemia damages
No neurotoxicity
[85]
21,4-benzoquinone and polyamine structure of caproctamine Antioxidants 10 00367 i002↓Aβ aggregation
↓tau phosphorylation
↑antioxidant activity
↓AChE
↓BACE-1
[87]
3Ferulic acid-memoquin hybrids Antioxidants 10 00367 i003↓AChE
↓BuChE
↓self-induced Aβ1-42 aggregation
no cytotoxicity in SH-SY5Y cells
good BBB predicted permeability
[88]
4Genistein with polyamines Antioxidants 10 00367 i004↓AChE
↓BuChE
Fe3+/Cu2+/Zn2+ chelation
no HepG-2 cell cytotoxicity
[89]
53,5-dibenzylidenepiperidin-4-
one functionalized with spermine
Antioxidants 10 00367 i005↓Aβ42 aggregation
no antioxidant properties in T67 cells
neuroprotection and no cytoxicity in vitro
[90]
6Dicaffeoylsper-midine cyclized derivatives Antioxidants 10 00367 i006Antioxidant activity
↑memory and learning in fruit flies model
[92]
Table 3. Phenolic acids hybrids and their main biological effects.
Table 3. Phenolic acids hybrids and their main biological effects.
EntryScaffolds CombinationStructureEffectsRef.
7Tacrine linked with ferulic acid Antioxidants 10 00367 i007↓ Aβ-aggregation
↓ ROS production
↓ AChE
↑ cognitive functions
↑ SOD/ChAT
[110]
8Tacrine and functionalized ferulic acid Antioxidants 10 00367 i008↓ Aβ-self aggregation
↓ AChE
↓ BuChE
↑ memory
no hepatotoxicity
[111]
9Ferulic and caffeic merged with serotonin Antioxidants 10 00367 i009↑ antioxidant activity
↓ BACE-1
[112]
10Aromatic amides and esters of caffeic acid Antioxidants 10 00367 i010↓MAO-A/MAO-B
↑ antioxidant activity
[113]
11Hydroxycinnamic acids and NBP (donepezil) Antioxidants 10 00367 i011↓MAO-A/MAO-B
↓AChE
↓BuChE
↑ antioxidant activity
[114]
12Hydroxycinnamic scaffolds and DBMA (AP2238) Antioxidants 10 00367 i012↓MAO-A/MAO-B
↓AChE
↓BuChE
↑ antioxidant activity
[114]
13Caffeic acid with hydrophobic moieties Antioxidants 10 00367 i013↓Aβ1-40 self-aggregation
↑ antioxidant activity
neuroprotection in SH-SY5Y cells
[115]
14Rivastigmine with GA Antioxidants 10 00367 i014↑ antioxidant activity
Cu2+ chelating properties
↓ChEs
↓Aβ self-aggregation
neuroprotective effects in vitro
no cytotoxicity
[116]
15Caffeic acid and diallyl sulfide Antioxidants 10 00367 i015↓Aβ42 self-aggregation
↑cytoprotection against H2O2-induced damages
↓p53 alteration induced by Aβ
[117]
16Ferulic core merged with 1,2,3,4-tetrahydroisoquinoline and (benzyl(ethyl)amino)butoxy scaffold Antioxidants 10 00367 i016↑antioxidant activity
↓AChE
↓BuChE
↓ MAO-A/MAO-B
↓ Aβ self-aggregation
↑ self-induced Aβ1-42 fibrils disaggregation
↑neuroprotective effect in SH-5YSY cells
↑autophagy in U87 cells
↑motility in Zebrafish model
↓ Aβ1-40-induced vascular injury in Zebrafish model
↑In vivo cognitive functions
[118]
17Ferulic acid merged with 1,3,4-oxadiazole scaffold Antioxidants 10 00367 i017↓ChEs
↓BACE-1
↓ Aβ self-aggregation
↓ Aβ AChE-induced aggregation
neuroprotective effects in vitro
↑In vivo cognitive functions
[119]
Table 4. Urolithins derivatives and their main biological effects.
Table 4. Urolithins derivatives and their main biological effects.
EntryScaffolds CombinationStructureEffectsRef.
18Urolithin scaffold with rivastigmine portion Antioxidants 10 00367 i018↓AChE
↓BuChE
[136]
19Urolithin scaffold with donepezil-like moieties Antioxidants 10 00367 i019↓AChE
↓BuChE
[136]
20Donepezil-like urolithin and tetrahydrourolithin derivatives Antioxidants 10 00367 i020↓AChE
↓BuChE
↓ AChE induced Aβ aggregation
[137]
21Nitro- and bromo-derivatives of urolithins Antioxidants 10 00367 i021CK2 inhibition
Selectivity in other kinases panel
[140]
22Tetrahydrourolithin scaffold linked with donepezil moiety Antioxidants 10 00367 i022AChE/BuChE inhibition
MAO-B inhibition
BBB permeability
no cytotoxicity in brain and liver cells
[141]
Table 5. Lipoic acid hybrids and their main biological effects.
Table 5. Lipoic acid hybrids and their main biological effects.
EntryScaffold CombinationStructureEffectsRef.
23Lipoic acid and tacrine Antioxidants 10 00367 i023↑ROS protection
↓AChE
↓BuChE
↓ AChE-induced Aβ aggregation
[160]
24Dopamine and LA linked by tetrazole ring Antioxidants 10 00367 i024↑antioxidant activity
neuroprotection in vitro
[162]
25LA-NBP and LA-DBMA conjugation Antioxidants 10 00367 i025↓AChE
↓BuChE
↓BACE-1
↑antioxidant activity
σ1R agonism
good BBB permeability prediction
neuroprotection in vitro
[163]
26LA-4-Phenyl-1H-pyrazole
derivatives
Antioxidants 10 00367 i026ROCK1/ROCK2 inhibition
↓ROS
↑GSH
vasorelaxant activity
[164]
27Lipoic isosorbide-2-benzylcarbamate Antioxidants 10 00367 i027↓ROS
↓BuChE
↓cytotoxicity in treated HT-22 cells
[165]
28LA and coumarin scaffold linked bridged with triazole Antioxidants 10 00367 i028↓AChE
↓BuChE
↓ Aβ peptide aggregation
↓intracellular ROS
neuroprotection against H2O2− or Aβ1-42-induced cytotoxicity in SH-SY5Y cell lines
Selective Cu/Fe chelation
[166]
29FA/CA-LA hybrids Antioxidants 10 00367 i029↓Aβ1-42-induced neurotoxicity in SH-SY5Y cells
↑protection in H2O2-insulted cells
no cytotoxicity
[167]
30Lipoic-functionalized benzodiazepine Antioxidants 10 00367 i030↑ROS scavenging
↑Nrf2-ARE pathway
↑HO-1/GCLc
neuroprotection in in vitro model
no cytotoxicity
no hepatotoxicity
[168]
31Lipoic-melatonin hybrids Antioxidants 10 00367 i031↑ROS scavenging
↑Nrf2-ARE pathway
antioxidant
activity and neuroprotection in vitro
no cytotoxicity
[169]
32LA-niacin hybrids Antioxidants 10 00367 i032↓Aβ1-42-induced cytotoxicity in HT22 cells
↓mitochondrial dysfunctions
↓intracellular ROS
↑SOD, CAT, GPx
↓apoptosis in Aβ1-42treated cells
[170]
33LA-3-n-butylphthalide amide Antioxidants 10 00367 i033↓intracellular ROS
↑direct ROS-scavenger
↓ H2O2-induced cell death
↑GSH
↓ H2O2-induced damage in cortical neurons
↓6-OHDA-induced neuronal damage in SH-5YSY cells
[171]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bacci, A.; Runfola, M.; Sestito, S.; Rapposelli, S. Beyond Antioxidant Effects: Nature-Based Templates Unveil New Strategies for Neurodegenerative Diseases. Antioxidants 2021, 10, 367. https://doi.org/10.3390/antiox10030367

AMA Style

Bacci A, Runfola M, Sestito S, Rapposelli S. Beyond Antioxidant Effects: Nature-Based Templates Unveil New Strategies for Neurodegenerative Diseases. Antioxidants. 2021; 10(3):367. https://doi.org/10.3390/antiox10030367

Chicago/Turabian Style

Bacci, Andrea, Massimiliano Runfola, Simona Sestito, and Simona Rapposelli. 2021. "Beyond Antioxidant Effects: Nature-Based Templates Unveil New Strategies for Neurodegenerative Diseases" Antioxidants 10, no. 3: 367. https://doi.org/10.3390/antiox10030367

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