Next Article in Journal
Association of Statin Therapy with Functional Outcomes and Survival in Intracerebral and Subarachnoid Hemorrhage
Next Article in Special Issue
Cardiac Autonomic Modulation and Cognitive Performance in Community-Dwelling Older Adults: A Preliminary Study
Previous Article in Journal
Cardiac CT in Large Vessel Occlusion Stroke for the Evaluation of Non-Thrombotic and Non-Atrial-Fibrillation-Related Embolic Causes
Previous Article in Special Issue
Impact of Mast Cell Activation on Neurodegeneration: A Potential Role for Gut–Brain Axis and Helicobacter pylori Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Role of Heterocyclic Compounds in Neurodegenerative Diseases: Insights from Alzheimer’s and Parkinson’s Diseases

Department of Life Sciences, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Authors to whom correspondence should be addressed.
Neurol. Int. 2025, 17(2), 26; https://doi.org/10.3390/neurolint17020026
Submission received: 6 December 2024 / Revised: 17 January 2025 / Accepted: 21 January 2025 / Published: 7 February 2025
(This article belongs to the Collection Advances in Neurodegenerative Diseases)

Abstract

:
Alzheimer’s and Parkinson’s are the most common neurodegenerative diseases (NDDs). The development of aberrant protein aggregates and the progressive and permanent loss of neurons are the major characteristic features of these disorders. Although the precise mechanisms causing Alzheimer’s disease (AD) and Parkinson’s disease (PD) are still unknown, there is a wealth of evidence suggesting that misfolded proteins, accumulation of misfolded proteins, dysfunction of neuroreceptors and mitochondria, dysregulation of enzymes, and the release of neurotransmitters significantly influence the pathophysiology of these diseases. There is no effective protective medicine or therapy available even with the availability of numerous medications. There is an urgent need to create new and powerful bioactive compounds since the number of people with NDDs is rising globally. Heterocyclic compounds have consistently played a pivotal role in drug discovery due to their exceptional pharmaceutical properties. Many clinically approved drugs, such as galantamine hydrobromide, donepezil hydrochloride, memantine hydrochloride, and opicapone, feature heterocyclic cores. As these heterocyclic compounds have exceptional therapeutic potential, heterocycles are an intriguing research topic for the development of new effective therapeutic drugs for PD and AD. This review aims to provide current insights into the development and potential use of heterocyclic compounds targeting diverse therapeutic targets to manage and potentially treat patients with AD and PD.

1. Introduction

A broad spectrum of neurological conditions characterized by alterations in the composition and operation of the central nervous system (CNS) are referred to as neurodegenerative diseases (NDDs). Parkinson’s disease (PD) and Alzheimer’s disease (AD) are the two most common NDDs. The main characteristic of these diseases is the massive deposition of misfolded protein aggregates consequent to aberrant production or overexpression of specific proteins [1,2]. In AD, the most prevalent cause of dementia worldwide, amyloid-β (Aβ) plaques and tau tangles are common hallmarks. The second most widespread protein misfolding disorder associated with dementia involves Lewy body (LB) pathology. This condition is defined by the intracellular aggregation of misfolded α-synuclein (α-syn) within cells, forming Lewy bodies and Lewy dystrophic neurites. These abnormalities are linked to disorders like dementia with Lewy bodies (DLB) and PD, collectively known as Lewy body disease (LBD). AD is a progressive NDD that affects many people worldwide [3,4], characterized by declining memory and cognitive dysfunction [5], and is the most prevalent type of dementia. The loss of synapses, the build-up of hyperphosphorylated tau protein inside cells, and the existence of extracellular Aβ peptide plaques are some of the main characteristics of AD [6]. Biochemical studies, along with research on transplanted neurons in PD patients and investigations using cell and animal models, suggest that the abnormal aggregation of α-synuclein (α-syn) and the spreading of pathology between the gut, brainstem, and higher brain regions may play a crucial role in the onset and progression of PD [7]. This condition is characterized by a progressive asymmetric slowness of movement (bradykinesia), rigidity, tremors, and gait disturbances, which occur alongside neuronal loss and the formation of α-syn-rich protein aggregates in the substantia nigra, referred to as Lewy bodies and Lewy neurites [8]. PD has both motor and nonmotor symptoms; whereas motor symptoms include tremors, resting tremors, bradykinesia, and stiffness, which are linked to a deficiency of dopamine in the striatum, nonmotor symptoms encompass sleep disorders, feelings of sadness, and cognitive impairments [9,10].
The pathogeneses of AD and PD share common pathways of degeneration. The most common etiological factors include oxidative stress (OS), mitochondrial stress, neuroinflammation, neurodegeneration, and the build-up of insoluble proteins, as shown in Figure 1.

2. Therapeutic Targets in Alzheimer’s and Parkinson’s Disease

The exact etiologies of AD and PD are still unknown, and there is not a safe and efficient treatment to stop these diseases’ progression or cure them [11,12]. A recent review article highlights the ongoing drug trials for AD and mentions that 187 trials are currently underway to evaluate 141 AD drugs, with over 57,000 participants [13], while another article for PD identified 145 registered and ongoing clinical trials [14]. For safe and effective treatment development, the well-detailed pathogenesis needs to be understood first. The primary causes of AD pathogenesis include tau NFTs, Aβ plaques, neuroinflammation, OS, cholinergic dysfunction, glutamate excitotoxicity, and alterations in neurotrophin levels [15,16,17,18,19,20,21]. These pathogenic factors are the possible major therapeutic targets for AD treatment and management [18,19,21,22,23,24,25].
PD is a multifactorial disorder with similar phenotypes. According to recent research studies, classifying patients based on genes involved in PD could be the most practical approach, as clinical phenotypes or biomarkers may not accurately represent the disorder. Mutations are common genetic causes of PD, with patients with these variants having varying clinical characteristics and pathologies [26]. Several autosomal dominant and autosomal recessive genes have been identified, with SNCA (Synuclein Alpha), LRRK2 (Leucine-Rich Repeat Kinase 2), PRKN (Parkin RBR E3 Ubiquitin Protein Ligase), PINK1 (PTEN-induced kinase 1), and GBA1 (glucosylceramidase beta 1) being the most commonly linked genes associated with PD pathology [12]. α-Syn aggregates are Lewy bodies’ primary components, which are PD’s pathological hallmarks [27]. Experimental evidence suggested that a decrease in glucocerebrosidase activity is associated with the accumulation of α-syn [28]. Similarly, a decrease in the TMEM175 (Transmembrane protein 175) level causes the accumulation of α-syn in neurons, leading to the loss of dopaminergic neurons [29]. The PRKN gene encodes the parkin protein that preferentially protects dopaminergic neurons from mitochondrial stress, and loss of this exacerbates mitochondrial dysfunction in neurons [30]. The PINK1/parkin pathway has been shown to significantly enhance the function of PINK1 and parkin, a crucial process in mitophagy essential for maintaining mitochondrial health, and the most potent therapeutic target for PD [31]. Several recent preclinical and clinical candidates are being developed to target various proteins and receptors associated with NDDs, as listed in Figure 2. These potential targets include tau protein [32,33,34,35], amyloid beta [36,37,38,39], α-syn [40,41], acetylcholinesterase (AChE) [42,43,44,45,46], butyrylcholinesterase (BChE) [47,48], tyrosine kinases [49], glycogen synthase kinase-3 [50,51,52], γ- and β-secretases [53,54], monoamine oxidase (MAO) [55], N-methyl-D-aspartate (NMDA) receptors [56,57,58], muscarinic acetylcholine receptors [59,60], dopamine D2 receptors [61,62], GABA-A receptors [63,64], the 5-hydroxytryptamine (5-HT6) receptor [65], the OS pathway, and mitochondrial dysfunction [66,67,68,69]. Major therapeutic targets are discussed briefly one by one.
Medications aimed at alleviating dementia symptoms fall into two main categories: those that address cognitive decline and those that target behavioral and psychological symptoms of dementia (BPSD). BPSD mitigators include suvorexant, an orexin receptor antagonist for sleep disorders, and brexpiprazole, an atypical antipsychotic suitable for treating agitation in dementia patients with its once-daily dosing. Examples of cognitive decline mitigators include glutamate inhibitors, like memantine, and cholinesterase inhibitors, such as donepezil, rivastigmine, and galantamine [70].
Pharmaceuticals that may slow down the clinical decline of AD patients and treatments that may temporarily relieve some of the disease’s related symptoms are the two types of medications approved by the US Food and Drug Administration (FDA) [71]. Figure 3 and Figure 4 illustrate the FDA-approved therapeutic drugs for AD and PD, respectively. While several therapies have received FDA approval to alleviate symptoms, none can slow, stop, or reverse the progression of these diseases. The primary constraint of these symptomatic treatments is the failure to address the underlying disease progression. AChE inhibitors, such as Razadyne, Aricept, and Exelon, are used in the management of dementia and cognitive decline symptoms. These inhibitors facilitate neuronal transmission and cognitive function by maintaining the levels of acetylcholine (ACh) in synaptic gaps and preventing AChE from degrading it [72]. Moreover, memantine hydrochloride (Namenda), an NMDA receptor antagonist, aims to prevent excess glutamate from overstimulating neurons and causing excitotoxic damage. Additionally, Belsomra, a dual orexin receptor antagonist originally used for insomnia, is available to address sleep disorders associated with AD [72]. All these FDA-approved drugs are heterocyclic, and their therapeutic categories, targets, and applications are summarized in Table 1.
Drugs like Sinemet, Paracopa, Rytary, and Duopa which contain levodopa and carbidopa are the most widely used approved treatments for PD motor symptoms. Carbidopa, a dopa decarboxylase inhibitor, decreases levodopa’s extracerebral metabolism before it crosses the protective blood–brain barrier, increasing the neurotransmitter’s brain bioavailability. Levodopa is a dopamine precursor. COMT inhibitors, which include Comtan, Tasmar, and Ogentys, are another family of FDA-approved therapies. These drugs have shown promise when used with carbidopa/levodopa therapy. By blocking the COMT enzyme, these medications stop levodopa from being broken down extracerebrally and raise its plasma levels. By attaching to dopamine receptors, dopamine agonists can penetrate the blood–brain barrier and replicate the effects of dopamine, hence lowering dyskinesia. Levodopa dosage requirements are decreased with carbidopa and COMT inhibitors. Currently, Mirapex, Requip, Apokyn, Kynmobi, and Neupro are FDA-approved medications that function as dopamine agonists. In the brain, the enzyme MAO-B degrades dopamine. Eldepryl, Zelapar, Azilect, and Xadago are examples of MAO-B inhibitors that have been approved by the FDA. The FDA has approved NMDA receptor antagonists such as Osmolex, Gocovri, and Symmetrel to treat dyskinesia. In PD patients, adenosine 2A antagonists like Nourianz exhibit neuroprotective effects. Anticholinergic medications, like Cogentin and Artane, have been licensed for the treatment of PD patients’ tremors by balancing the imbalance between dopamine and ACh. It is important to remember that every approved treatment is symptomatic [72].

2.1. Proteins

2.1.1. Aggregated Protein

The amyloid peptide is a major component of amyloid plaques and has been implicated in numerous studies as a major pathogenic factor of AD. Tau protein (another important histopathological characteristic of AD) clumps inside neurons make up the majority of NFTs [73].
Multiple locations of tau phosphorylation regulate the protein’s ability to connect to microtubules; hyperphosphorylation, on the other hand, causes AD pathogenesis. According to research, tau impairment is most likely the effector molecule of neurodegeneration, and Aβ may start the chain of events leading to tau hyperphosphorylation. By mediating the activation of several pathways, Aβ speeds up the hyperphosphorylation of tau and exacerbates tau-induced neurotoxicity, which results in AD [74]. More significantly, it has now been demonstrated that downregulating tau partially restores transcriptional perturbations and that Aβ and tau cooperate to hinder the transcription of genes implicated in synaptic function [75,76]. Recent research has shown links between abnormal tau and Aβ protein behavior and several neurological conditions, including AD, DPD, and ALS, as well as retinal neurodegenerative conditions, including age-related macular degeneration and glaucoma [77].
To investigate the possible molecular routes through which tau may contribute to the pathogenesis of PD, several studies have used transgenic PD mouse lines in which tau is altered, decreased, or removed [78]. α-Syn is predominantly found at the presynaptic terminals and associated with synaptic vesicles [79]. α-Syn exhibits a range of species, including monomers, oligomers, and fibrils, each of which has distinct neurotoxic characteristics [80]. One of the main molecular mechanisms behind the pathophysiology of PD is the aggregation of α-syn. These aggregates, which disturb cellular homeostasis and contribute to neurodegeneration, can form at synaptic terminals and are frequently caused by a breakdown in proteostatic defenses or aberrant interactions between α-syn and lipids [81,82]. PD patients’ autopsies also showed that tau and α-syn were colocalized in LBs. Tau interacts with α-syn and affects the pathogenesis of α-syn in Parkinson’s disease, according to experimental studies [78]. Tau’s interaction with α-syn and the acceleration of its aggregation were demonstrated experimentally in a recent study. Tau-modified α-syn fibrils exhibit increased seeding activity in comparison to pure α-syn fibrils, causing neurotoxicity, synaptic impairment, and mitochondrial dysfunction in vitro. Compared to mice injected with pure α-syn fibrils, mice injected with tau-modified α-syn fibrils in the striatum experience more severe α-syn disease, motor dysfunction, and cognitive impairment. Both in mice injected with α-syn fibrils and α-syn A53T transgenic animals, tau knockout reduces the spread of α-syn pathology and PD-like symptoms. To sum up, tau promotes α-syn aggregation and spread in PD [83]. Therefore, to improve therapeutic efficacy, it is essential to understand the molecular properties of pathogenic proteins deposited in each non-AD NDD [84].

2.1.2. Dopamine Transporter (DAT)

The dopamine transporter (DAT) is a transmembrane protein that belongs to the Na+/Cl-dependent neurotransmitter transporter family. It consists of 12 transmembrane helices and undergoes conformational changes upon ligand binding. These structural changes are essential for facilitating the translocation of dopamine into the neuron [85]. DAT facilitates the transport of extracellular dopamine into the intracellular space, playing a crucial role in regulating dopamine neurotransmission. A reduction in DAT density is implicated in PD, and, simultaneously, dopamine turnover is elevated in early symptomatic PD [85,86]. Recent studies have revealed that blocking DAT shows the potential to increase dopamine levels and slow disease progression, highlighting its therapeutic promise [87]. Nigrostriatal dopaminergic dysfunction contributes to parkinsonism and cognition independently of extranigral cortical thinning in patients with AD [88].

2.2. Enzymes

2.2.1. Cholinesterase Enzymes

Cholinesterases are crucial enzymes found in both the central and peripheral nervous systems. These enzymes facilitate neurotransmission by enabling the conduction of nerve impulses at cholinergic synapses. The two main forms of cholinesterases are AChE and BChE. BChE, closely related to AChE, acts as a co-regulator of cholinergic neurotransmission by hydrolyzing ACh. Research has shown that during the progression of AD, there is an increase in BChE activity in the most affected brain regions, such as the temporal cortex and hippocampus. This heightened BChE activity is also significant in the early stages of senile plaque development, particularly in relation to Aβ aggregation. Consequently, inhibiting both AChE and BChE has been recognized as an important strategy for effectively managing AD, as it leads to increased availability of ACh in the brain and a reduction in Aβ deposition. One common method for treating a variety of mental illnesses is the inhibition of the AChE and BChE enzymes. Cholinesterase inhibitors are used in this context to alleviate the symptoms of neurological conditions such as AD and dementia. The role of several heterocyclic scaffolds comprising N, O, and S in the design and development of novel potential AChE and BChE inhibitors to treat AD has been investigated in recent years. Additionally, a thorough structure–activity relationship (SAR) has been developed for the potential future development of new medications to treat AD. New powerful cholinesterase inhibitors have been designed using the majority of heterocyclic motifs [89]. The oldest and least expensive antiparkinsonian medicine is an anticholinergic agent used to treat PD [90].

2.2.2. Carbonic Anhydrases (CAs)

CAs, metalloenzymes that reversibly hydrate carbon dioxide, play an antioxidant role in cells during oxidative processes. They are crucial in brain pH control, neuronal excitability, and cognition and can impact animal learning. Moreover, CA II is associated with abundant plaque proteins, suggesting a central role in plaque development. Inhibiting mitochondrial CAs may slow NDD progression [91]. Mitochondrial proteomic profiling reveals increased CA II in aging and neurodegeneration, suggesting that targeting CA II associated with mitochondria could specifically modulate age-related impairments and neurological diseases [92]. Recent research indicates that CA activation may be a viable therapeutic strategy for the treatment of AD since it is implicated in brain functions necessary for the transmission of neuronal messages [93] and various illnesses that are marked by cognitive issues and memory loss. Certain substances that can activate CAs have been discovered and suggested as a potential solution to neurodegenerative issues [94].

2.2.3. Monoamine Oxidase Enzyme (MAO)

MAO, an enzyme anchored to the mitochondrial membrane, catalyzes the oxidative deamination of both endogenous and exogenous monoamines, including key neurotransmitters such as dopamine, serotonin, and adrenaline. Recent research has highlighted MAO as a promising therapeutic target for NDDs, as its role in neurodegeneration has been increasingly recognized [95,96,97]. The MAO enzyme, present in both the brain and peripheral tissues, plays a crucial role in regulating neurotransmitter levels by deaminating biogenic amines. MAO exists in two isoforms, MAO-A and MAO-B, which are differentiated by their three-dimensional structures, substrate preferences, inhibitor selectivity, and amino acid sequences [55,98]. Both isoforms of MAO and their inhibitors, such as clorgiline and selegiline—commonly prescribed for neurological and neurodegenerative disorders—interact with substrates like dopamine, tyramine, and tryptamine. MAO-A inhibitors are primarily utilized for treating depression, while MAO-B inhibitors are widely used in managing PD. Additionally, the potential efficacy of MAO-B inhibitors in treating AD is an area of ongoing research [55].
Activated MAO facilitates two consecutive cleavages of APP by β-secretase and γ-secretase, leading to the aggregation of Aβ. Additionally, active MAO exacerbates cholinergic neuronal damage and dysfunction in the cholinergic system, contributing to the formation of NFTs and subsequent cognitive decline. The overall anti-AD effect of MAO inhibition is primarily attributed to the reduction in OS induced by MAO enzymes [99] and neurochemicals like dopamine, serotonin, and adrenaline [100].
MAO inhibitors have recently been developed using a variety of structural scaffolds, including chalcones, conjugated dienones, isatins, chromones, coumarins, pyrazolines, quinazolines, β-carbolines, and compounds generated from benzyloxy [100,101,102].

2.2.4. Catechol-O-Methyltransferase (COMT) Inhibitors

COMT is an enzyme that catalyzes the transfer of a methyl group from S-adenosylmethionine and degrades catecholamines. COMT regulates dopamine metabolism in the prefrontal cortex, affecting neuropsychiatric disorders and cognitive disruptions. It plays a crucial role in regulating prefrontal levels, affecting both the CNS and peripheral regions [103]. The two primary isoforms of the COMT protein are soluble (S-COMT) and membrane-bound (MB-COMT). In the brain, MB-COMT is the predominant kind, whereas the S-COMT isoform is more common in other organs. Through precise dopamine flux modulation, COMT plays a critical role in the prefrontal cortex’s and striatum’s dopamine degradation, which impacts cognition and cognitive control [104]. COMT inhibitors function by preventing catechols, such as dopamine, from degrading into inactive chemicals. This reduces the symptoms of CNS diseases by making L-dopa available to the brain [105]. Through its effects on the metabolism of catecholamine neurotransmitters and estrogen, COMT is emerging as a key player in the pathogenesis of AD. A subgroup of AD patients has been found to have impaired executive functioning, which is linked to a more severe pathology, a faster rate of disease development, and a lower survival rate [106].
COMT inhibitors are among the recommended first-line therapies to be combined with levodopa for managing end-of-dose motor fluctuations in patients with advanced PD [107]. The three most important COMT inhibitors are entacapone, tolcapone, and opicapone used in PD treatment and management, as represented in Figure 5 [108].

2.3. Mitochondrial Calcium Homeostasis and Oxidative Stress

Mitochondria are a hub for cellular communication and metabolism. In addition to being necessary for a variety of mitochondrial processes, ions like Ca2+ and iron can be stored within the mitochondria to preserve cellular ionic equilibrium. Ion intake, transit, and storage are coordinated by ion channels and transporters in the mitochondrial and plasma membranes. Via voltage- or ligand-gated Ca2+ channels located throughout the cell surface, Ca2+ enters the cell. Iron is transported by endocytosis from iron-bound carrier proteins like transferrin, and internal iron can be absorbed by mitochondria through a number of IMM transporters. Even though iron and Ca2+ homeostasis disruptions have been documented separately in PD and AD models, their molecular relationship and roles in disease vulnerability are still unclear [109].
A complex web of interrelated factors, including high metabolic activity, neurotransmitter autoxidation, elevated redox-active transition metal content, modest antioxidant defense, glutamate excitotoxicity, and altered Ca2+ influx and related signaling processes, make the brain especially vulnerable to oxidative insults [110,111]. Redox balance changes and corresponding changes in redox-sensitive signaling pathways can be caused by a compromised antioxidant system or by abnormal and persistent free radical production. Although brain ROS and RNS are second messengers engaged in intracellular signaling at the normal level, elevated free radical levels have detrimental effects on biological macromolecules that are linked to the pathophysiology of NDDs and the aging process [112,113]. Although the precise mechanisms behind the etiopathogenesis of AD and PD are still unknown, there is a wealth of evidence suggesting that the pathophysiology of these neurological disorders is significantly influenced by the excessive production of ROS and RNS, as well as by a depleted antioxidant system, mitochondrial dysfunction, and intracellular Ca2+ dyshomeostasis. The incidence of age-related NDDs has dramatically grown as life expectancy has increased. However, only very limited palliative care is accessible, and no effective protective medicine or therapy is available. Therefore, the development of disease-modifying treatments and preventive measures to treat AD/PD is urgently needed. Since oxidative damage and neuropathology in these disorders are caused by dysregulated Ca2+ metabolism, finding or creating substances that can restore Ca2+ homeostasis and signaling could offer a neuroprotective therapy option for NDDs [114].
According to preclinical research, controlling calcium levels may help in both in vitro and in vivo models of AD by directly blocking calcium channels or calcium-dependent downstream cascades. Calcium channel blockers, either alone or in combination with antioxidants, GSK-3 inhibitors, and cholinesterase inhibitors, have been a part of the majority of effective models to date. However, in both acute and chronic settings, it has also been demonstrated that activating specific calcium channels, including TRPV1, TRPC6, and TRPML1, has positive effects. Thus, the molecular target and its corresponding cellular physiological role, the afflicted brain circuit, and the stage of AD pathology should all be taken into account when adjusting intracellular calcium levels. For instance, the hippocampus’s synaptic plasticity processes depend on a homeostatic rise in calcium levels, but an overabundance of these levels can cause late-onset neurodegeneration, excitotoxicity, and glial-induced neuroinflammation problems across the board [115].
The study of mitochondria-targeted medication has gained attention concerning NDDs, PD, and AD. It is well accepted that the pathophysiology of PD and other related NDDs is significantly influenced by mitochondrial dysfunction [116]. However, intricate genetic control governs mitochondrial malfunction. There is mounting evidence that genes linked to Parkinson’s disease either directly or indirectly impact mitochondrial integrity. As a result, there are several clinical opportunities for tailored modulation of mitochondrial function in the therapy of PD [117]. Disturbances in iron and calcium homeostasis and accumulation are strongly linked to pathological features of PD, including intracellular α-syn deposition and dopaminergic neuronal death [118]. Several NDDs associated with OS have been documented, including PD, AD, spinocerebellar ataxia, ALS, and HD [119].

3. Therapeutic Potential of Heterocyclic Compounds in AD and PD Management

In many active medications and natural compounds, heteroatoms and heterocyclic scaffolds are commonly found as fundamental components. More than 85% of all physiologically active chemicals are heterocycles or contain a heterocycle, with nitrogen heterocycles being the most common backbone in their complex structures. These statistics reveal and highlight how important heterocycles are to contemporary medication design and discovery [120].
Heterocycles are important in medicinal chemistry and pharmaceutical industries because of their various actions and stability with different substituents. Several diseases such as cancer, AIDS, CNS disorders, CVDs, and diabetes involve the usage of drugs with heterocyclic ring systems as their major pharmacophore [121]. In essence, the ring structures of heterocycles are made up of atoms other than carbon, with oxygen, nitrogen, and sulfur serving as the most common substituents. Heterocycles can be categorized as oxygen-, nitrogen-, or sulfur-based based on the heteroatom or atoms that are present in the ring structures. Within each class, compounds are arranged according to the size of the ring structure, which is defined by the total number of atoms [122].
Numerous lead compounds with CNS action have been developed with success using polycyclic cage scaffolds. Polycyclic cage derivatives can affect a number of NDDs, including drug misuse, schizophrenia, stroke, PD, HD, and AD. When creating therapeutically active medicines to treat neurological illnesses, these cage moieties—which include derivatives of adamantane and pentacycloundecane—improve the pharmacokinetic and pharmacodynamic characteristics of conjugated parent medications [123].
Heterocycles that include nitrogen are a significant and distinct class of molecules that are used extensively in organic chemistry. In medicinal chemistry, they are unique in that they are a valuable source of therapeutic compounds. Over 75% of FDA-approved medications that are presently on the market have heterocyclic moieties that contain nitrogen. N-heterocyclic compounds are widely found in nature, have a variety of physiological and pharmacological characteristics, and are building blocks of several molecules that are significant to biology. Vitamins, nucleic acids, medications, antibiotics, colors, and agrochemicals are also included [124]. S-heterocycles continue to be a crucial component of FDA-approved medications and medicinally active substances. Researchers’ focus has switched to other heterocycles, particularly S-heterocycles, as a result of the thorough investigation of nitrogen heterocycles in medicinal chemistry. In contrast to earlier N-heterocycles, numerous attempts have been made to synthesize a range of novel sulfur-containing molecules with a low toxicity profile and significant therapeutic utility. There have been reports in the past five years that highlight the importance and use of sulfur-containing heterocycles in the drug development process, including thiirane, thiophene, thiazole, thiopyran, and thiazolidine [125]. In organic chemistry, the oxygen-containing heterocycles constitute a significant class of molecules. In medical research, these substances—coumarin and oxazole—are utilized as medications [126]. Major therapeutic applications of heterocyclic motifs in AD/PD treatment are shown in Figure 6. Various studies show the potential role of synthetic heterocyclic compounds in AD/PD treatment is summarized in Table 2 and Table 3 providing a summary of in vitro studies showing the inhibitory impact of synthetic heterocyclic compounds on major therapeutic enzymatic targets of AD and PD.

3.1. Nitrogen-Based Heterocyclic Moieties

Nitrogenous heterocyclic moieties as the core nucleus and/or in combination as an antidepressant chemical include quinazoline, pyridine, pyrimidine, pyrrolidine, imidazole, pyrazole, piperidine, oxadiazole, benzimidazole, benzothiazole, piperazine, triazine, purine, benzoxazole, and isoxazole. These compounds with various structural characteristics have been shown to have antidepressant effects through a variety of methods, including COMT inhibitors, MAO inhibitors, and selective serotonin reuptake inhibitors (SSRIs) [127]. Currently, privileged structures such as benzopyrans, arylpiperazines, biphenyls, and indoles are recognized as valuable strategies. Various nitrogen- and oxygen-containing heterocycles, including pyrazoles, hydrazinylthiazoles, xanthones, coumarins, and chromones, have been extensively explored as scaffolds for developing new MAO-B inhibitors. Among these, nitrogen-based derivatives play a pivotal role, with numerous studies emphasizing hydrazines, thiazoles, and indoles as essential scaffolds for the design of novel MAO-B inhibitors [90].
As a nonaromatic heterocyclic nucleus, piperidine has a six-membered ring with one secondary amine group (−NH−) and five methylene groups (−CH2−). Among heterocyclic compounds, piperidines are well known for their several pharmacological uses, which include antipsychotic, antidepressant, and neuroprotective qualities. Many drugs, including clarinex, donezepil, raloxifene, and methylphenidate, have been found to have piperidine as their principal chromophore [128].
Piper nigrum, the plant from which black and white pepper grains are derived, contains the alkaloid piperine. It has been demonstrated that piperine exhibits a variety of activities, including MAO inhibitory action. An MAO-A and -B test was used to screen several compounds associated with piperine and antiepilepsirine, which may have been used in PD [129]. The main alkaloid in a family of structurally related substances present in Lobelia inflata is α-lobeline, also known as lobeline, a lipophilic, nonpyridino, alkaloidal component of Indian tobacco. By interacting with vesicular monoamine transporter 2, lobeline has been shown to suppress DA uptake into synaptic vesicles and promote the reverse transportation of DA from synaptic vesicles. Lobeline reduces behavioral impairments in rats and exhibits protective benefits against MPTP-induced dopaminergic neuron death [130].
As with methylphenidate analogs, its 3,4-dichloroaryl cousin is one of the more powerful benzoylpiperidines. Here, one study shows that these chemicals probably bind similarly at hDAT using homology models. The effectiveness of these hybrids may be influenced by the electronic nature of the substituents, as evidenced by the fact that the 3,4-dichlorobenzoylpiperidine analog of 1a is more potent than its 3,4-dimethyl equivalent. In addition, the 3,4-benz-fused (naphthyl) benzoylpiperidine analog functions at hDAT in the same way as its methylphenidate cousin. Similar to its methylphenidate analog, the naphthyl molecule also operates at the human serotonin transporter (hSERT), albeit in a somewhat different way and with less effectiveness than the other members of the two series. The benzoylpiperidines are a new structural class of hDAT reuptake inhibitors that work similarly to their methylphenidate equivalents [131].
In the Aβ(1-42) rat model of AD, one study examined the potential memory-boosting and antioxidant qualities of the methanolic extract of Piper nigrum L. fruits. In vivo methods were used to investigate the plant extract’s memory-boosting properties. Additionally, the total content of reduced glutathione, malondialdehyde, and protein carbonyl levels, as well as the activities of superoxide dismutase, catalase, and glutathione peroxidase, were used to evaluate the antioxidant activity in the hippocampus. Rats treated with Aβ(1-42) showed an increase in working memory and reference memory errors in the radial arm maze test and a decrease in the percentage of spontaneous alternations in the Y-maze task. The administration of the plant extract demonstrated antioxidant properties and markedly enhanced memory performance. Overall, the results suggest that the plant extract ameliorates Aβ(1-42)-induced spatial memory impairment by attenuating the OS in the rat hippocampus [132].
Both plants and animals naturally contain the heterocyclic aromatic ring system indole. It has electronic and steric properties that are advantageous for attaining bioavailability and pharmacological effects. It is composed of a six-membered benzene ring fused with a five-membered pyrrole ring. Because it may bind with different receptors and produce a wide range of biological functions, indole is known as a flexible pharmacophore. It is therefore regarded as one of the most advantageous frameworks in pharmaceutical chemistry [73]. Zhou L and colleagues developed, produced, and assessed a number of diosgenin–indole compounds for the potential therapy of AD. These compounds were tested for their neuroprotective effects against Aβ, 6-hydroxydopamine (6-OHDA), and hydrogen peroxide (H2O2) damages. The addition of an indole fragment and an electron-donating group at C-5 on the indole ring may have neuroprotective effects, according to preliminary structure–activity relationships. The most promising contender against cellular damage caused by H2O2, 6-OHDA, and Aβ1-42 was compound 5b, according to the results [133]. To treat AD, thirteen new oxathiolanyl, pyrazolyl, and pyrimidinyl indole derivatives were created and produced. The method was as follows: Both the AChE and BChE enzymes were tested in an in vitro enzyme assay. Furthermore, cytotoxicity on a normal cell line and antioxidant assay results were established. Dynamic simulations and molecular docking were used to validate the binding mode in the active regions of both esterases. Studies on toxicity, metabolism, excretion, distribution, and absorption were also conducted in silico. Superior inhibitory action against AChE and BChE was demonstrated by a few of them [134].
As demonstrated by its ranking first among the top five most prevalent five-membered nonaromatic nitrogen heterocycles and its presence in thirty-seven new molecular and/or first-in-class entities approved by the FDA, such as captopril, lincomycin, clemastin, and remoxipride, the pyrrolidine nucleus is one of the preferred scaffolds in pharmaceutical sciences and drug design. Many new pyrrolidine compounds have been created in recent decades for their potential in the CNS, and they have shown promise in treating neurodegenerative illnesses, mental illness, and other CNS problems [135,136]. Numerous pyrrolidine derivatives have been developed as conformationally limited analogs of profadol for analgesic action, and various N-(pyrrolidin-3-yl)-naphthamide analogs bind with high affinity for both the D2 and D3 dopamine receptor subtypes. In addition to these, several pyrrolidine-containing substances have been suggested over time to treat AD [135]. After preparing some new N-benzoylthiourea-pyrrolidine carboxylic acid derivatives with an imidazole moiety and assessing their diverse biological activities, it was discovered that the produced compounds may have inhibited AChE and BChE, the primary targets for AD [137].
Researchers have developed two novel hMAO-B inhibitors using benzimidazole as a scaffold and a primary amide group. The most potent inhibitor inhibits hMAO-B competitively and reversibly. The compound shows a good safety profile, ideal pharmacokinetic properties, and blood–brain barrier permeability. In a PD mouse model, the compound significantly alleviated motor impairment, making it a promising lead compound for further research in PD treatment [138].
Curcumin, a dietary polyphenol, has been studied for its potential treatment against neurodegenerative diseases due to its pharmacological activities and ability to cross the blood–brain barrier. However, its potency and bioavailability are limited due to its physical and metabolic instability. Researchers have attempted to chemically modify curcumin to increase its potency, but one study screened curcumin isoxazole and curcumin pyrazole for their inhibitory potency against α-synuclein aggregation. The results showed that curcumin pyrazole derivatives inhibit α-synuclein aggregation and reduce the associated neurotoxicity [139].
A series of carbazole-based stilbene derivatives were designed by Patel DV et al. [140] to develop multitarget-directed ligands for AD treatment. The compounds were evaluated for anti-AD activities, including cholinesterase inhibition, Aβ aggregation inhibition, and antioxidant and metal chelation properties. The best candidate, (E)-1-(4-(2-(9-ethyl-9H-carbazol-3-yl)vinyl)phenyl)-3-(2-(pyrrolidin-1-yl)ethyl)thiourea), showed good inhibitory activities against AChE and BChE, and significant inhibition of self-mediated Aβ1-42 aggregation.
Table 2. Studies showing the potential therapeutic role of synthesized heterocyclic compounds in the treatment and management of AD and PD.
Table 2. Studies showing the potential therapeutic role of synthesized heterocyclic compounds in the treatment and management of AD and PD.
CompoundsConditionStudy ModelDose and TimeExperimental AssaysOutcomesReference
N-Substituted 2-aryloxymethylpyrrolidinesADHepG2 and SH-SY5Y cells0.1–100 µM for 24 hAChE and BChE inhibition activity; antioxidant activity; molecular modeling; cell viability assayThe compound shows dual BChE/FAAH inhibition activity and a high potential to cross BBB [135]
Diosgenin–indole derivativesADSH-SY5Y cells
ICR mice
0.01–100 mM for 2 hIn vitro: H2O2- and anti-6-OHDA-induced oxidant assay; anti-Ab assay
In vivo: Morris water maze (MWM) test
Effectively improved memory and learning impairments in mice damaged by Aβ[133]
Indole derivativesPDHuman HMC3 microglial cells
MPTP mice
In vitro: 1–10 µM for 8 h+ 20 h
In vivo: 40 mg/kg for 6 weeks (5 times/week)
In vitro: antioxidant assay; In vivo: behavioral tests; neuroinflammation and OS analysesIt reduced MPP+-induced cytotoxicity, reduced NO, IL-1β, IL-6, and TNF-α production, suppressed NLRP3 inflammasome activation, and ameliorated motor deficits, nonmotor depression, and OS in mice[141]
Indole-3-carbinolPDLPS rats25 and 50 mg/kg for 21 daysCytokine assay, NF-κB inhibition assay; balance test, open field test (OFT), and MWM A delay in neurodegeneration of neurons and improvement in motor functions and cognitive function[142]
Benzothiazole and indole derivativesPDM17D-TR/αS-3 K::YFP neuroblastoma cells5 μM to 40 μM for 24 hα-Syn inclusion-forming neuroblastoma cell experiment; tau fibril inhibition assayInhibit α-syn oligomer activity, but not tau oligomers[143]
Indanone derivativesAD and PDPerphenazine (PPZ)-induced (PD)
LPS-induced mice (AD)
PD mice model: 20 mg/kg, p.o.;
AD mice model: 250 µg/kg, i.p
Memory assayImprove cognitive function[144]
1,4-Dihydropyridine derivativesNDDSH-SY5Y; rat hippocampal slices; glia
from cerebral cortex of Sprague Dawley rats
Cell line: 1 μM for 24 h; hippocampal slices: 10 µM for 6 hNeuroprotection studies; anti-inflammatory capacity; GSK-3 inhibitory capacity; voltage-dependent calcium channel blockade assayHave high antioxidant
activity, potent anti-inflammatory capacity,
GSK-3 inhibitory capacity, and VDCC antagonist activity; can cross the BBB
[145]
3,5-Diarylpyrazole analogsADMC65 cell lines; adult female Swiss albino miceCell line: 1–50 µM; mice: 30 mg/kgCholinesterase inhibitory activity and SAR studies; AChE enzyme kinetic assay; behavioral studies (in vivo)Decreased metal-induced Aβ1-42 aggregation;
better spontaneous alternation
score and novel arm entries without influencing the locomotor activity
[146]
Curcumin pyrazole and its derivative
(N-(3-Nitrophenylpyrazole)
PDSHSY5Y neuroblastoma cell line210 µM for 24 h Determination of cytotoxicity; aggregation assaysInhibit α-syn aggregation [139]
CNB-001, a pyrazole derivativePDAdult male C57BL/6 mice24 mg/kg body wt. for 7 daysBehavioral assay; analysis of striatal dopamine and its metabolitesMitigated motor impairments, reduced behavioral impairments, and significantly reduced striatal dopamine and its metabolite levels in mice while also protecting dopaminergic neurons from MPTP toxicity[147]
Pyrazolo[3,4-b]quinoline and benzo[b]pyrazolo[4,3-g][1,8]naphthyridine
derivatives
AD10 nM to 1 µMSH-SY5Y cellsAChE/BuChE inhibitory activity; neuroprotective effectProtect against rotenone/oligomycin A-induced neuronal death; inhibit AChE/BuChE activity in vitro[148]
3,4-Dimethyl coumarin scaffold derivativesADSH-SY5Y cells1 and 10 μM for 24 hAChE and MAO-B inhibitory assayInhibit hAChE and hMAO-B; reduce OS[149]
Dibenzo[1,4,5]thiadiazepineNDDsHuman neuroblastoma cell line SH-SY5Y3 mM for 24 hCholinesterase inhibitory activities, cell viability experiments, neuroprotection studies, cytosolic calcium concentrationShows significant calcium channel modulation activity and is found to be effective in sequestering mitochondrial ROS[150]
N-Heterocyclic amineADIn vitro biochemical assays
and FRDA and neuronal (HT-22) cell line
1 μM for 20 minAntioxidant capacity
and amyloid disaggregation
The compound protects amyloid from copper ions and disaggregated amyloid aggregates, with antioxidant activity observed in both cell lines[151]
3-Amidocoumarin derivatives (coumarins 1–17)NDDsRat cortical neuron culture100 mm for 24 hMAO in vitro inhibition; neuronal survival;
PAMPA
Cross the BBB and thus exert activity
in the CNS; notable neuroprotection from OS
[152]
Ethyl nipecotate (ethyl-piperidine-3-carboxylate)
nipecotic acid
ADWistar rats0.15 mmol/kg for 3.5 hIn vitro lipid peroxidation inhibition; AChE inhibition activity Significant antioxidant potential; GABA reuptake inhibitor; inhibits AChE and reduces rat paw edema[153]
Pyridine-based hybrids linked to the 1,2,3-triazole unitNeurological diseasesOutbred mice50 to 500 mg/kgAnticonvulsant effect; psychotropic properties; OFT, elevated plus maze, forced swimming, and test for learning
and memory
High anticonvulsant and psychotropic properties[154]
1,2,4-Thiadiazolylnitrones and furoxanylnitronesNDDsSH-SY5Y cells 0.05–10 μM for 24 hPAMPA, antioxidant activityCompounds exhibit strong free radical scavenger properties and potential therapeutic applications in preventing cell death from OS and damage[155]
Rosiglitazone ADK670N/M671L-transgenic mice overexpressing human hAPP 5 mg/kg for 4 weeksObject recognition and MWM tests; Aβ plaque deposition assay (ELISA)Ameliorates memory deficits; rosiglitazone shown to decrease brain Aβ levels and Aβ plaque deposition; reduces p-tau aggregates[156]
Riluzole ADAβ25-35-induced rat10 mg/kg/day p.o. MWM tests; AChE activity and OS marker assayImproves spatial memory, retention, and recall in MWM and passive avoidance tasks, but is not neutralized by muscarinic or nicotinic receptor antagonists[157]
In recent experimental work, two novel derivative series acting as hMAO-B inhibitors were developed and evaluated by the researchers. The primary amide group, which is known to be a crucial pharmacophore in the subsequent activity screening and reversible mode of action, has been carefully added to these series, which use benzimidazole as a scaffold. With an IC50 value of 67.3 nM in vitro, 16d is the most effective hMAO-B inhibitor among these substances. Furthermore, 16d demonstrated a favorable safety profile in mouse tests for acute toxicity and cellular damage. Additionally, it demonstrated optimal pharmacokinetic characteristics and blood–brain barrier permeability in vivo, which are crucial requirements for medications that target the CNS. 16d dramatically reduced motor impairment, particularly muscle relaxation and motor coordination, in the MPTP-induced PD mouse model. Being a lead compound, 16d thus has instructional value for further research on its use in the treatment of PD [138].
Based on the location of individual nitrogen atoms arising from various imidazole and pyridine ring annulations, imidazopyridines are classified into four categories. According to the proper nomenclature, all of the scaffolds are classified as imidazopyridines, even if they vary in their structures or even in the number of nitrogen atoms (imidazo[1,2-a]pyridine I and imidazo[4,5-c]pyridine IV). Derivatives of imidazopyridines interact with several biological targets in the CNS, making them a flexible structural template for the development of new CNS modulators [158].
Table 3. In vitro studies showing the inhibitory impact of synthetic heterocyclic compounds on major therapeutic enzymatic targets of AD and PD.
Table 3. In vitro studies showing the inhibitory impact of synthetic heterocyclic compounds on major therapeutic enzymatic targets of AD and PD.
CompoundsConditionExperimental AssaysOutcomesReference
6-Amino-substituted imidazo[1,2-b]
pyridazines
NDDsDPPH radical scavenging activity; AChE inhibition assay; molecular docking Antioxidative/antiparkinsonian agents for important
metabolic functions
[159]
4-(Benzylideneamino)-
and 4-(benzylamino)-benzenesulfonamide derivatives
ADAChE activity determination, in vitro inhibition studies, ADMET analysisPotential inhibitor properties for AChE[160]
Thiazolyl-pyrazoline derivatives (3a-k)ADAChE activity assay, CA activity assay, AChE and CA kinetic analysis Inhibit AChE and hCA activity[161]
Isoindole-1,3-dione-substituted sulfonamides AChE enzyme activity; molecular docking studyInhibit AChE and hCA activity[162]
Isoindolines/isoindoline-1,3-dionesADAnticholinesterase activity assay;
molecular docking
AChE inhibitors [163]
Imidazo
[2,1-B][1,3,4] thiadiazole
ADBACE1 enzymatic assay; inhibitory activities against AChE and BChESuperior BACE1 inhibitory activity; potential inhibitory activity against cholinesterase
(AChE and BChE)
[164]
1-Hydroxy-2(1H)-pyridinone-based chelatorsPDModeling methodsPotential to inhibit COMT[165]
N-substituted pyrazole-derived α-aminophosphonatesADInhibition assay on cholinesterase; evaluation of the cytotoxic activity; antioxidant activity assayBetter AChE inhibitory activity; did not show any cytotoxicity and have promising antioxidant activities against DPPH and H2O2
scavenging
[166]
Nitrocatechol derivatives of chalcone
PDMAO-A, MAO-B, and COMT inhibition assayPotent inhibitors of MAO and COMT[167]
1,3-
Oxazole analogs
ADAChE and BChE inhibition activityAbility
to inhibit AChE and BChE
[168]
Nitrogen-based novel heterocyclic
compounds
ADAChE and α-glycosidase
enzymes were evaluated
Potentially inhibit AChE and α-glycosidase
[124]
Heterocyclic amines (F3S4-m, F2S4-m, and F2S4-p)ADBACE1 and AChE inhibition activity and Aβ
oligomerization assay
Inhibit Aβ1–42 aggregation and AChE and BACE1 enzyme activities[169]
6-Benzothiazolyl urea, thiourea, and guanidine derivativesADABAD’s enzymatic activity Potent inhibitors of ABAD/17β-HSD10 and potential drugs for AD treatment[170]
As a heterocyclic aromatic molecule with six members, pyrimidine has nitrogen atoms at positions one and three. This moiety has been used in medicinal chemistry in a variety of ways in recent years. Researchers are drawn to pyrimidine derivatives because of their therapeutic value and adaptable scaffolds [171]. Marketed medications like piribedil, which is used to treat PD, demonstrate their promise for use in NDD medicine. Additionally, a number of molecules containing pyrimidines have been identified as anti-neurodegenerative medicines in preclinical and clinical trials. By inhibiting different enzymes and targets involved in NDDs, the drugs’ activity is explained. Pyrimidine analogs show promise as anti-neurodegenerative drugs. These compounds’ biological evidence supports their potential as multitarget enzyme inhibitors, demonstrating their pharmacological potential by inhibiting targets like ChE, Aβ, MAO-B, KMO, SOD1, Kv, and CX3CR1 that are implicated in several neurodegenerative disorders. Therefore, it is evident from the above that molecules having pyrimidine scaffolds have several uses in the development of drugs to treat NDDs [172].
Several mental illnesses, including anxiety and depression, have etiological relationships with monoamine levels. In theory, a suitable way to evaluate the antidepressant qualities of novel medication candidates is to block MAO [173].
From apixaban, an anticoagulant used to treat and prevent blood clots and stroke, to bixafen, a pyrazole-carboxamide fungicide used to control diseases of rapeseed and cereal plants, the remarkable prevalence of pyrazole scaffolds in a diverse array of bioactive molecules has prompted medicinal and organic chemists to investigate novel approaches in creating pyrazole-containing compounds for various applications [174]. Recently, a study investigated the neuroprotective properties of CNB-001, a novel pyrazole derivative of curcumin and cyclohexyl bisphenol A, in PD. The study found that CNB-001 significantly reduces motor impairments, lowers dopamine levels, and upregulates inflammatory and apoptotic markers. However, co-treatment with CNB-001 attenuated these effects. The results suggest CNB-001’s potential as a therapeutic candidate for PD treatment [147].
Turkan and colleagues synthesized and evaluated the inhibitory effects of pyrazole derivatives (1–8) against the human carbonic anhydrase isoenzymes I and II, as well as the AChE, BChE, and α-glycosidase metabolic enzymes. Their studies revealed that a series of substituted pyrazol-4-yl-diazene derivatives effectively inhibit α-glycosidase, cytosolic hCA I and II, BChE, and AChE. Recently, the inhibition of these metabolic enzymes has emerged as a promising avenue for pharmacologic intervention in a variety of conditions, including NDDs [175].
The pyrazoline (five-membered) heterocyclic ring compound is of great importance pharmaceutically, and its derivatives are of research interest, currently being explored for drug development. Because of their high stability, medicinal chemists have been motivated to experiment with the structure of the ring in a wide range of ways to produce a wide range of pharmacological actions. Pyrazoline rings are found in the architecture of several pharmaceutical products that are sold commercially. The ability of pyrazolines to cure neurodegenerative illnesses is widely established. The NDDs that impact large populations worldwide include psychiatric diseases, AD, and PD [176].
Hitge et al.’s [177] study proposes molecular docking to study potential binding modes and interactions with COMT. The study identifies dual MAO-B/COMT inhibitors by synthesizing chalcone derivatives and converting them to pyrazoline derivatives. The pyrazoline derivatives were found to be more potent than chalcones in inhibiting human MAO and rat COMT. However, these derivatives were weak MAO inhibitors [177].
The goal of another study was to find new dual inhibitors of COMT and MAO, which are involved in the metabolism of dopamine and L-dopa. Chalcone nitrocatechol compounds were created for this aim and assessed as COMT and MAO inhibitors. While nitrocatechol derivatives, such as tolcapone and entacapone, are therapeutically employed COMT inhibitors, the chalcone class of drugs is well known for its strong inhibition of MAO-B. All of the compounds are highly effective in vitro inhibitors of rat liver COMT, according to the data. The most effective inhibitors for the in vitro inhibition of human MAO-B are chalcones, which are less effective as MAO inhibitors. This study suggests a generic method for boosting MAO-B inhibition while maintaining the strong COMT inhibitory activity of this class. It also demonstrates that nitrocatechol derivatives of chalcone may function as COMT and MAO-B inhibitors [167].
Abdelgawad and colleagues produced halogenated chalcones and assessed how well they inhibited MAOs. Compared to MAO-A, all of the synthesized chalcones exhibited stronger and greater inhibitory efficacy against MAO-B. Interestingly, CHB3 and CHF3 inhibited MAO-B more strongly than the reference compounds pargyline and lazabemide. Both compounds inhibited MAO-B reversibly and competitively. According to the findings, CHB3 may be a promising treatment for neurological conditions like PD [178]. Because of its dual inhibitory effects on MAOs and ChEs, the oxygen-containing bicyclic molecule coumarin is frequently employed to create medicines for NDDs [179]. Coumarin–triazole trihybrids with phenylpiperazine, triazole, carbazole, donepezil pharamcophore, and arylisoxazole moieties exhibit good AChE enzyme inhibitory action, representing potential neuroprotective agents for AD treatment [180].

3.2. Sulfur-Based Heterocycle Moieties

Many studies on S-containing compounds in the context of AD have been carried out because of the inherent antioxidant potential of S-based heterocycles, and some are presently undergoing clinical trials. Thiazole, which has the formula C3H3NS, is a heterocyclic chemical molecule having a five-membered molecular ring structure. Because of the acidic proton at position C-2, the thiazole ring is extremely reactive and has become a key synthon to produce a range of NCEs (new chemical entities). Numerous new compounds with a broad range of pharmacological actions, including antioxidant, antibacterial, antifungal, antitubercular, diuretic, anti-inflammatory, anticancer, and antipsychotic properties, have been produced by diverse modifications of the thiazole ring at different locations [181]. A diverse range of therapeutic medicines targeting various CNS targets have been developed using thiazoles. Several thiazole derivatives are currently undergoing clinical studies, and thiazole-containing therapeutic molecules are currently being used to treat a variety of CNS illnesses. Thiazole and its analogs have been the subject of numerous studies, which have demonstrated their effectiveness in treating several CNS disorders in both rodent and primate models [182].
CA inhibitors are based on sulfonamide groups; their primary disadvantage is that they nonspecifically block all CA isoforms, which might result in undesirable side effects. Askin et al. (2021) [183] reported an investigation of novel derivatives of imidazole and thiadiazole that do not possess the zinc-binding sulfonamide group for inhibiting human CA I and II isoforms and AChE. Imidazo[2,1-b][1,3,4]thiadiazoles showed low nanomolar inhibitory activity against hCA I, hCA II, and AChE. Compound 9b inhibited hCA I 18-fold more than acetazolamide, while compound 10a selectively targeted hCA II. These compounds could be potential lead compounds for further design [183].
The FDA-approved adenosine A receptor antagonist (anti-PD medication) istradefylline is effective. However, it has been shown that when istradefylline is exposed to direct light or an indoor atmosphere in a diluted solution, its double bond readily transforms into a cis structure. The compound’s series (12 compounds) was created by keeping the xanthine skeleton of istradefylline the same while substituting thiazole orbenzothiazole and other physiologically active heterocyclic compounds for the trans-double bond in order to find more stable adenosine A receptor antagonists with comparable pharmacological efficacy to istradefylline. These compounds were created by a multi-step procedure, and their capacity to prevent the generation of cAMP in cells that overexpress AAR was successfully verified using several characterization approaches. In contrast to istradefylline, the thiazole derivative of istradefylline demonstrated notable activity. Furthermore, benzothiazole derivatives and thiazole compounds with a higher rate of inhibition were molecularly docked and compared to istradefylline. Therefore, this research may serve as a foundation for the logical development of adenosine, a potent receptor antagonist [184].
For more than a century, methylene blue (MB), a tricyclic phenothiazine often referred to as methylthionine hydrochloride, has been used for a variety of medical purposes, including targeting specific cellular targets. MB has been shown to prevent tau aggregation in vitro and has beneficial effects on AD and memory improvement [185]. According to Akour E and colleagues’ research, the interaction between the reduction in and oxidation of tau’s native cysteine residues forms the basis of the process that inhibits tau aggregation. Because MB and its metabolites keep tau in a monomeric disordered shape, they stop filaments and their harmful precursors from forming [186].
Derivatives of fused thiophene are important heterocycles in medicinal chemistry, with fascinating uses in many different domains. They have a wide variety of biological activities including anti-inflammatory, analgesic, and antidepressant activities. Fused thiophene is a potent therapeutic molecule for AD treatment due to its targeting of Aβ aggregation and enzymes including cholinesterases, MAO, and glycogen synthase kinase-3 [187].
In their work, González-Muñoz and colleagues described a novel family of dibenzo[1,4,5]thiadiazepines (1-12) that exhibited an intriguing pharmacological profile in vitro, including antioxidant and neuroprotective qualities, in addition to blocking cytosolic calcium entry. Most of them were anticipated to be CNS-permeable drugs and did not exhibit any deleterious effects. When substances were treated with human neuroblastoma cells 24 h before the addition of toxic stimuli, the cells demonstrated good neuroprotective capabilities against mitochondrial OS, often reaching near-complete protection (>90%). These numbers were lower under co-incubation settings, but certain compounds still exhibited an intriguing degree of neuroprotection—above 50%. Four chosen compounds were discovered to be efficient antioxidants by containing mitochondrial ROS. Additionally, the molecules demonstrated a surprising ability to modulate calcium channels. The fact that dibenzo[1,4,5]thiadiazepine is a little-known structure that is easy to synthesize and has few reported derivatives adds to the interest in these compounds and opens up a new and expansive field of study in medicinal chemistry [150].
Porcal W et al. (2008) synthesized the new compounds 1,2,4-thiadiazolylnitrones and furoxanylnitrones as neuroprotective agents for human neuroblastoma cells. They inhibited oxidative damage and death induced by hydrogen peroxide exposure, demonstrating their potential as neuroprotective agents [155].

3.3. Oxygen-Based Heterocycle Moieties

Oxygen-based heterocycles form the core structure of many biologically active molecules as well as U.S. FDA-approved drugs. Moreover, they possess a broad range of biological activities. Often found in remarkably bioactive compounds and a wide range of natural products, oxazole is a biologically active scaffold on which pharmacophores are built to produce powerful, selective medicines. Using nineteen hybrid 1,3-oxazole-based benzoxazole analogs, Hussain and his team assessed their capacity to inhibit AChE and BChE. All analogs showed varied levels of AChE and BChE inhibition in comparison to the conventional donepezil [168]. Opicapone inhibits COMT and is commonly used to manage the symptoms of PD [188].
Flavone derivatives and 1,2,4-oxadiazole have strong anti-inflammatory and antioxidant properties. Based on the impressive anti-inflammatory and antioxidant properties of the 1,2,4-oxadiazole and flavonoid pharmacophores, one researcher created and synthesized a novel series of 3-methyl-8-(3-methyl-1,2,4-oxadiazol-5-yl)-2-phenyl-4H-chromen-4-one derivatives by pharmacophore combination to find highly effective drugs for the treatment of Parkinson’s disease. They then assessed the anti-inflammatory and antioxidant properties of these compounds for the treatment of PD. Their inhibitory actions against ROS and NO release in LPS-induced BV2 microglial cells were used to conduct a preliminary SAR study. The best compound, Flo8, showed the strongest anti-inflammatory and antioxidant properties. By blocking inflammatory and apoptotic signaling pathways, Flo8 prevented neuronal apoptosis, according to both in vitro and in vivo data. Additionally, in vivo research revealed that the chemical Flo8 raised serum dopamine levels and improved behavioral and motor impairments in MPTP-induced PD model mice. All things considered, this study showed that the chemical Flo8 may be a potential treatment for PD [189].
Chalcones are seen as attractive possibilities for the treatment of NDDs, including PD, because of their straightforward structure and advantageous biological characteristics. Chalcones are characterized as promising α-syn imaging probes, enzyme inhibitors (MAO-B, COMT, AChE), antagonists of adenosine A1 and/or A2A receptors, and agents with anti-neuroinflammatory properties (iNOS suppression or Nrf2 signaling activation) [190].
In a recent study by Ceyhun I et al. [191], twelve novel chalcones (2a–l) were synthesized, and their neuroprotective role was studied by investigating the AChE and BchE inhibitory potentials. The synthetic compounds showed notable action against AChE. Additionally, docking modeling showed that these substances interacted with the enzyme active site in a manner akin to donepezil, suggesting they might be powerful medications for the treatment of AD [191].
Sever B and colleagues designed new thiazolyl-pyrazoline (3a-k) compounds to inhibit AChE and hCA activity. The in vitro and in silico results revealed that among 3a to 3k, the most promising derivatives were 3a, 3f, and 3d, with significant effects on AChE, hCA I, and hCA II [161]. Two tolcapone analogs, containing a 1H-pyridinone replacement, have been synthesized by Bergin JCJ et al. (2022) to inhibit COMT and protect neurons from oxidative damage. Measurements of pKa, stability constants, and in silico modeling suggest that these compounds are promising candidates for further evaluation, potentially reducing dopamine in the brain and protecting neurons in PD [165].
Various synthetic compounds currently under clinical trials for AD and PD treatment are listed in Table 4.

4. Current Challenges and Future Prospects

L-dopa, dopamine agonists, and MAO-B and COMT inhibitors are the only pharmacological classes that are approved for the treatment of motor-related symptoms of PD, despite all the new research. These treatments primarily act on the dopaminergic neuron system. Although they are also accessible, anticholinergic medications and glutamate antagonists are not frequently utilized in daily practice. Since there is currently no successful therapy approach, alternative approaches need to be looked into. One area of research that has always proven essential to enhancing human health is the hunt for new medicinal possibilities. The largest problem the scientific world is still confronting is drug discovery research. A hybrid approach using multitarget-directed ligands (MTDLs) has been widely used to develop therapeutic agents for AD and PD.
Heterocycles are chemical compounds with at least one element other than carbon in their ring structure and have long been acknowledged as recurring scaffolds in medicinal chemistry because of their importance in drug discovery and development. They are used as appealing building blocks in drug design for neuroprotective drugs and thus, because of their structural diversity and adaptability, open the door to novel treatments and better patient results [192].
In addition to the intricacy of the research, the difficulties encountered in the identification, development, and approval of novel chemical entities are frequently costly and time-consuming [193]. Traditional trial-and-error experiments, which are expensive, time-consuming, and unpredictable, are still used in the development of current formulations. A new field called “computational pharmaceutics” has emerged in the last ten years as a result of the exponential growth of computing power and algorithms. This field combines big data, artificial intelligence, and multi-scale modeling techniques with pharmaceutics, and it has the potential to revolutionize drug delivery. Although molecular modeling is a large subject, the three most popular computer modeling components—molecular docking, MD simulation, and ADMET modeling—have proven essential to the simple identification of leads for experimental in vitro and in vivo testing [193,194,195,196].

Author Contributions

Conceptualization, N.P.; writing—original draft preparation, N.P.; writing—review and editing, N.P.; supervision, M.S. 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 conflicts of interest.

References

  1. Sweeney, P.; Park, H.; Baumann, M.; Dunlop, J.; Frydman, J.; Kopito, R.; McCampbell, A.; Leblanc, G.; Venkateswaran, A.; Nurmi, A.; et al. Protein misfolding in neurodegenerative diseases: Implications and strategies. Transl. Neurodegener. 2017, 6, 6. [Google Scholar] [CrossRef]
  2. Khanam, H.; Ali, A.; Asif, M.; Shamsuzzaman. Neurodegenerative diseases linked to misfolded proteins and their therapeutic approaches: A review. Eur. J. Med. Chem. 2016, 124, 1121–1141. [Google Scholar] [CrossRef] [PubMed]
  3. Puranik, N.; Yadav, D.; Song, M. Advancements in the Application of Nanomedicine in Alzheimer’s Disease: A Therapeutic Perspective. Int. J. Mol. Sci. 2023, 24, 14044. [Google Scholar] [CrossRef] [PubMed]
  4. 2023 Alzheimer’s disease facts and figures. Alzheimer’s Dement. 2023, 19, 1598–1695. [CrossRef] [PubMed]
  5. Zhang, J.; Zhang, Y.; Wang, J.; Xia, Y.; Zhang, J.; Chen, L. Recent advances in Alzheimer’s disease: Mechanisms, clinical trials and new drug development strategies. Signal Transduct. Target. Ther. 2024, 9, 211. [Google Scholar] [CrossRef]
  6. Puranik, N.; Song, M. Insights into the Role of microRNAs as Clinical Tools for Diagnosis, Prognosis, and as Therapeutic Targets in Alzheimer’s Disease. Int. J. Mol. Sci. 2024, 25, 9936. [Google Scholar] [CrossRef]
  7. Gómez-Benito, M.; Granado, N.; García-Sanz, P.; Michel, A.; Dumoulin, M.; Moratalla, R. Modeling Parkinson’s Disease With the Alpha-Synuclein Protein. Front. Pharmacol. 2020, 11, 356. [Google Scholar] [CrossRef]
  8. Morris, H.R.; Spillantini, M.G.; Sue, C.M.; Williams-Gray, C.H. The pathogenesis of Parkinson’s disease. Lancet 2024, 403, 293–304. [Google Scholar] [CrossRef]
  9. Sayyaed, A.; Saraswat, N.; Vyawahare, N.; Kulkarni, A. A detailed review of pathophysiology, epidemiology, cellular and molecular pathways involved in the development and prognosis of Parkinson’s disease with insights into screening models. Bull. Natl. Res. Cent. 2023, 47, 70. [Google Scholar] [CrossRef]
  10. Rademacher, K.; Nakamura, K. Role of dopamine neuron activity in Parkinson’s disease pathophysiology. Exp. Neurol. 2024, 373, 114645. [Google Scholar] [CrossRef]
  11. Chakraborty, A.; Diwan, A. Alzheimer and It’s Possible Therapy: A Review. J. Exp. Neurol. 2020, 1, 115–122. [Google Scholar] [CrossRef]
  12. Gouda, N.A.; Elkamhawy, A.; Cho, J. Emerging Therapeutic Strategies for Parkinson’s Disease and Future Prospects: A 2021 Update. Biomedicines 2022, 10, 371. [Google Scholar] [CrossRef] [PubMed]
  13. Cummings, J.; Zhou, Y.; Lee, G.; Zhong, K.; Fonseca, J.; Cheng, F. Alzheimer’s disease drug development pipeline: 2023. Alzheimer’s Dement. Transl. Res. Clin. Interv. 2023, 9, e12385. [Google Scholar] [CrossRef]
  14. McFarthing, K.; Buff, S.; Rafaloff, G.; Dominey, T.; Wyse, R.K.; Stott, S.R.W. Parkinson’s Disease Drug Therapies in the Clinical Trial Pipeline: 2020. J. Park. Dis. 2020, 10, 757–774. [Google Scholar] [CrossRef] [PubMed]
  15. Srivastava, S.; Ahmad, R.; Khare, S.K. Alzheimer’s disease and its treatment by different approaches: A review. Eur. J. Med. Chem. 2021, 216, 113320. [Google Scholar] [CrossRef] [PubMed]
  16. Singh, M.K.; Shin, Y.; Ju, S.; Han, S.; Kim, S.S.; Kang, I. Comprehensive Overview of Alzheimer’s Disease: Etiological Insights and Degradation Strategies. Int. J. Mol. Sci. 2024, 25, 6901. [Google Scholar] [CrossRef]
  17. Wu, M.; Zhang, M.; Yin, X.; Chen, K.; Hu, Z.; Zhou, Q.; Cao, X.; Chen, Z.; Liu, D. The role of pathological tau in synaptic dysfunction in Alzheimer’s diseases. Transl. Neurodegener. 2021, 10, 45. [Google Scholar] [CrossRef]
  18. Bagga, S.; Kumar, M. Current Status of Alzheimer’s Disease and Pathological Mechanisms Investigating the Therapeutic Molecular Targets. Curr. Mol. Med. 2023, 23, 492–508. [Google Scholar] [CrossRef]
  19. Ju, Y.; Tam, K. Pathological mechanisms and therapeutic strategies for Alzheimer’s disease. Neural Regen. Res. 2022, 17, 543–549. [Google Scholar] [CrossRef]
  20. Cheong, S.L.; Tiew, J.K.; Fong, Y.H.; Leong, H.W.; Chan, Y.M.; Chan, Z.L.; Kong, E.W.J. Current Pharmacotherapy and Multi-Target Approaches for Alzheimer’s Disease. Pharmaceuticals 2022, 15, 1560. [Google Scholar] [CrossRef]
  21. Hynd, M.R.; Scott, H.L.; Dodd, P.R. Glutamate-mediated excitotoxicity and neurodegeneration in Alzheimer’s disease. Neurochem. Int. 2004, 45, 583–595. [Google Scholar] [CrossRef] [PubMed]
  22. Rajah Kumaran, K.; Yunusa, S.; Perimal, E.; Wahab, H.; Müller, C.P.; Hassan, Z. Insights into the Pathophysiology of Alzheimer’s Disease and Potential Therapeutic Targets: A Current Perspective. J. Alzheimer’s Dis. 2023, 91, 507–530. [Google Scholar] [CrossRef] [PubMed]
  23. Conway, M.E.; Conway, M.E. Alzheimer’s disease: Targeting the glutamatergic system. Biogerontology 2020, 21, 257–274. [Google Scholar] [CrossRef] [PubMed]
  24. Burns, S.; Selman, A.; Sehar, U.; Rawat, P.; Reddy, A.P.; Reddy, P.H. Therapeutics of Alzheimer’s Disease: Recent Developments. Antioxidants 2022, 11, 2402. [Google Scholar] [CrossRef] [PubMed]
  25. Xiao, D.; Zhang, C. Current therapeutics for Alzheimer’s disease and clinical trials. Open Explor. 2024, 3, 255–271. [Google Scholar] [CrossRef]
  26. Senkevich, K.; Rudakou, U.; Gan-Or, Z. New therapeutic approaches to Parkinson’s disease targeting GBA, LRRK2 and Parkin. Neuropharmacology 2022, 202, 108822. [Google Scholar] [CrossRef]
  27. Caramiello, A.M.; Pirota, V. Novel Therapeutic Horizons: SNCA Targeting in Parkinson’s Disease. Biomolecules 2024, 14, 949. [Google Scholar] [CrossRef]
  28. Polinski, N.K.; Martinez, T.N.; Ramboz, S.; Sasner, M.; Herberth, M.; Switzer, R.; Ahmad, S.O.; Pelligrino, L.J.; Clark, S.W.; Marcus, J.N.; et al. The GBA1 D409V mutation exacerbates synuclein pathology to differing extents in two alpha-synuclein models. Dis. Model. Mech. 2022, 15, dmm049192. [Google Scholar] [CrossRef]
  29. Tang, T.; Jian, B.; Liu, Z. Transmembrane Protein 175, a Lysosomal Ion Channel Related to Parkinson’s Disease. Biomolecules 2023, 13, 802. [Google Scholar] [CrossRef]
  30. Pickrell, A.M.; Huang, C.H.; Kennedy, S.R.; Ordureau, A.; Sideris, D.P.; Hoekstra, J.G.; Harper, J.W.; Youle, R.J. Endogenous Parkin Preserves Dopaminergic Substantia Nigral Neurons following Mitochondrial DNA Mutagenic Stress. Neuron 2015, 87, 371–381. [Google Scholar] [CrossRef]
  31. Vizziello, M.; Borellini, L.; Franco, G.; Ardolino, G. Disruption of Mitochondrial Homeostasis: The Role of PINK1 in Parkinson’s Disease. Cells 2021, 10, 3022. [Google Scholar] [CrossRef] [PubMed]
  32. Guo, Y.; Li, S.; Zeng, L.-H.; Tan, J. Tau-targeting therapy in Alzheimer’s disease: Critical advances and future opportunities. Ageing Neurodegener. Dis. 2022, 2, 11. [Google Scholar] [CrossRef]
  33. Congdon, E.E.; Ji, C.; Tetlow, A.M.; Jiang, Y.; Sigurdsson, E.M. Tau-targeting therapies for Alzheimer disease: Current status and future directions. Nat. Rev. Neurol. 2023, 19, 715–736. [Google Scholar] [CrossRef] [PubMed]
  34. Ye, J.; Wan, H.; Chen, S.; Liu, G.P. Targeting tau in Alzheimer’s disease: From mechanisms to clinical therapy. Neural Regen. Res. 2024, 19, 1489–1498. [Google Scholar] [CrossRef] [PubMed]
  35. VandeVrede, L.; Boxer, A.L.; Polydoro, M. Targeting tau: Clinical trials and novel therapeutic approaches. Neurosci. Lett. 2020, 731, 134919. [Google Scholar] [CrossRef]
  36. Rao, C.V.; Asch, A.S.; Carr, D.J.J.; Yamada, H.Y. “Amyloid-beta accumulation cycle” as a prevention and/or therapy target for Alzheimer’s disease. Aging Cell 2020, 19, e13109. [Google Scholar] [CrossRef]
  37. Karran, E.; De Strooper, B. The amyloid hypothesis in Alzheimer disease: New insights from new therapeutics. Nat. Rev. Drug Discov. 2022, 21, 306–318. [Google Scholar] [CrossRef]
  38. Jeremic, D.; Jiménez-Díaz, L.; Navarro-López, J.D. Past, present and future of therapeutic strategies against amyloid-β peptides in Alzheimer’s disease: A systematic review. Ageing Res. Rev. 2021, 72, 101496. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Chen, H.; Li, R.; Sterling, K.; Song, W. Amyloid β-based therapy for Alzheimer’s disease: Challenges, successes and future. Signal Transduct. Target. Ther. 2023, 8, 248. [Google Scholar] [CrossRef]
  40. Jasutkar, H.G.; Oh, S.E.; Mouradian, M.M. Therapeutics in the Pipeline Targeting α-Synuclein for Parkinson’s Disease. Pharmacol. Rev. 2022, 74, 207–237. [Google Scholar] [CrossRef]
  41. Teil, M.; Arotcarena, M.L.; Faggiani, E.; Laferriere, F.; Bezard, E.; Dehay, B. Targeting α-Synuclein for PD Therapeutics: A Pursuit on All Fronts. Biomolecules 2020, 10, 391. [Google Scholar] [CrossRef]
  42. Walczak-Nowicka, Ł.J.; Herbet, M. Acetylcholinesterase Inhibitors in the Treatment of Neurodegenerative Diseases and the Role of Acetylcholinesterase in their Pathogenesis. Int. J. Mol. Sci. 2021, 22, 9290. [Google Scholar] [CrossRef] [PubMed]
  43. Zhao, X.; Hu, Q.; Wang, X.; Li, C.; Chen, X.; Zhao, D.; Qiu, Y.; Xu, H.; Wang, J.; Ren, L.; et al. Dual-target inhibitors based on acetylcholinesterase: Novel agents for Alzheimer’s disease. Eur. J. Med. Chem. 2024, 279, 116810. [Google Scholar] [CrossRef] [PubMed]
  44. Benfante, R.; Di Lascio, S.; Cardani, S.; Fornasari, D. Acetylcholinesterase inhibitors targeting the cholinergic anti-inflammatory pathway: A new therapeutic perspective in aging-related disorders. Aging Clin. Exp. Res. 2019, 33, 823–834. [Google Scholar] [CrossRef] [PubMed]
  45. Akıncıoğlu, H.; Gülçin, İ. Potent Acetylcholinesterase Inhibitors: Potential Drugs for Alzheimer’s Disease. Mini-Rev. Med. Chem. 2020, 20, 703–715. [Google Scholar] [CrossRef]
  46. Marucci, G.; Buccioni, M.; Ben, D.D.; Lambertucci, C.; Volpini, R.; Amenta, F. Efficacy of acetylcholinesterase inhibitors in Alzheimer’s disease. Neuropharmacology 2021, 190, 108352. [Google Scholar] [CrossRef]
  47. Nordberg, A.; Ballard, C.; Bullock, R.; Darreh-Shori, T.; Somogyi, M. A Review of Butyrylcholinesterase as a Therapeutic Target in the Treatment of Alzheimer’s Disease. Prim. Care Companion CNS Disord. 2013, 15, PCC.12r01412. [Google Scholar] [CrossRef]
  48. Ha, Z.Y.; Mathew, S.; Yeong, K.Y. Butyrylcholinesterase: A Multifaceted Pharmacological Target and Tool. Curr. Protein Pept. Sci. 2019, 21, 99–109. [Google Scholar] [CrossRef]
  49. Fu, L.; Liu, B. Role of receptor tyrosine kinases in neurodegenerative disorders. In Receptor Tyrosine Kinases in Neurodegenerative and Psychiatric Disorders; Elsevier: Amsterdam, The Netherlands, 2023; pp. 279–299. [Google Scholar] [CrossRef]
  50. Turkistani, A.; Al-kuraishy, H.M.; Al-Gareeb, A.I.; Albuhadily, A.K.; Alexiou, A.; Papadakis, M.; Elfiky, M.M.; Saad, H.M.; Batiha, G.E. Therapeutic Potential Effect of Glycogen Synthase Kinase 3 Beta (GSK-3β) Inhibitors in Parkinson Disease: Exploring an Overlooked Avenue. Mol. Neurobiol. 2024, 61, 7092–7108. [Google Scholar] [CrossRef]
  51. Lauretti, E.; Dincer, O.; Praticò, D. Glycogen synthase kinase-3 signaling in Alzheimer’s disease. Biochim. Biophys. Acta Mol. Cell Res. 2020, 1867, 118664. [Google Scholar] [CrossRef]
  52. Yu, H.; Xiong, M.; Zhang, Z. The role of glycogen synthase kinase 3 beta in neurodegenerative diseases. Front. Mol. Neurosci. 2023, 16, 1209703. [Google Scholar] [CrossRef] [PubMed]
  53. Swarup, G.; Kanik, P.; Shekhar, V.; Gupta, S. β- and γ-secretases as therapeutic targets for Alzheimer’s disease. In Targeted Therapy for the Central Nervous System; Elsevier: Amsterdam, The Netherlands, 2025; pp. 239–263. [Google Scholar] [CrossRef]
  54. Wolfe, M.S. γ-Secretase: Once and future drug target for Alzheimer’s disease. Expert. Opin. Drug Discov. 2024, 19, 5–8. [Google Scholar] [CrossRef] [PubMed]
  55. Özdemir, Z.; Alagöz, M.A.; Bahçecioğlu, Ö.F.; Gök, S. Monoamine Oxidase-B (MAO-B) Inhibitors in the Treatment of Alzheimer’s and Parkinson’s Disease. Curr. Med. Chem. 2021, 28, 6045–6065. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, W.; Ross, P.J.; Ellis, J.; Salter, M.W. Targeting NMDA receptors in neuropsychiatric disorders by drug screening on human neurons derived from pluripotent stem cells. Transl. Psychiatry 2022, 12, 243. [Google Scholar] [CrossRef]
  57. Bading, H. Therapeutic targeting of the pathological triad of extrasynaptic NMDA receptor signaling in neurodegenerations. J. Exp. Med. 2017, 214, 569. [Google Scholar] [CrossRef]
  58. Carles, A.; Freyssin, A.; Perin-Dureau, F.; Rubinstenn, G.; Maurice, T. Targeting N-Methyl-d-Aspartate Receptors in Neurodegenerative Diseases. Int. J. Mol. Sci. 2024, 25, 3733. [Google Scholar] [CrossRef]
  59. Jiang, S.; Li, Y.; Zhang, C.; Zhao, Y.; Bu, G.; Xu, H.; Zhang, Y.W. M1 muscarinic acetylcholine receptor in Alzheimer’s disease. Neurosci. Bull. 2014, 30, 295–307. [Google Scholar] [CrossRef]
  60. Scarpa, M.; Hesse, S.; Bradley, S.J. M1 muscarinic acetylcholine receptors: A therapeutic strategy for symptomatic and disease-modifying effects in Alzheimer’s disease? Adv. Pharmacol. 2020, 88, 277–310. [Google Scholar] [CrossRef]
  61. Tozzi, A.; Tantucci, M.; Marchi, S.; Mazzocchetti, P.; Morari, M.; Pinton, P.; Mancini, A.; Calabresi, P. Dopamine D2 receptor-mediated neuroprotection in a G2019S Lrrk2 genetic model of Parkinson’s disease. Cell Death Dis. 2018, 9, 204. [Google Scholar] [CrossRef]
  62. Rangel-Barajas, C.; Coronel, I.; Florán, B. Dopamine Receptors and Neurodegeneration. Aging Dis. 2015, 6, 349. [Google Scholar] [CrossRef]
  63. Alharbi, B.; Al-kuraishy, H.M.; Al-Gareeb, A.I.; Elekhnawy, E.; Alharbi, H.; Alexiou, A.; Papadakis, M.; El-Saber Batiha, G. Role of GABA pathway in motor and non-motor symptoms in Parkinson’s disease: A bidirectional circuit. Eur. J. Med. Res. 2024, 29, 205. [Google Scholar] [CrossRef] [PubMed]
  64. Carello-Collar, G.; Bellaver, B.; Ferreira, P.C.L.; Ferrari-Souza, J.P.; Ramos, V.G.; Therriault, J.; Tissot, C.; De Bastiani, M.A.; Soares, C.; Pascoal, T.A.; et al. The GABAergic system in Alzheimer’s disease: A systematic review with meta-analysis. Mol. Psychiatry 2023, 28, 5025–5036. [Google Scholar] [CrossRef] [PubMed]
  65. Ramírez, M.J. 5-HT6 receptors and Alzheimer’s disease. Alzheimers Res. Ther. 2013, 5, 15. [Google Scholar] [CrossRef] [PubMed]
  66. Rai, S.N.; Singh, P.; Varshney, R.; Chaturvedi, V.K.; Vamanu, E.; Singh, M.P.; Singh, B.K. Promising drug targets and associated therapeutic interventions in Parkinson’s disease. Neural Regen. Res. 2021, 16, 1730. [Google Scholar] [CrossRef]
  67. Pingale, T.; Gupta, G.L. Current and emerging therapeutic targets for Parkinson’s disease. Metab. Brain Dis. 2021, 36, 13–27. [Google Scholar] [CrossRef]
  68. Kumari, S.; Maddeboina, K.; Bachu, R.D.; Boddu, S.H.S.; Trippier, P.C.; Tiwari, A.K. Pivotal role of nitrogen heterocycles in Alzheimer’s disease drug discovery. Drug Discov. Today 2022, 27, 103322. [Google Scholar] [CrossRef]
  69. Hiremathad, A.; Piemontese, L. Heterocyclic compounds as key structures for the interaction with old and new targets in Alzheimer’s disease therapy. Neural Regen. Res. 2017, 12, 1256–1261. [Google Scholar] [CrossRef]
  70. Varadharajan, A.; Davis, A.D.; Ghosh, A.; Jagtap, T.; Xavier, A.; Menon, A.J.; Roy, D.; Gandhi, S.; Gregor, T. Guidelines for pharmacotherapy in Alzheimer’s disease—A primer on FDA-approved drugs. J. Neurosci. Rural. Pr. 2023, 14, 566. [Google Scholar] [CrossRef]
  71. Alhazmi, H.A.; Albratty, M. An update on the novel and approved drugs for Alzheimer disease. Saudi Pharm. J. SPJ 2022, 30, 1755. [Google Scholar] [CrossRef]
  72. Chopade, P.; Chopade, N.; Zhao, Z.; Mitragotri, S.; Liao, R.; Chandran Suja, V. Alzheimer’s and Parkinson’s disease therapies in the clinic. Bioeng. Transl. Med. 2023, 8, e10367. [Google Scholar] [CrossRef]
  73. Barresi, E.; Baglini, E.; Poggetti, V.; Castagnoli, J.; Giorgini, D.; Salerno, S.; Taliani, S.; Da Settimo, F. Indole-Based Compounds in the Development of Anti-Neurodegenerative Agents. Molecules 2024, 29, 2127. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, H.; Wei, W.; Zhao, M.; Ma, L.; Jiang, X.; Pei, H.; Cao, Y.; Li, H. Interaction between Aβ and Tau in the Pathogenesis of Alzheimer’s Disease. Int. J. Biol. Sci. 2021, 17, 2181. [Google Scholar] [CrossRef] [PubMed]
  75. Roda, A.; Serra-Mir, G.; Montoliu-Gaya, L.; Tiessler, L.; Villegas, S. Amyloid-beta peptide and tau protein crosstalk in Alzheimer’s disease. Neural Regen. Res. 2022, 17, 1666–1674. [Google Scholar] [CrossRef] [PubMed]
  76. Busche, M.A.; Hyman, B.T. Synergy between amyloid-β and tau in Alzheimer’s disease. Nat. Neurosci. 2020, 23, 1183–1193. [Google Scholar] [CrossRef]
  77. Abyadeh, M.; Gupta, V.; Paulo, J.A.; Mahmoudabad, A.G.; Shadfar, S.; Mirshahvaladi, S.; Gupta, V.; Nguyen, C.T.O.; Finkelstein, D.I.; You, Y.; et al. Amyloid-beta and tau protein beyond Alzheimer’s disease. Neural Regen. Res. 2024, 19, 1262–1276. [Google Scholar] [CrossRef]
  78. Pan, L.; Meng, L.; He, M.; Zhang, Z. Tau in the Pathophysiology of Parkinson’s Disease. J. Mol. Neurosci. 2021, 71, 2179–2191. [Google Scholar] [CrossRef]
  79. Srinivasan, E.; Chandrasekhar, G.; Chandrasekar, P.; Anbarasu, K.; Vickram, A.S.; Karunakaran, R.; Rajasekaran, R.; Srikumar, P.S. Alpha-Synuclein Aggregation in Parkinson’s Disease. Front. Med. 2021, 8, 736978. [Google Scholar] [CrossRef]
  80. Vidović, M.; Rikalovic, M.G. Alpha-Synuclein Aggregation Pathway in Parkinson’s Disease: Current Status and Novel Therapeutic Approaches. Cells 2022, 11, 1732. [Google Scholar] [CrossRef]
  81. Burré, J.; Edwards, R.H.; Halliday, G.; Lang, A.E.; Lashuel, H.A.; Melki, R.; Murayama, S.; Outeiro, T.F.; Papa, S.M.; Stefanis, L.; et al. Research Priorities on the Role of α-Synuclein in Parkinson’s Disease Pathogenesis. Mov. Disord. 2024, 39, 1663–1678. [Google Scholar] [CrossRef]
  82. Tofaris, G.K. Initiation and progression of α-synuclein pathology in Parkinson’s disease. Cell. Mol. Life Sci. 2022, 79, 210. [Google Scholar] [CrossRef]
  83. Pan, L.; Li, C.; Meng, L.; Tian, Y.; He, M.; Yuan, X.; Zhang, G.; Zhang, Z.; Xiong, J.; Chen, G.; et al. Tau accelerates α-synuclein aggregation and spreading in Parkinson’s disease. Brain 2022, 145, 3454–3471. [Google Scholar] [CrossRef] [PubMed]
  84. Ichimata, S.; Yoshida, K.; Li, J.; Rogaeva, E.; Lang, A.E.; Kovacs, G.G. The molecular spectrum of amyloid-beta (Aβ) in neurodegenerative diseases beyond Alzheimer’s disease. Brain Pathol. 2024, 34, e13210. [Google Scholar] [CrossRef] [PubMed]
  85. Bu, M.; Farrer, M.J.; Khoshbouei, H. Dynamic control of the dopamine transporter in neurotransmission and homeostasis. Npj Park. Dis. 2021, 7, 22. [Google Scholar] [CrossRef] [PubMed]
  86. Palermo, G.; Giannoni, S.; Bellini, G.; Siciliano, G.; Ceravolo, R. Dopamine Transporter Imaging, Current Status of a Potential Biomarker: A Comprehensive Review. Int. J. Mol. Sci. 2021, 22, 11234. [Google Scholar] [CrossRef]
  87. Shaikh, A.; Ahmad, F.; Teoh, S.L.; Kumar, J.; Yahaya, M.F. Targeting dopamine transporter to ameliorate cognitive deficits in Alzheimer’s disease. Front. Cell Neurosci. 2023, 17, 1292858. [Google Scholar] [CrossRef]
  88. Kang, S.; Jeon, S.; Lee, Y.-G.; Ye, B.S. Striatal dopamine transporter uptake, parkinsonism and cognition in Alzheimer’s disease. Eur. J. Neurol. 2023, 30, 3105–3113. [Google Scholar] [CrossRef]
  89. Obaid, R.J.; Naeem, N.; Mughal, E.U.; Al-Rooqi, M.M.; Sadiq, A.; Jassas, R.S.; Moussa, Z.; Ahmed, S.A. Inhibitory potential of nitrogen, oxygen and sulfur containing heterocyclic scaffolds against acetylcholinesterase and butyrylcholinesterase. RSC Adv. 2022, 12, 19764–19855. [Google Scholar] [CrossRef]
  90. Reis, J.; Encarnacao, I.; Gaspar, A.; Morales, A.; Milhazes, N.; Borges, F. Parkinson’s Disease Management. Part II- Discovery of MAO-B Inhibitors Based on Nitrogen Heterocycles and Analogues. Curr. Top Med. Chem. 2013, 12, 2116–2130. [Google Scholar] [CrossRef]
  91. Carradori, S.; Fantacuzzi, M.; Ammazzalorso, A.; Angeli, A.; De Filippis, B.; Galati, S.; Petzer, A.; Petzer, J.P.; Poli, G.; Tuccinardi, T.; et al. Resveratrol Analogues as Dual Inhibitors of Monoamine Oxidase B and Carbonic Anhydrase VII: A New Multi-Target Combination for Neurodegenerative Diseases? Molecules 2022, 27, 7816. [Google Scholar] [CrossRef]
  92. Pollard, A.; Shephard, F.; Freed, J.; Liddell, S.; Chakrabarti, L. Mitochondrial proteomic profiling reveals increased carbonic anhydrase II in aging and neurodegeneration. Aging 2016, 8, 2425. [Google Scholar] [CrossRef]
  93. Lemon, N.; Canepa, E.; Ilies, M.A.; Fossati, S. Carbonic Anhydrases as Potential Targets Against Neurovascular Unit Dysfunction in Alzheimer’s Disease and Stroke. Front. Aging Neurosci. 2021, 13, 772278. [Google Scholar] [CrossRef] [PubMed]
  94. Poggetti, V.; Salerno, S.; Baglini, E.; Barresi, E.; Da Settimo, F.; Taliani, S. Carbonic Anhydrase Activators for Neurodegeneration: An Overview. Molecules 2022, 27, 2544. [Google Scholar] [CrossRef] [PubMed]
  95. Duarte, P.; Cuadrado, A.; León, R. Monoamine Oxidase Inhibitors: From Classic to New Clinical Approaches. Handb. Exp. Pharmacol. 2020, 264, 229–259. [Google Scholar] [CrossRef]
  96. Jones, D.N.; Raghanti, M.A. The role of monoamine oxidase enzymes in the pathophysiology of neurological disorders. J. Chem. Neuroanat. 2021, 114, 101957. [Google Scholar] [CrossRef]
  97. Sharma, P.; Srivastava, P.; Seth, A.; Tripathi, P.N.; Banerjee, A.G.; Shrivastava, S.K. Comprehensive review of mechanisms of pathogenesis involved in Alzheimer’s disease and potential therapeutic strategies. Prog. Neurobiol. 2019, 174, 53–89. [Google Scholar] [CrossRef]
  98. Behl, T.; Kaur, D.; Sehgal, A.; Singh, S.; Sharma, N.; Zengin, G.; Andronie-Cioara, F.L.; Toma, M.M.; Bungau, S.; Bumbu, A.G. Role of Monoamine Oxidase Activity in Alzheimer’s Disease: An Insight into the Therapeutic Potential of Inhibitors. Molecules 2021, 26, 3724. [Google Scholar] [CrossRef]
  99. Manzoor, S.; Hoda, N. A comprehensive review of monoamine oxidase inhibitors as Anti-Alzheimer’s disease agents: A review. Eur. J. Med. Chem. 2020, 206, 112787. [Google Scholar] [CrossRef]
  100. Jayan, J.; Chandran, N.; Thekkantavida, A.C.; Abdelgawad, M.A.; Ghoneim, M.M.; Shaker, M.E.; Prerna, U.; Feba, B.; Zachariah, M.S.; Kumar, S.; et al. Piperidine: A Versatile Heterocyclic Ring for Developing Monoamine Oxidase Inhibitors. ACS Omega 2023, 8, 37731–37751. [Google Scholar] [CrossRef]
  101. Dhiman, P.; Malik, N.; Khatkar, A. Natural based piperine derivatives as potent monoamine oxidase inhibitors: An in silico ADMET analysis and molecular docking studies. BMC Chem. 2020, 14, 12. [Google Scholar] [CrossRef]
  102. Chavarria, D.; Fernandes, C.; Silva, V.; Silva, C.; Gil-Martins, E.; Soares, P.; Silva, T.; Silva, R.; Remião, F.; Oliveira, P.J.; et al. Design of novel monoamine oxidase-B inhibitors based on piperine scaffold: Structure-activity-toxicity, drug-likeness and efflux transport studies. Eur. J. Med. Chem. 2020, 185, 111770. [Google Scholar] [CrossRef]
  103. Dauvilliers, Y.; Tafti, M.; Landolt, H.P. Catechol-O-methyltransferase, dopamine, and sleep-wake regulation. Sleep. Med. Rev. 2015, 22, 47–53. [Google Scholar] [CrossRef] [PubMed]
  104. Perkovic, M.N.; Strac, D.S.; Tudor, L.; Konjevod, M.; Erjavec, G.N.; Pivac, N. Catechol-O-methyltransferase, Cognition and Alzheimer’s Disease. Curr. Alzheimer Res. 2018, 15, 408–419. [Google Scholar] [CrossRef] [PubMed]
  105. Akhtar, M.J.; Yar, M.S.; Grover, G.; Nath, R. Neurological and psychiatric management using COMT inhibitors: A review. Bioorg. Chem. 2020, 94, 103418. [Google Scholar] [CrossRef] [PubMed]
  106. Serretti, A.; Olgiati, P. Catechol-o-methyltransferase and Alzheimer’s disease: A review of biological and genetic findings. CNS Neurol. Disord. Drug Targets 2012, 11, 299–305. [Google Scholar] [CrossRef]
  107. Fabbri, M.; Ferreira, J.J.; Rascol, O. COMT Inhibitors in the Management of Parkinson’s Disease. CNS Drugs 2022, 36, 261–282. [Google Scholar] [CrossRef]
  108. Salamon, A.; Zádori, D.; Szpisjak, L.; Klivényi, P.; Vécsei, L. What is the impact of catechol-O-methyltransferase (COMT) on Parkinson’s disease treatment? Expert. Opin. Pharmacother. 2022, 23, 1123–1128. [Google Scholar] [CrossRef]
  109. Bharat, V.; Durairaj, A.S.; Vanhauwaert, R.; Li, L.; Muir, C.M.; Chandra, S.; Kwak, C.S.; Le Guen, Y.; Nandakishore, P.; Hsieh, C.H.; et al. A mitochondrial inside-out iron-calcium signal reveals drug targets for Parkinson’s disease. Cell Rep. 2023, 42, 113544. [Google Scholar] [CrossRef]
  110. Jurcau, A. Insights into the Pathogenesis of Neurodegenerative Diseases: Focus on Mitochondrial Dysfunction and Oxidative Stress. Int. J. Mol. Sci. 2021, 22, 11847. [Google Scholar] [CrossRef]
  111. Lee, K.H.; Cha, M.; Lee, B.H. Neuroprotective Effect of Antioxidants in the Brain. Int. J. Mol. Sci. 2020, 21, 7152. [Google Scholar] [CrossRef]
  112. Misrani, A.; Tabassum, S.; Yang, L. Mitochondrial Dysfunction and Oxidative Stress in Alzheimer’s Disease. Front. Aging Neurosci. 2021, 13, 617588. [Google Scholar] [CrossRef]
  113. Jomova, K.; Raptova, R.; Alomar, S.Y.; Alwasel, S.H.; Nepovimova, E.; Kuca, K.; Valko, M. Reactive oxygen species, toxicity, oxidative stress, and antioxidants: Chronic diseases and aging. Arch. Toxicol. 2023, 97, 2499. [Google Scholar] [CrossRef] [PubMed]
  114. Tapias, V.; González-Andrés, P.; Peña, L.F.; Barbero, A.; Núñez, L.; Villalobos, C. Therapeutic Potential of Heterocyclic Compounds Targeting Mitochondrial Calcium Homeostasis and Signaling in Alzheimer’s Disease and Parkinson’s Disease. Antioxidants 2023, 12, 1282. [Google Scholar] [CrossRef] [PubMed]
  115. Baracaldo-Santamaría, D.; Avendaño-Lopez, S.S.; Ariza-Salamanca, D.F.; Rodriguez-Giraldo, M.; Calderon-Ospina, C.A.; González-Reyes, R.E.; Nava-Mesa, M.O. Role of Calcium Modulation in the Pathophysiology and Treatment of Alzheimer’s Disease. Int. J. Mol. Sci. 2023, 24, 9067. [Google Scholar] [CrossRef] [PubMed]
  116. Pizarro-Galleguillos, B.M.; Kunert, L.; Brüggemann, N.; Prasuhn, J. Neuroinflammation and Mitochondrial Dysfunction in Parkinson’s Disease: Connecting Neuroimaging with Pathophysiology. Antioxidants 2023, 12, 1411. [Google Scholar] [CrossRef] [PubMed]
  117. Yao, M.F.; Dang, T.; Wang, H.J.; Zhu, X.Z.; Qiao, C. Mitochondrial homeostasis regulation: A promising therapeutic target for Parkinson’s disease. Behav. Brain Res. 2024, 459, 114811. [Google Scholar] [CrossRef]
  118. Wang, J.; Zhao, J.; Zhao, K.; Wu, S.; Chen, X.; Hu, W. The Role of Calcium and Iron Homeostasis in Parkinson’s Disease. Brain Sci. 2024, 14, 88. [Google Scholar] [CrossRef]
  119. Orfali, R.; Alwatban, A.Z.; Orfali, R.S.; Lau, L.; Chea, N.; Alotaibi, A.M.; Nam, Y.W.; Zhang, M. Oxidative stress and ion channels in neurodegenerative diseases. Front. Physiol. 2024, 15, 1320086. [Google Scholar] [CrossRef]
  120. Heravi, M.M.; Zadsirjan, V. Prescribed drugs containing nitrogen heterocycles: An overview. RSC Adv. 2020, 10, 44247–44311. [Google Scholar] [CrossRef]
  121. Chauhan, A.; Salahuddin, S.; Mazumder, A.; Kumar, R.; Jawed Ahsan, M.; Shahar Yar, M.; Maqsood, R.; Singh, K. Targeted Development of Pyrazoline Derivatives for Neurological Disorders: A Review. ChemistrySelect 2024, 9, e202400738. [Google Scholar] [CrossRef]
  122. Martins, P.; Jesus, J.; Santos, S.; Raposo, L.R.; Roma-Rodrigues, C.; Baptista, P.V.; Fernandes, A.R. Heterocyclic Anticancer Compounds: Recent Advances and the Paradigm Shift towards the Use of Nanomedicine’s Tool Box. Molecules 2015, 20, 16852. [Google Scholar] [CrossRef]
  123. Joubert, J.; Geldenhuys, W.J.; VanderSchyf, C.J.; Oliver, D.W.; Kruger, H.G.; Govender, T.; Malan, S.F. Polycyclic Cage Structures as Lipophilic Scaffolds for Neuroactive Drugs. ChemMedChem 2012, 7, 375–384. [Google Scholar] [CrossRef] [PubMed]
  124. Gulcin, İ.; Petrova, O.V.; Taslimi, P.; Malysheva, S.F.; Schmidt, E.Y.; Sobenina, L.N.; Gusarova, N.K.; Trofimov, B.A.; Tuzun, B.; Farzaliyev, V.M.; et al. Synthesis, Characterization, Molecular Docking, Acetylcholinesterase and α-Glycosidase Inhibition Profiles of Nitrogen-Based Novel Heterocyclic Compounds. ChemistrySelect 2022, 7, e202200370. [Google Scholar] [CrossRef]
  125. Pathania, S.; Narang, R.K.; Rawal, R.K. Role of sulphur-heterocycles in medicinal chemistry: An update. Eur. J. Med. Chem. 2019, 180, 486–508. [Google Scholar] [CrossRef] [PubMed]
  126. Venkatachalam, H.; Kumar, N.V.A.; Venkatachalam, H.; Kumar, N.V.A. The Oxygen-Containing Fused Heterocyclic Compounds. In Heterocycles—Synthesis and Biological Activities; IntechOpen: Geneva, Switzerland, 2019. [Google Scholar] [CrossRef]
  127. Singh, K.; Pal, R.; Khan, S.A.; Kumar, B.; Akhtar, M.J. Insights into the structure activity relationship of nitrogen-containing heterocyclics for the development of antidepressant compounds: An updated review. J. Mol. Struct. 2021, 1237, 130369. [Google Scholar] [CrossRef]
  128. Khan, B.A.; ur Rehman, O.; Alsfouk, A.A.; Ejaz, S.A.; Channar, P.A.; Saeed, A.; Ghafoor, A.; Ujan, R.; Mughal, E.U.; Kumar, R.; et al. Substituted piperidine as a novel lead molecule for the treatment of Parkinson’s disease: Synthesis, crystal structure, hirshfeld surface analysis, and molecular modeling. J. Mol. Struct. 2022, 1265, 133350. [Google Scholar] [CrossRef]
  129. Al-Baghdadi, O.B.; Prater, N.I.; Van Der Schyf, C.J.; Geldenhuys, W.J. Inhibition of monoamine oxidase by derivatives of piperine, an alkaloid from the pepper plant Piper nigrum, for possible use in Parkinson’s disease. Bioorg. Med. Chem. Lett. 2012, 22, 7183–7188. [Google Scholar] [CrossRef]
  130. Li, C.Y.; Zhao, L.M.; Shi, X.W.; Zhang, J.D. Lobeline shows protective effects against MPTP-induced dopaminergic neuron death and attenuates behavior deficits in animals. Exp. Ther. Med. 2014, 7, 375–378. [Google Scholar] [CrossRef]
  131. Jones, C.B.; Eltit, J.M.; Dukat, M. Do 2-(Benzoyl)piperidines Represent a Novel Class of hDAT Reuptake Inhibitors? ACS Chem. Neurosci. 2023, 14, 741–748. [Google Scholar] [CrossRef]
  132. Hritcu, L.; Noumedem, J.A.; Cioanca, O.; Hancianu, M.; Kuete, V.; Mihasan, M. Methanolic extract of Piper nigrum fruits improves memory impairment by decreasing brain oxidative stress in amyloid beta(1–42) rat model of Alzheimer’s disease. Cell Mol. Neurobiol. 2014, 34, 437–449. [Google Scholar] [CrossRef]
  133. Zhou, L.C.; Liang, Y.F.; Huang, Y.; Yang, G.X.; Zheng, L.L.; Sun, J.M.; Li, Y.; Zhu, F.L.; Qian, H.W.; Wang, R.; et al. Design, synthesis, and biological evaluation of diosgenin-indole derivatives as dual-functional agents for the treatment of Alzheimer’s disease. Eur. J. Med. Chem. 2021, 219, 113426. [Google Scholar] [CrossRef]
  134. Azmy, E.M.; Nassar, I.F.; Hagras, M.; Fawzy, I.M.; Hegazy, M.; Mokhtar, M.M.; Yehia, A.M.; Ismail, N.S.; Lashin, W.H. New Indole Derivatives As Multitarget anti-Alzheimer’s Agents: Synthesis, Biological Evaluation and Molecular Dynamics. Future Med. Chem. 2023, 15, 473–495. [Google Scholar] [CrossRef] [PubMed]
  135. Carrieri, A.; Barbarossa, A.; de Candia, M.; Samarelli, F.; Altomare, C.D.; Czarnota-Łydka, K.; Sudoł-Tałaj, S.; Latacz, G.; Handzlik, J.; Brunetti, L.; et al. Chiral pyrrolidines as multipotent agents in Alzheimer and neurodegenerative diseases. Bioorg. Med. Chem. 2024, 110, 117829. [Google Scholar] [CrossRef] [PubMed]
  136. Li Petri, G.; Raimondi, M.V.; Spanò, V.; Holl, R.; Barraja, P.; Montalbano, A. Pyrrolidine in Drug Discovery: A Versatile Scaffold for Novel Biologically Active Compounds. Top. Curr. Chem. 2021, 379, 34. [Google Scholar] [CrossRef] [PubMed]
  137. Poyraz, S.; Döndaş, H.A.; Sansano, J.M.; Belveren, S.; Yamali, C.; Ülger, M.; Döndaş, N.Y.; Sağlık, B.N.; Pask, C.M. N-Benzoylthiourea-pyrrolidine carboxylic acid derivatives bearing an imidazole moiety: Synthesis, characterization, crystal structure, in vitro ChEs inhibition, and antituberculosis, antibacterial, antifungal studies. J. Mol. Struct. 2023, 1273, 134303. [Google Scholar] [CrossRef]
  138. Lv, Y.; Fan, M.; He, J.; Song, X.; Guo, J.; Gao, B.; Zhang, J.; Zhang, C.; Xie, Y. Discovery of novel benzimidazole derivatives as selective and reversible monoamine oxidase B inhibitors for Parkinson’s disease treatment. Eur. J. Med. Chem. 2024, 274, 116566. [Google Scholar] [CrossRef]
  139. Ahsan, N.; Mishra, S.; Jain, M.K.; Surolia, A.; Gupta, S. Curcumin Pyrazole and its derivative (N-(3-Nitrophenylpyrazole) Curcumin inhibit aggregation, disrupt fibrils and modulate toxicity of Wild type and Mutant α-Synuclein. Sci. Rep. 2015, 5, 9862. [Google Scholar] [CrossRef]
  140. Patel, D.V.; Patel, N.R.; Kanhed, A.M.; Teli, D.M.; Patel, K.B.; Joshi, P.D.; Patel, S.P.; Gandhi, P.M.; Chaudhary, B.N.; Prajapati, N.K.; et al. Novel carbazole-stilbene hybrids as multifunctional anti-Alzheimer agents. Bioorg. Chem. 2020, 101, 103977. [Google Scholar] [CrossRef]
  141. Chiu, Y.J.; Lin, C.H.; Lin, C.Y.; Yang, P.N.; Lo, Y.S.; Chen, Y.C.; Chen, C.M.; Wu, Y.R.; Yao, C.F.; Chang, K.H.; et al. Investigating Therapeutic Effects of Indole Derivatives Targeting Inflammation and Oxidative Stress in Neurotoxin-Induced Cell and Mouse Models of Parkinson’s Disease. Int. J. Mol. Sci. 2023, 24, 2642. [Google Scholar] [CrossRef]
  142. Saini, N.; Akhtar, A.; Chauhan, M.; Dhingra, N.; Pilkhwal Sah, S. Protective effect of Indole-3-carbinol, an NF-κB inhibitor in experimental paradigm of Parkinson’s disease: In silico and in vivo studies. Brain Behav. Immun. 2020, 90, 108–137. [Google Scholar] [CrossRef]
  143. Elbatrawy, A.A.; Ademoye, T.A.; Alnakhala, H.; Tripathi, A.; Zami, A.; Ostafe, R.; Dettmer, U.; Fortin, J.S. Discovery of small molecule benzothiazole and indole derivatives tackling tau 2N4R and α-synuclein fibrils. Bioorg Med. Chem. 2024, 100, 117613. [Google Scholar] [CrossRef]
  144. Bansal, R.; Singh, R.; Rana, P. Synthesis, in silico Studies and Pharmacological Evaluation of a New Series of Indanone Derivatives as Anti-Parkinsonian and Anti-Alzheimer’s Agents. Curr. Comput. Aided Drug Des. 2023, 19, 94–107. [Google Scholar] [CrossRef]
  145. Michalska, P.; Mayo, P.; Fernández-Mendívil, C.; Tenti, G.; Duarte, P.; Buendia, I.; Ramos, M.T.; López, M.G.; Menéndez, J.C.; León, R. Antioxidant, Anti-inflammatory and Neuroprotective Profiles of Novel 1,4-Dihydropyridine Derivatives for the Treatment of Alzheimer’s Disease. Antioxidants 2020, 9, 650. [Google Scholar] [CrossRef] [PubMed]
  146. Gutti, G.; Kumar, D.; Paliwal, P.; Ganeshpurkar, A.; Lahre, K.; Kumar, A.; Krishnamurthy, S.; Singh, S.K. Development of pyrazole and spiropyrazoline analogs as multifunctional agents for treatment of Alzheimer’s disease. Bioorg. Chem. 2019, 90, 103080. [Google Scholar] [CrossRef]
  147. Jayaraj, R.L.; Elangovan, N.; Dhanalakshmi, C.; Manivasagam, T.; Essa, M.M. CNB-001, a novel pyrazole derivative mitigates motor impairments associated with neurodegeneration via suppression of neuroinflammatory and apoptotic response in experimental Parkinson’s disease mice. Chem. Biol. Interact. 2014, 220, 149–157. [Google Scholar] [CrossRef] [PubMed]
  148. Silva, D.; Chioua, M.; Samadi, A.; Carmo Carreiras, M.; Jimeno, M.L.; Mendes, E.; Ríos Cde, L.; Romero, A.; Villarroya, M.; López, M.G.; et al. Synthesis and pharmacological assessment of diversely substituted pyrazolo[3,4-b]quinoline, and benzo[b]pyrazolo[4,3-g][1,8]naphthyridine derivatives. Eur. J. Med. Chem. 2011, 46, 4676–4681. [Google Scholar] [CrossRef] [PubMed]
  149. Pisani, L.; Farina, R.; Catto, M.; Iacobazzi, R.M.; Nicolotti, O.; Cellamare, S.; Mangiatordi, G.F.; Denora, N.; Soto-Otero, R.; Siragusa, L.; et al. Exploring Basic Tail Modifications of Coumarin-Based Dual Acetylcholinesterase-Monoamine Oxidase B Inhibitors: Identification of Water-Soluble, Brain-Permeant Neuroprotective Multitarget Agents. J. Med. Chem. 2016, 59, 6791–6806. [Google Scholar] [CrossRef]
  150. González-Muñoz, G.C.; Arce, M.P.; Pérez, C.; Romero, A.; Villarroya, M.; López, M.G.; Conde, S.; Rodríguez-Franco, M.I. Dibenzo[1,4,5]thiadiazepine: A hardly-known heterocyclic system with neuroprotective properties of potential usefulness in the treatment of neurodegenerative diseases. Eur. J. Med. Chem. 2014, 81, 350–358. [Google Scholar] [CrossRef]
  151. Lincoln, K.M.; Richardson, T.E.; Rutter, L.; Gonzalez, P.; Simpkins, J.W.; Green, K.N. An N-heterocyclic amine chelate capable of antioxidant capacity and amyloid disaggregation. ACS Chem. Neurosci. 2012, 3, 919–927. [Google Scholar] [CrossRef]
  152. Matos, M.J.; Rodríguez-Enríquez, F.; Borges, F.; Santana, L.; Uriarte, E.; Estrada, M.; Rodríguez-Franco, M.I.; Laguna, R.; Viña, D. 3-Amidocoumarins as Potential Multifunctional Agents against Neurodegenerative Diseases. ChemMedChem 2015, 10, 2071–2079. [Google Scholar] [CrossRef]
  153. Papagiouvannis, G.; Theodosis-Nobelos, P.; Rekka, E.A. Nipecotic Acid Derivatives as Potent Agents against Neurodegeneration: A Preliminary Study. Molecules 2022, 27, 6984. [Google Scholar] [CrossRef]
  154. Sirakanyan, S.N.; Spinelli, D.; Petrou, A.; Geronikaki, A.; Kartsev, V.G.; Hakobyan, E.K.; Yegoryan, H.A.; Zuppiroli, L.; Zuppiroli, R.; Ayvazyan, A.G.; et al. New Bicyclic Pyridine-Based Hybrids Linked to the 1,2,3-Triazole Unit: Synthesis via Click Reaction and Evaluation of Neurotropic Activity and Molecular Docking. Molecules 2023, 28, 921. [Google Scholar] [CrossRef] [PubMed]
  155. Porcal, W.; Hernández, P.; González, M.; Ferreira, A.; Olea-Azar, C.; Cerecetto, H.; Castro, A. Heteroarylnitrones as drugs for neurodegenerative diseases: Synthesis, neuroprotective properties, and free radical scavenger properties. J. Med. Chem. 2008, 51, 6150–6159. [Google Scholar] [CrossRef] [PubMed]
  156. Escribano, L.; Simón, A.M.; Gimeno, E.; Cuadrado-Tejedor, M.; De Maturana, R.L.; García-Osta, A.; Pérez-Mediavilla, A.; Del Río, J.; Frechilla, D. Rosiglitazone Rescues Memory Impairment in Alzheimer’s Transgenic Mice: Mechanisms Involving a Reduced Amyloid and Tau Pathology. Neuropsychopharmacology 2010, 35, 1593–1604. [Google Scholar] [CrossRef] [PubMed]
  157. Mokhtari, Z.; Baluchnejadmojarad, T.; Nikbakht, F.; Mansouri, M.; Roghani, M. Riluzole ameliorates learning and memory deficits in Aβ25-35-induced rat model of Alzheimer’s disease and is independent of cholinoceptor activation. Biomed. Pharmacother. 2017, 87, 135–144. [Google Scholar] [CrossRef]
  158. Vanda, D.; Zajdel, P.; Soural, M. Imidazopyridine-based selective and multifunctional ligands of biological targets associated with psychiatric and neurodegenerative diseases. Eur. J. Med. Chem. 2019, 181, 111569. [Google Scholar] [CrossRef]
  159. Sridhar, P.; Alagumuthu, M.; Ram, B.; Arumugam, S.; Reddy, S.R. Drugs Against Neurodegenerative Diseases: Design and Synthesis of 6-Amino–substituted Imidazo[1,2-b]pyridazines as Acetylcholinesterase Inhibitors. ChemistrySelect 2017, 2, 842–847. [Google Scholar] [CrossRef]
  160. Işık, M.; Demir, Y.; Durgun, M.; Türkeş, C.; Necip, A.; Beydemir, Ş. Molecular docking and investigation of 4-(benzylideneamino)- and 4-(benzylamino)-benzenesulfonamide derivatives as potent AChE inhibitors. Chem. Pap. 2020, 74, 1395–1405. [Google Scholar] [CrossRef]
  161. Sever, B.; Türkeş, C.; Altıntop, M.D.; Demir, Y.; Beydemir, Ş. Thiazolyl-pyrazoline derivatives: In vitro and in silico evaluation as potential acetylcholinesterase and carbonic anhydrase inhibitors. Int. J. Biol. Macromol. 2020, 163, 1970–1988. [Google Scholar] [CrossRef]
  162. Gündoğdu, S.; Türkeş, C.; Arslan, M.; Demir, Y.; Beydemir, Ş. New Isoindole-1,3-dione Substituted Sulfonamides as Potent Inhibitors of Carbonic Anhydrase and Acetylcholinesterase: Design, Synthesis, and Biological Evaluation. ChemistrySelect 2019, 4, 13347–13355. [Google Scholar] [CrossRef]
  163. Andrade-Jorge, E.; Sánchez-Labastida, L.A.; Soriano-Ursúa, M.A.; Guevara-Salazar, J.A.; Trujillo-Ferrara, J.G. Isoindolines/isoindoline-1,3-diones as AChE inhibitors against Alzheimer’s disease, evaluated by an improved ultra-micro assay. Med. Chem. Res. 2018, 27, 2187–2198. [Google Scholar] [CrossRef]
  164. Azimi, S.; Firuzi, O.; Iraji, A.; Zonouzi, A.; Khoshneviszadeh, M.; Mahdavi, M.; Edraki, N. Synthesis and In Vitro Biological Activity Evaluation of Novel Imidazo [2,1-B][1,3,4] Thiadiazole as Anti-Alzheimer Agents. Lett. Drug Des. Discov. 2018, 17, 610–617. [Google Scholar] [CrossRef]
  165. Bergin, J.C.J.; Tan, K.K.; Nelson, A.K.; Amarandei, C.A.; Hubscher-Bruder, V.; Brandel, J.; Voinarovska, V.; Dejaegere, A.; Stote, R.H.; Tétard, D. 1-Hydroxy-2(1H)-pyridinone-Based Chelators with Potential Catechol O-Methyl Transferase Inhibition and Neurorescue Dual Action against Parkinson’s Disease. Molecules 2022, 27, 2816. [Google Scholar] [CrossRef] [PubMed]
  166. Shaikh, S.; Dhavan, P.; Pavale, G.; Ramana, M.M.V.; Jadhav, B.L. Design, synthesis and evaluation of pyrazole bearing α-aminophosphonate derivatives as potential acetylcholinesterase inhibitors against Alzheimer’s disease. Bioorg Chem. 2020, 96, 103589. [Google Scholar] [CrossRef] [PubMed]
  167. Engelbrecht, I.; Petzer, J.P.; Petzer, A. Nitrocatechol Derivatives of Chalcone as Inhibitors of Monoamine Oxidase and Catechol-O-Methyltransferase. Cent. Nerv. Syst. Agents Med. Chem. 2018, 18, 115–127. [Google Scholar] [CrossRef]
  168. Hussain, R.; Rahim, F.; Rehman, W.; Khan, S.; Rasheed, L.; Maalik, A.; Taha, M.; Alanazi, M.M.; Alanazi, A.S.; Khan, I. Synthesis, in vitro analysis and molecular docking study of novel benzoxazole-based oxazole derivatives for the treatment of Alzheimer’s disease. Arab. J. Chem. 2023, 16, 105244. [Google Scholar] [CrossRef]
  169. García Marín, I.D.; Camarillo López, R.H.; Martínez, O.A.; Padilla-Martínez, I.I.; Correa-Basurto, J.; Rosales-Hernández, M.C. New compounds from heterocyclic amines scaffold with multitarget inhibitory activity on Aβ aggregation, AChE, and BACE1 in the Alzheimer disease. PLoS ONE 2022, 17, e0269129. [Google Scholar] [CrossRef]
  170. Benek, O.; Hroch, L.; Aitken, L.; Dolezal, R.; Guest, P.; Benkova, M.; Soukup, O.; Musil, K.; Kuca, K.; Smith, T.K.; et al. 6-benzothiazolyl ureas, thioureas and guanidines are potent inhibitors of ABAD/17β-HSD10 and potential drugs for Alzheimer’s disease treatment: Design, synthesis and in vitro evaluation. Med. Chem. 2017, 13, 345–358. [Google Scholar] [CrossRef]
  171. Patil, S.B. Recent medicinal approaches of novel pyrimidine analogs: A review. Heliyon 2023, 9, e16773. [Google Scholar] [CrossRef]
  172. Pant, S.; Kapri, A.; Nain, S. Pyrimidine analogues for the management of neurodegenerative diseases. Eur. J. Med. Chem. Rep. 2022, 6, 100095. [Google Scholar] [CrossRef]
  173. Costa, R.F.; Turones, L.C.; Cavalcante, K.V.N.; Rosa Júnior, I.A.; Xavier, C.H.; Rosseto, L.P.; Napolitano, H.B.; Castro, P.F.D.S.; Neto, M.L.F.; Galvão, G.M.; et al. Heterocyclic Compounds: Pharmacology of Pyrazole Analogs From Rational Structural Considerations. Front. Pharmacol. 2021, 12, 666725. [Google Scholar] [CrossRef]
  174. Li, X.; Yu, Y.; Tu, Z. Pyrazole Scaffold Synthesis, Functionalization, and Applications in Alzheimer’s Disease and Parkinson’s Disease Treatment (2011–2020). Molecules 2021, 26, 1202. [Google Scholar] [CrossRef] [PubMed]
  175. Turkan, F.; Cetin, A.; Taslimi, P.; Gulçin, İ. Some pyrazoles derivatives: Potent carbonic anhydrase, α-glycosidase, and cholinesterase enzymes inhibitors. Arch. Pharm. 2018, 351, 1800200. [Google Scholar] [CrossRef] [PubMed]
  176. Ahsan, M.J.; Ali, A.; Ali, A.; Thiriveedhi, A.; Bakht, M.A.; Yusuf, M.; Salahuddin Afzal, O.; Altamimi, A.S.A. Pyrazoline Containing Compounds as Therapeutic Targets for Neurodegenerative Disorders. ACS Omega 2022, 7, 38207–38245. [Google Scholar] [CrossRef] [PubMed]
  177. Hitge, R.; Smit, S.; Petzer, A.; Petzer, J.P. Evaluation of nitrocatechol chalcone and pyrazoline derivatives as inhibitors of catechol-O-methyltransferase and monoamine oxidase. Bioorg Med. Chem. Lett. 2020, 30, 127188. [Google Scholar] [CrossRef]
  178. Abdelgawad, M.A.; Oh, J.M.; Parambi, D.G.T.; Kumar, S.; Musa, A.; Ghoneim, M.M.; Nayl, A.A.; El-Ghorab, A.H.; Iqrar Ahmad Patel, H.; Kim, H.; et al. Development of bromo- and fluoro-based α, β-unsaturated ketones as highly potent MAO-B inhibitors for the treatment of Parkinson’s disease. J. Mol. Struct. 2022, 1266, 133545. [Google Scholar] [CrossRef]
  179. Mathew, B.; Parambi, D.G.T.; Mathew, G.E.; Uddin, M.S.; Inasu, S.T.; Kim, H.; Marathakam, A.; Unnikrishnan, M.K.; Carradori, S. Emerging therapeutic potentials of dual-acting MAO and AChE inhibitors in Alzheimer’s and Parkinson’s diseases. Arch. Pharm. 2019, 352, 1900177. [Google Scholar] [CrossRef]
  180. Husain, A.; Balushi, K.A.; Akhtar, M.J.; Khan, S.A. Coumarin linked heterocyclic hybrids: A promising approach to develop multi target drugs for Alzheimer’s disease. J. Mol. Struct. 2021, 1241, 130618. [Google Scholar] [CrossRef]
  181. Petrou, A.; Fesatidou, M.; Geronikaki, A. Thiazole Ring—A Biologically Active Scaffold. Molecules 2021, 26, 3166. [Google Scholar] [CrossRef]
  182. Mishra, C.B.; Kumari, S.; Tiwari, M. Thiazole: A promising heterocycle for the development of potent CNS active agents. Eur. J. Med. Chem. 2015, 92, 1–34. [Google Scholar] [CrossRef]
  183. Askin, S.; Tahtaci, H.; Türkeş, C.; Demir, Y.; Ece, A.; Akalın Çiftçi, G.; Beydemir, Ş. Design, synthesis, characterization, in vitro and in silico evaluation of novel imidazo[2,1-b][1,3,4]thiadiazoles as highly potent acetylcholinesterase and non-classical carbonic anhydrase inhibitors. Bioorg. Chem. 2021, 113, 105009. [Google Scholar] [CrossRef]
  184. Wang, Y.; Wang, H.; Xu, H.; Zheng, Z.; Meng, Z.; Xu, Z.; Li, J.; Xue, M. Design and synthesis of five-membered heterocyclic derivatives of istradefylline with comparable pharmacological activity. Chem. Biol. Drug Des. 2022, 100, 534–552. [Google Scholar] [CrossRef] [PubMed]
  185. Oz, M.; Lorke, D.E.; Hasan, M.; Petroianu, G.A. Cellular and molecular actions of Methylene Blue in the nervous system. Med. Res. Rev. 2011, 31, 93–117. [Google Scholar] [CrossRef] [PubMed]
  186. Akoury, E.; Pickhardt, M.; Gajda, M.; Biernat, J.; Mandelkow, E.; Zweckstetter, M. Mechanistic basis of phenothiazine-driven inhibition of Tau aggregation. Angew. Chem. Int. Ed. Engl. 2013, 52, 3511–3515. [Google Scholar] [CrossRef]
  187. Kassab, A.E.; Gedawy, E.M.; Sayed, A.S. Fused thiophene as a privileged scaffold: A review on anti-Alzheimer’s disease potentials via targeting cholinesterases, monoamine oxidases, glycogen synthase kinase-3, and Aβ aggregation. Int. J. Biol. Macromol. 2024, 265, 131018. [Google Scholar] [CrossRef] [PubMed]
  188. Kiss, L.E.; Bonifácio, M.J.; Rocha, J.F.; Soares-da-Silva, P. Opicapone, a Novel Catechol-O-Methyltranferase Inhibitor (COMT) to Manage the Symptoms of Parkinson’s Disease. Success. Drug Discov. 2018, 3, 319–340. [Google Scholar] [CrossRef]
  189. Shen, Z.B.; Meng, H.W.; Meng, X.S.; Lv, Z.K.; Fang, M.Y.; Zhang, L.L.; Lv, Z.L.; Li, M.S.; Liu, A.K.; Han, J.H.; et al. Design, synthesis, and SAR study of novel flavone 1,2,4-oxadiazole derivatives with anti-inflammatory activities for the treatment of Parkinson’s disease. Eur. J. Med. Chem. 2023, 255, 115417. [Google Scholar] [CrossRef]
  190. Królicka, E.; Kieć-Kononowicz, K.; Łażewska, D. Chalcones as Potential Ligands for the Treatment of Parkinson’s Disease. Pharmaceuticals 2022, 15, 847. [Google Scholar] [CrossRef]
  191. Ceyhun, İ.; Karaca, Ş.; Osmaniye, D.; Sağlık, B.N.; Levent, S.; Özkay, Y.; Kaplancıklı, Z.A. Design and synthesis of novel chalcone derivatives and evaluation of their inhibitory activities against acetylcholinesterase. Arch. Pharm. 2022, 355, e2100372. [Google Scholar] [CrossRef]
  192. Pibiri, I. Recent Advances: Heterocycles in Drugs and Drug Discovery. Int. J. Mol. Sci. 2024, 25, 9503. [Google Scholar] [CrossRef]
  193. Adelusi, T.I.; Oyedele, A.Q.K.; Boyenle, I.D.; Ogunlana, A.T.; Adeyemi, R.O.; Ukachi, C.D.; Idris, M.O.; Olaoba, O.T.; Adedotun, I.O.; Kolawole, O.E.; et al. Molecular modeling in drug discovery. Inf. Med. Unlocked 2022, 29, 100880. [Google Scholar] [CrossRef]
  194. Lim, H.-S. Evolving role of modeling and simulation in drug development. Transl. Clin. Pharmacol. 2019, 27, 19. [Google Scholar] [CrossRef] [PubMed]
  195. Wang, W.; Ye, Z.; Gao, H.; Ouyang, D. Computational pharmaceutics—A new paradigm of drug delivery. J. Control. Release 2021, 338, 119–136. [Google Scholar] [CrossRef] [PubMed]
  196. AlRawashdeh, S.; Barakat, K.H. Applications of Molecular Dynamics Simulations in Drug Discovery. Methods Mol. Biol. 2024, 2714, 127–141. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Pathogenic factors of Alzheimer’s disease and Parkinson’s disease. These factors include oxidative stress/mitochondrial stress, neuroinflammation and neurodegeneration, and plaques of insoluble proteins including Aβ, tau, and α-syn.
Figure 1. Pathogenic factors of Alzheimer’s disease and Parkinson’s disease. These factors include oxidative stress/mitochondrial stress, neuroinflammation and neurodegeneration, and plaques of insoluble proteins including Aβ, tau, and α-syn.
Neurolint 17 00026 g001
Figure 2. Major therapeutic targets of Alzheimer’s and Parkinson’s disease. As AD and PD share major common etiological factors, they also share some therapeutic targets, which mostly include proteins, enzymes, receptors, and mitochondrial/oxidative stress.
Figure 2. Major therapeutic targets of Alzheimer’s and Parkinson’s disease. As AD and PD share major common etiological factors, they also share some therapeutic targets, which mostly include proteins, enzymes, receptors, and mitochondrial/oxidative stress.
Neurolint 17 00026 g002
Figure 3. FDA-approved drugs for treatment and management of Alzheimer’s disease.
Figure 3. FDA-approved drugs for treatment and management of Alzheimer’s disease.
Neurolint 17 00026 g003
Figure 4. FDA-approved drugs for treatment and management of Parkinson’s disease.
Figure 4. FDA-approved drugs for treatment and management of Parkinson’s disease.
Neurolint 17 00026 g004
Figure 5. Chemical structure of COMT inhibitors entacapone, opicapone, and tolcapone used in PD treatment and management.
Figure 5. Chemical structure of COMT inhibitors entacapone, opicapone, and tolcapone used in PD treatment and management.
Neurolint 17 00026 g005
Figure 6. General overview of the significant roles of heterocyclic compounds in the treatment of Alzheimer’s disease and Parkinson’s disease. Various heterocyclic motifs exhibit diverse activities, including anticholinesterase activity, antioxidant properties, inhibition of protein aggregation, regulation of neurotransmitter release in synapses, maintenance of mitochondrial homeostasis, modulation of neurotransmitter receptor activation, and reduction in behavioral impairments.
Figure 6. General overview of the significant roles of heterocyclic compounds in the treatment of Alzheimer’s disease and Parkinson’s disease. Various heterocyclic motifs exhibit diverse activities, including anticholinesterase activity, antioxidant properties, inhibition of protein aggregation, regulation of neurotransmitter release in synapses, maintenance of mitochondrial homeostasis, modulation of neurotransmitter receptor activation, and reduction in behavioral impairments.
Neurolint 17 00026 g006
Table 1. FDA-approved heterocyclic drugs for AD and PD treatment.
Table 1. FDA-approved heterocyclic drugs for AD and PD treatment.
DrugTherapeutic CategoryBrand NameTargetFDA ApprovalApplication
BrexpiprazoleAtypical antipsychoticRexultiNovel D2 dopamine and serotonin 1A partial agonist2015Approved for the treatment of depression, schizophrenia, and agitation associated with dementia due to AD
DonepezilParasympathomemeticAdlarity, Aricept, NamzaricAChE inhibitor1996Used to treat the behavioral and cognitive effects of AD and other types of dementia
RivastigmineParasympathomemeticExelon, Nimvastid, PrometaxInhibits both BChE and AChE2000Used to treat mild to moderate dementia in AD and PD
GalantamineParasympathomemeticRazadyneCompetitive inhibitor of AChE2001Reduces the severity of dementia in patients with AD
MemantineNAMD
antagonist
Axura, Ebixa, Marixino, Namenda, NamzaricNMDA receptor antagonist2013Blocks the effects of Glu in the brain that lead to neuronal excitability and excessive stimulation in AD
SafinamideMAO-B
inhibitor
XadagoSelective and reversible inhibition of MAO-B with blockade of voltage-dependent Na+ and Ca2+ channels and inhibition of Glu release2017Neuroprotective and neuro-rescuing effects in PD
Istradefylline-NourianzTargets adenosine A2A receptors in the basal ganglia2019Used to treat reduced GABAergic action and motor control in PD patients
PimavanserinAtypical antipsychoticNuplazidInteracts with the serotonin receptors, particularly the 5-HT2A and HT2C receptors2016Used to treat PD- associated psychotic symptoms without causing extrapyramidal or worsening motor symptoms
AmantadineInfluenza A M2 protein inhibitorGocovri, OsmolexReleasing dopamine from the nerve endings of the brain cells, and stimulation of norepinephrine response; it also has NMDA receptor antagonistic effects1973 (for PD)Used to treat dyskinesia in Parkinson’s patients
BenzatropineAnticholinergic agent and histamine antagonistCogentinSelective inhibition of dopamine transporters1996Used as an adjunct in the therapy of all forms of parkinsonism
BiperidenAnticholinergic agentNACompetitive antagonism of ACh at cholinergic receptors in the corpus striatum1959An adjunct in the therapy of all forms of parkinsonism and control of extrapyramidal disorders secondary to neuroleptic drug therapy
TrihexyphenidylAnticholinergic agent and
muscarinic antagonist
NAACh receptor (M1 subtype) antagonist1949Anticholinergic activity useful in the treatment of symptoms associated with PD
CarbidopaAADC inhibitorCrexont, Dhivy, Duodopa, Duopa, Lodosyn, Parcopa, Rytary, Sinemet, StalevoDopa decarboxylase inhibitor used in combination with levodopa2014Symptomatic treatment of idiopathic PD and other conditions associated with parkinsonian symptoms
EntacaponeCOMT inhibitorComtan, Comtess, StalevoAdministered concomittantly with levodopa and carbidopa; increased and more sustained plasma levodopa concentrations are reached as compared to the administration of levodopa and a decarboxylase inhibitor1999Symptomatic treatment of patients with idiopathic PD
TolcaponeCOMT inhibitorTasmarInhibits the enzyme COMT used as an adjunct1998Helps to improve the symptoms of PD such as trembling, difficulty with movement, stiffness, and other symptoms
OpicaponeCOMT inhibitorOngentysCOMT inhibitor used as an adjunct2020Managing motor and some nonmotor symptoms associated with PD
NA—not available; COMT—catechol-O-methyl-transferase; MAO-B—monoamine oxidase B; NAMD—N-methyl-D-aspartate receptor; ACh—acetylcholine; AChE—acetylcholinesterase; BChE—butyrylcholinesterase; AADC—aromatic L-amino acid decarboxylase; Glu—glutamate.
Table 4. Summary of synthetic organic compounds under clinical trials for the management of Alzheimer’s and Parkinson’s.
Table 4. Summary of synthetic organic compounds under clinical trials for the management of Alzheimer’s and Parkinson’s.
DrugIUPAC NameCompound and ClassStructureDiseasesMechanism of ActionClinical Trial NoClinical PhaseSponsor Company
Semagacestat
(LY-450139)
(2S)-2-hydroxy-3-methyl-N-[(2S)-1-[[(1S)-3-methyl-2-oxo-4,5-dihydro-1H-3-benzazepin-1-yl]amino]-1-oxopropan-2-yl]butanamideOrganic acids and derivatives
Class: carboxylic acids and derivatives
Neurolint 17 00026 i001ADSmall-molecule γ-secretase inhibitorNCT01035138Phase 3 Eli Lilly and Company
ANAVEX2-73
(Blarcamesine)
[(2,2-diphenyloxolan-3-yl)methyl]dimethylamineNANeurolint 17 00026 i002ADCognition and functionNCT03790709Phase 2b/3Anavex Life Sciences Corp.
Dimebon
(Latrepirdine)
5-(2-{2,8-dimethyl-1H,2H,3H,4H,5H-pyrido[4,3-b]indol-5-yl}ethyl)-2-methylpyridineOrganoheterocyclic compounds
Class: indoles and derivatives
Neurolint 17 00026 i003ADImproves cognition in models of ADNCT00912288Phase 3Pfizer
ALZ- 801
(Valiltramiprosate)
3-[[(2S)-2-amino-3-methylbutanoyl]amino]propane-1-sulfonic acidNANeurolint 17 00026 i004ADPrevents Aβ42 from forming oligomersNCT06304883Phase 3Alzheon Inc.
Isradipine3-methyl 5-propan-2-yl 4-(2,1,3-benzoxadiazol-4-yl)-2,6-dimethyl-1,4-dihydropyridine-3,5-dicarboxylateOrganoheterocyclic compounds
Class:
benzoxadiazoles
Neurolint 17 00026 i005PDCalcium channel blockerNCT02168842Phase 3University of Rochester
Nilvadipine3-O-methyl 5-O-propan-2-yl 2-cyano-6-methyl-4-(3-nitrophenyl)-1,4-dihydropyridine-3,5-dicarboxylateOrganoheterocyclic compounds
Class: pyridines and derivatives
Neurolint 17 00026 i006 Calcium channel blockerNCT02017340Phase 3Prof Brian Lawlor, St. James’s Hospital, Ireland
NA: not available.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Puranik, N.; Song, M. Therapeutic Role of Heterocyclic Compounds in Neurodegenerative Diseases: Insights from Alzheimer’s and Parkinson’s Diseases. Neurol. Int. 2025, 17, 26. https://doi.org/10.3390/neurolint17020026

AMA Style

Puranik N, Song M. Therapeutic Role of Heterocyclic Compounds in Neurodegenerative Diseases: Insights from Alzheimer’s and Parkinson’s Diseases. Neurology International. 2025; 17(2):26. https://doi.org/10.3390/neurolint17020026

Chicago/Turabian Style

Puranik, Nidhi, and Minseok Song. 2025. "Therapeutic Role of Heterocyclic Compounds in Neurodegenerative Diseases: Insights from Alzheimer’s and Parkinson’s Diseases" Neurology International 17, no. 2: 26. https://doi.org/10.3390/neurolint17020026

APA Style

Puranik, N., & Song, M. (2025). Therapeutic Role of Heterocyclic Compounds in Neurodegenerative Diseases: Insights from Alzheimer’s and Parkinson’s Diseases. Neurology International, 17(2), 26. https://doi.org/10.3390/neurolint17020026

Article Metrics

Back to TopTop