Next Article in Journal
Extracellular Vesicles and Antiphospholipid Syndrome: State-of-the-Art and Future Challenges
Next Article in Special Issue
Human-Induced Neural and Mesenchymal Stem Cell Therapy Combined with a Curcumin Nanoconjugate as a Spinal Cord Injury Treatment
Previous Article in Journal
Transcriptome and MiRNAomics Analyses Identify Genes Associated with Cytoplasmic Male Sterility in Cotton (Gossypium hirsutum L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Applying Computer-Aided Drug Design for Neurodegenerative Diseases

1
Department of Physiology, Anatomy and Genetics, University of Oxford, Parks Road, Oxford OX1 3QX, UK
2
Oxford Parkinson’s Disease Centre, University of Oxford, South Parks Road, Oxford OX1 3QX, UK
3
Department of Pharmaceutical Chemistry, College of Pharmacy, University of Alkafeel, Najaf 54001, Iraq
4
Department of Chemistry and Biochemistry, College of Medicine, University of Kerbala, Karbala 56001, Iraq
5
School of Biosciences, College of Health and Life Sciences, Aston University, Aston Triangle, Birmingham B4 7ET, UK
6
John Van Geest Centre for Brain Repair, University of Cambridge, Cambridge CB2 0PY, UK
*
Authors to whom correspondence should be addressed.
Equal contribution.
Int. J. Mol. Sci. 2021, 22(9), 4688; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22094688
Submission received: 29 March 2021 / Revised: 26 April 2021 / Accepted: 26 April 2021 / Published: 28 April 2021
(This article belongs to the Special Issue CNS Injuries)

Abstract

:
Neurodegenerative diseases (NDs) including Alzheimer’s disease, Parkinson’s disease, amyotrophic lateral sclerosis, and Huntington’s disease are incurable and affect millions of people worldwide. The development of treatments for this unmet clinical need is a major global research challenge. Computer-aided drug design (CADD) methods minimize the huge number of ligands that could be screened in biological assays, reducing the cost, time, and effort required to develop new drugs. In this review, we provide an introduction to CADD and examine the progress in applying CADD and other molecular docking studies to NDs. We provide an updated overview of potential therapeutic targets for various NDs and discuss some of the advantages and disadvantages of these tools.

1. Introduction

Neurodegenerative diseases (NDs) are incurable and debilitating conditions that result in progressive degeneration and/or death of nerve cells in the central nervous system (CNS) [1,2,3]. Dementia rates are alarmingly on the rise worldwide. There are over 50 million people worldwide living with dementia in 2020, with nearly 60% living in low- and middle-income countries [4]. This number will almost double every 20 years, reaching 82 million in 2030 and 152 million in 2050 [4]. The number of people with dementia in the UK is predicted to be around 1.14 million by 2025 and 2.1 million by 2051, an increase of 40% over the next 5 years and 157% over the next 31 years [5].
The UK Prime Minister’s Challenge on Dementia was launched in 2015 to identify strategies to tackle dementia by 2025 [6]. Current therapies for NDs treat symptoms, not the underlying pathological changes. There is a clear and unmet clinical need to develop new therapies based on understanding the molecular pathologies. One of the most promising approaches is to develop novel therapeutics using computer-aided drug design (CADD) [7,8].
In this review, we provide an introduction to CADD and different approaches involved in this technique. We provide a list of over 200 pieces of CADD software using a citation-based scoring system (Supplementary Table S1), with the 30 most commonly used software products listed in Table 1. We examine the progress in applying CADD and other molecular docking studies to NDs, provide an updated overview of potential therapeutic targets for various NDs, and discuss some of the advantages and disadvantages of these tools.

2. Computer-Aided Drug Design

“Computer-aided drug design” (CADD) refers to the application of computational modelling approaches to drug discovery. Drug discovery is an expensive and time-consuming process with the average approved drug requiring 10 to 15 years to develop with an estimated cost of 0.8–2 billion USD [9]. Many licensed drugs, such as captopril, dorzolamide, oseltamivir, aliskiren, and nolatrexed, were all optimized using CADD [10], and a large number of publications describe the successful design and discovery of leads/drugs using CADD [11,12,13]. The major steps involved in CADD are summarized in Figure 1A and discussed in the following sections. The main goal of CADD is to reduce these timescales and costs without affecting quality (Figure 1B) [14]. Importantly, CADD can be used in most stages of drug development: from target identification to target validation, from lead discovery to optimization, and in preclinical studies. It is therefore estimated that CADD could reduce the cost of drug development by up to 50% [15,16].

2.1. Drug Target Selection

Drug target selection is the first step of structure-based drug design. This involves identifying and determining the structures of the relevant proteins [17]. Understanding and characterization of the molecular biology of the targeted disease are therefore necessary before the initiation of any active compound search process.

2.2. Determination of the Protein Structure

An in-depth understanding of biological processes is still often hampered by a lack of detailed knowledge of protein structures [18]. The determination of the structure of the target protein is a prerequisite for CADD [19]. Structural elucidation of the target protein can be performed by experimental tools including, but not limited to, nuclear magnetic resonance (NMR) spectroscopy, Cryo-EM, and X-ray crystallography [20,21].

2.3. Homology Modelling

Despite the current revolution in structural studies, in particular the recent developments in cryo-EM, the detailed structures of a large number of proteins, and especially membrane proteins (which are over-represented amongst drug targets), have not been determined [18,22]. Homology modelling is an approach to estimate the structure of a target protein based on structural data from proteins with sequence homology to the target [23].
For instance, a homology model of human catechol-O-methyltransferase (COMT) was constructed utilizing the X-ray crystal structure of rat COMT to design anti-PD drugs by performing ligand docking, resulting in the discovery of nine putative inhibitors. Another example involves a cysteine protease from Xanthomonas campestris (an aerobic, Gram-negative rod-shaped bacterium known to cause black rot in crucifers by darkening the vascular tissues). The active site of this enzyme is homologous to human cathepsin B enzyme (hCB), the activity of which contributes to the reduction of the amyloid peptide by proteolytic cleavage of Aβ1-42, offering a protective role against AD [24].

2.4. Identification of Binding Sites

When the three-dimensional structure of the target protein is determined, the next step is the identification of potential binding sites for small molecules. This process can be conducted using various algorithms for computing and identifying binding pockets [25,26,27].

2.5. Molecular Dynamics Simulation

Molecular dynamics (MD) simulations are a theoretical tool to discover the configurations and dynamic behaviours of molecules, providing atomic-level insight into drug mechanisms of action [13]. MD may also help to reveal the aggregation pathway of neurotoxic protein aggregates and thus aid in the design of new inhibitors [28].

2.6. Molecular Docking Studies

Molecular docking is a computational procedure that predicts the lowest energy binding conformations of one molecule to a second (usually a small drug-like molecule to a protein). Accordingly, molecular docking procedures, along with their different scoring systems, are frequently utilized to predict the binding modes and affinities between chemical compounds and drug binding sites on biological macromolecules [29,30].

2.7. Virtual Screening

Virtual screening (VS) is the process of screening small molecule libraries in silico to identify chemical structures that may bind to a drug target [31,32,33].

2.8. Quantitative Structure—Activity Relationship Study

Quantitative structure—activity relationship (QSAR) methods are conducted to correlate a biological response (e.g., enzyme activity, cell viability, etc.) to the chemical properties of a set of molecules [34,35,36].

2.9. Pharmacophore Modelling

Pharmacophore modelling deals with finding the optimal shapes and charge distributions for binding of a small molecule to a biological macromolecule. Pharmacophore modelling is commonly implemented to rapidly specify potential lead compounds [37,38].

3. Neurodegenerative Diseases

NDs include Alzheimer’s disease (AD), Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS), and Huntington’s disease (HD) [39]. These diseases are diverse in their pathophysiology and effective treatments are urgently needed, but they will only be achieved with an in-depth understanding of the causes and mechanisms of each disease. These diseases and potential drug targets for each are discussed briefly below. Current molecular targets for these diseases, along with examples of drugs discovered in CADD projects, are summarised in Table 2. The molecular mechanisms of neurodegeneration and potential drug targets in these diseases are summarised in Figure 2.

3.1. Alzheimer’s Disease (AD)

AD is a chronic, progressive, and persistent neurodegenerative disease whose main symptoms are reduced motor and cognitive function and accelerated memory loss, resulting from the progressive loss of neurons and synapses in the cerebral cortex, ultimately leading to death [66].
The estimated number of people aged 65 years or older in the USA with AD in 2010 was 4.7 million and this number is predicted to reach 13.8 million by 2050 [67]. In 2013, the number of people in the UK with dementia was estimated at 815,827, of which 62% had AD [68]. Approximately 70% of the UK care home population suffers from dementia and more than 42,000 people below 65 years also have dementia [69].
AD is characterised by the presence of amyloid plaques, composed primarily of aggregated amyloid-β (Aβ) peptides proteolytically derived from the amyloid precursor protein (APP), and neurofibrillary tangles (NFTs) that are intracellular protein aggregates composed primarily of phosphorylated tau protein. Although amyloid deposits are thought to develop before NFTs, amyloid burden is poorly correlated with disease progression, whereas NFT burden is more strongly correlated [70]. The exact mechanisms by which NFTs and Aβ plaques lead to neurodegeneration are still poorly understood. Several genetic contributors to AD have been identified, including variants of presenilin 1 (PSEN1) and presenilin 2 (PSEN2), components of the γ-secretase complex that cleaves an APP intermediate to its amyloidogenic forms, as well as variants of APP itself. The strongest genetic risk factor not directly involved in amyloid formation is the APOE gene, encoding an apolipoprotein that is responsible for CNS cholesterol transport. Weaker genetic risk factors include a variety of genes involved in cholesterol metabolism, endocytosis, and neuroinflammation [71,72]. Recent work on the glymphatic waste clearance system suggests that reduced glymphatic function is correlated with Aβ and tau accumulation [73]. Bulk flow through the glymphatic system is elevated during sleep and mediated by the water channel protein AQP4, which is also implicated in various CNS pathologies [74,75,76]. Despite there being no single drug that has been approved to successfully target AQP4 [77], new studies suggest that modulators of sleep or AQP4 (by targeting the trafficking mechanism or membrane abundance rather than pore-blocking) could be novel targets for early intervention in AD and other protein-misfolding diseases [78,79].

3.1.1. Macromolecular Targets in AD

Acetylcholinesterase

Acetylcholinesterase inhibitors (AChEIs) have been considered as potential drugs to treat AD and other dementias for many years, due to the degeneration and loss of cholinergic neurons associated with AD symptoms. Indeed, three of the four currently approved drugs for AD are AChEIs (donepezil, galantamine, and rivastigmine). Accordingly, acetylcholinesterase is routinely targeted in docking studies [80]. For example, utilizing molecular docking, the binding of compounds found in Salvia miltiorrhiza (red sage) extract, e.g., miltirone and salvianolic acid A, to acetylcholinesterase [81], and the binding of cinerin C (a molecule extracted from Prosopis cineraria pods) to acetylcholinesterase [82] have been reported.

Beta-Secretase and Gamma-Secretase Enzymes

Aβ formation is catalysed by β-secretase (BACE) and γ-secretase (GS) enzymes and, thus, inhibiting these enzymes could prevent Aβ plaque formation and prevent AD [83]. Molecular docking has been utilized to score putative inhibitors of GS, and the highest scoring compound was used to identify chemically similar compounds for pharmacophore mapping [84].

Caspases

Caspases are important mediators of apoptosis in neurons (and indeed in most cell types); their inhibition might therefore be helpful in preventing neurodegeneration-associated neuronal death in ALS, AD, PD, and HD [85,86,87,88].
Several studies employed in silico drug design and molecular docking to target caspases to treat NDs. For example, ten non-cytotoxic nitrones were assessed for their capability to arrest apoptosis and reduce the levels of active caspase-3 and oxidative stress in the HT22 neuronal cell line. Molecular docking suggested that these nitrones bound to a site near the catalytic region of caspase-3. This suggested that medicinal chemistry using these nitrones as a starting point could be considered to begin the development of novel ND therapies [89].

Acetylcholine (ACh) Receptors

Many studies conducted both in vitro and in vivo have demonstrated that reduced cholinergic activity is a direct cause of memory loss in AD patients [90]. Consequently, one of the potential targets in AD is the nicotinic acetylcholine receptor (nAChR). Compounds discovered using multitarget CADD studies based on nicotinic receptors were found to improve memory, cognition, and spatial capabilities in animal models [91,92].

N-Methyl-D-Aspartate Receptor

N-methyl-D-aspartate (NMDA) receptors transduce glutamate and glycine signals that play crucial roles in CNS development and the synaptic plasticity that is essential for memory and learning processes [93]. However, overexposure to glutamate can result in neurotransmission disturbances correlated with the NMDA receptor, which are treatable with NDMA antagonists [94,95]. The identification of conantokins, MK-801 and memantine (memantine was approved by the FDA for AD in 2004), as NMDA receptor inhibitors led to the investigation of these structures using CADD to identify new NMDA receptor inhibitors. New compounds discovered in this way could be utilized as potential AD therapeutics [47,94,96,97].

ROCK-I and NOX2 Enzymes

One of the possible approaches to treat neuroinflammation is the inhibition of both NADPH oxidase 2 (NOX2) and Rho kinase 1 (ROCK-I). This might be an effective way to treat some progressive neurological diseases, including AD [98]. NOX2 is the catalytic subunit of a multi-protein complex that can be activated in host defence phagocytic processes (e.g., in microglia) to govern the generation of superoxide from oxygen. ROCK-I is a significant mediator of cell migration, proliferation, and adhesion. In disease states, NOX2 integration into the NADPH oxidase complex is activated by ROCK-I via Ras associated C3 botulinum toxin substrate (Rac). Consequently, microglial cells with high ROCK-I and NOX2 lead to progressive neuronal damage in the early development of neurological disease [99]. In one study, CADD was utilized to discover new molecules with the ability to inhibit both ROCK-I and NOX2, with 18 compounds identified from a library of 5 × 105. Of these 18 molecules, 7 had an inhibitory effect against both enzymes in cell-based assays [98].

3.2. Parkinson’s Disease (PD)

PD is the second most common neurodegenerative disorder with symptoms including tremors, muscle rigidity, and postural imbalance [100,101,102]. PD affected around 145,000 people in the UK in 2019 [103,104,105]. In the USA, the estimated number of annual PD diagnoses is 60,000 and approximately one million are affected with PD in 2020 [106,107]. PD is characterised by preferential and progressive loss of dopaminergic neurons starting in the substantia nigra pars compacta, and the presence of intracellular aggregates, known as Lewy bodies, composed primarily of the protein α-synuclein. Exactly how (or even if) Lewy bodies exert neurotoxic effects is poorly understood.

3.2.1. Macromolecular Targets in PD

COMT (Catechol-O-Methyltransferase) Inhibitors

COMT metabolises catechols by methylation. As dopamine is one of the catechols that is reduced in the CNS during PD, COMT is considered a drug target for the management of PD. Nitrocatechol-type inhibitors (e.g., tolcapone and entacapone), bisubstrate inhibitors (e.g., thiopyridine, purine, N-methyladenine, and 6-methylpurine), and other molecules (e.g., 4-phenyl-7,8-dihydroxycoumarin) were reported as potential COMT inhibitors from structure-based drug design studies [108].

Dopamine Agonists

Pergolide, pramipexole, ropinirole, bromocriptine, and piribedil are currently the most commonly prescribed dopamine-receptor agonists. They are generally combined with levodopa plus dopa decarboxylase inhibitors (DDIs), especially in patients with motor dysfunctions. They can be efficient as a monotherapy during early PD (they can delay the need for the introduction of levodopa plus DDIs in newly diagnosed patients) or in combination with levodopa plus DDIs for dyskinesia and motor fluctuations [109].
There are five subtypes of dopamine receptors, D1–5 and each one has a different function. Different patients may respond differently to different dopamine receptor agonists. Hence, clinicians often change the therapeutic choice from one dopamine receptor agonist to another in order to achieve better control of PD symptoms and avoid specific side-effects [110]. D1, D2, and D3 receptors primarily control locomotor activity. Moreover, D1 and D2 receptors (and to a lesser extent D3) are essential in memory and learning mechanisms, mainly in the prefrontal cortex [111]. D2 receptors have a crucial function in psychotic behaviours since almost all effective antipsychotic drugs antagonize D2 receptors. The D3 receptor is primarily expressed in the limbic area of the brain [112]. D4 receptors are associated with relapse to stimulant use and selective D4 inhibitors/antagonists might be potential therapies for drug-relapse.
Outside the CNS, dopamine is also implicated in cardiovascular and renal functions, mainly through D1 and D2 receptors. Heterodimerization of dopamine receptors in various biological systems further complicates the role of dopaminergic interactions in PD [113]; therefore, designing more specific effective drugs using molecular docking might be a viable strategy to achieve drugs with fewer adverse effects in PD patients.

Gene Variants

Variants in a variety of genes have been reported to be associated with PD, including SNCA (encoding α-synuclein), ADH1C, DJ-1, EIF4G1, FBXO7, GBA/GBAP1, GIGYF2, HTRA2, LRRK2 [114], MAPT, PARK2, PARK7 [115], PRKN, PINK1, PLA2G6, UCHL1, and VPS35 [116]. For example, several mutations to LRRK2, encoding the leucine-rich repeat kinase 2 (LRRK2), are associated with PD and it has been reported as a significant factor for drug resistance [117,118]. A panel of 160 kinase inhibitors was examined for their activity against LRRK2 in vitro employing a peptide substrate kinase assay and neuronal SH-SY5Y cells overexpressing LRRK2 [59]. In silico docking studies utilizing the LRRK2 kinase structure and some selected compounds found a correlation between docking scores for the LRRK2 ATP binding site and both in vitro and cellular compound activity [59].

Glutamate Antagonists

Glutamate receptors can be classified into two major classes: ionotropic (iGluRs) and metabotropic receptors (mGluRs). Glutamate antagonists have well-established neuroprotective effects through slowing the rate of dopaminergic neuron loss in the substantia nigra [119]. A number of glutamate antagonists improve motor function in PD animal models through acting on α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) and NMDA subtypes of ionotropic glutamate receptors. Nonetheless, systemic administration is associated with serious side-effects such as sedation and ataxia, especially for NMDA antagonists [120]. This has substantially affected their widespread use; therefore, developing selective antagonists against specific receptor isoforms that are preferentially expressed in the critical parts of the pathophysiological circuitry might be an interesting therapeutic approach in the future.

MAO-B

Monoamine oxidase inhibitors (MAOI) were one of the earliest drugs to be tried in PD and can be used with or without levodopa. Non-selective MAOI (such as tranylcypromine) have limited use in treating PD-associated depression due to their numerous side effects, while reversible and selective MAO-A inhibitors are more recommended. Selective and irreversible MAO-B inhibitors such as selegiline and rasagiline are recommended for the control of motor fluctuations and akinesia.
Selegiline is a selective, irreversible MAO-B inhibitor that has been widely used for PD treatment. It has been shown to delay the need for levodopa during early stages of PD and managing the end-of-dose akinesia in fully developed PD patients. A number of further irreversible and reversible MAO-B inhibitors have been developed.
Safinamide is a relatively new selective reversible MAO-B inhibitor with ion channel activity that does not cause a cheese-reaction, unlike other MAO-B inhibitors [121]. This drug enhances motor function in early PD [122].

3.3. Amyotrophic Lateral Sclerosis (ALS)

ALS is a lethal condition that is characterised by progressive muscular paralysis and wasting, reflecting degeneration of neurons controlling voluntary muscles, including both the upper motor neurons in the motor cortex and lower motor neurons in the brainstem and spinal cord [123].
Around 5000 people in the USA are diagnosed with ALS each year. Cumulatively, there are more than 30,000 and 5000 people affected with ALS in the USA and UK, respectively [124,125,126].
The pathogenesis of ALS is relatively poorly understood. Only two drugs are approved for ALS: one of these is a glutamate antagonist (riluzole) and the other (edaravone) works by an unknown mechanism.

3.3.1. Macromolecular Targets in ALS

SOD1

Superoxide dismutase (SOD1) is an antioxidant enzyme involved in the detoxification of superoxide radicals. The SOD1 enzyme requires bound zinc and copper ions to maintain intra-molecular disulphide bonds [127]. Variation in zinc and copper ion binding to SOD1 leads to misfolded enzymes and can initiate aggregation and facilitate the protein instability associated with ALS.
In one study, 32,791 molecules were virtually screened by establishing an in silico assay system to screen for inhibitors of the aberrant interaction between mutant SOD1 and tubulin, with the aim of identifying lead compounds for ALS [128].
Molecular docking studies have been used to develop inhibitors of dimer destabilization and aggregation of the human SOD1 G85R mutant. CADD studies have predicted a number of inhibitors such as linear tripeptides [129], the tubulin binding site of G85R SOD1 [128], resveratrol [130], natural polyphenols of curcumin [131], kaempferol, and kaempferide [132] as potential lead compounds for treating ALS.

MAPK

Many processes within the cell, such as mitogenesis, apoptosis, oncogenesis, and differentiation, are associated with the mitogen activated protein kinases (MAPKs) [133]. MAPKs are activated by upstream kinases called MAPK kinases (termed MAPKK, MEK, or MKK) and an MAPK kinase kinase (termed MAPKKK, MEKK, or MKKK) [134], and are linked to the inhibition of proinflammatory cytokines [135]. An in silico and in vitro study of a MAPKK inhibitor (silibinin) used molecular docking to address the interactions of silibinin with p38 MAPK, which is an important kinase associated with glial cell activation and neuroinflammation [136].

Casein Kinase 1 (CK-1) Inhibitors

The protein kinase CK-1 was reported to directly phosphorylate Tyrosyl-DNA phosphodiesterase (TDP3). The latter is a DNA repair enzyme and is considered a promising target for antitumor and neurodegenerative therapy [137], and up-regulation of CK-1 is correlated with ALS [138]. Accordingly, CK-1δ inhibitors crossing the blood–brain barrier (BBB), such as riluzole and others, may be a novel approach to treat ALS [139,140,141].

3.4. Huntington’s Disease

HD is a genetic, incurable, and fatal neurodegenerative condition characterized by progressive degeneration of neurons, starting specifically with medium spiny neurons (MSNs) in the striatum, and leading to inevitable deterioration of the mental and physical abilities of those affected [149,150].
In the UK, the number of people diagnosed with HD is around 6000 people, whereas the number is around 30,000 in the USA [151,152,153,154].
HD is a monogenic disease caused by expansion of a CAG trinucleotide repeat in the HTT gene, leading to expansion of a polyglutamine tract in the Huntingtin protein, which is expressed ubiquitously throughout the brain. Mutant Huntingtin is prone to aggregation, but how this causes selective degeneration of striatal MSNs is poorly understood. Currently, no disease-modifying therapies or cures are available.
Reducing levels of mutant HTT is, understandably, a major therapeutic goal in HD. A recent study showed that intrathecal administration of the antisense oligonucleotide (ASO) IONIS-HTTRx (Tominersen) to HD patients resulted in a dose-dependent reduction of mutant HTT in the cerebrospinal fluid (CSF) [155]. Tominersen was rapidly moved to a Phase III trial. However, a press release by Roche in March 2021 announced the decision to discontinue dosing of Tominersen in manifest HD in the Phase III trial.
Despite the unfortunate news, these studies suggest that ASOs administration is a viable therapeutic strategy to reduce levels of toxic proteins in NDs. How and to what extent ASOs reach different parts of the central nervous system is not fully understood yet.

3.4.1. Macromolecular Targets in HD

4-Aminobutyrate Aminotransferase

4-Aminobutyrate aminotransferase (ABAT) (PDB ID: 1OHY) is responsible for the degradation of gamma-aminobutyric acid (GABA), a major inhibitory mediator for synaptic transmission in the mammalian CNS [156]. Reduction in GABAergic transmission is the result of many genetic disorders and chronic neurological diseases, including HD, AD, PD, and epilepsy. Unfortunately, GABA is unable to cross the BBB, preventing the direct use of exogenous GABA [157]. Enhancing the levels of GABA by decreasing its degradation by ABAT is an alternative strategy. In one study, the structures of thirty-two molecules from thirty-one medicinal plants were obtained from a chemical database and were chosen with the aid of previous literature reports. These 32 natural molecules were examined in a molecular docking study in which the researchers concluded that the top-ranked compounds may be suitable candidates for in vitro and in vivo studies of ABAT inhibition [158]. Moreover, GABA derivatives have been tested for ABAT binding in silico [156].

4. A Roadmap for Implementing CADD in ND Drug Design

Even with the number of successful implementations of CADD in modern drug discovery, it has its limitations. Molecules designed in silico utilizing computational and theoretical chemistry sometimes do not work in real biological systems [159,160]. In general, poor pharmacokinetics and/or pharmacodynamics result in only 40% of drug candidates passing phase I clinical trials [161]. Moreover, each computational technique depends on pre-determined algorithms that have their own limitations. CADD results must be validated in real biological systems, as many molecules that appear to bind in silico do not show the desired activity in vitro. Another limitation of CADD is that all tools for designing and discovery of new drugs are based on algorithms that, by necessity, simplify the underlying physics and chemistry and, therefore, have a variety of limitations that necessitate the continuous updating of these algorithms to enhance the accuracy and thus the provision of new drugs [162,163,164,165,166,167,168]. Furthermore, the shortage of experimental data regarding predicted absorption, distribution, metabolism, excretion, and toxicity results has led to several published failures [169,170,171,172,173].
To overcome the limitations and improve the accuracy of CADD it is necessary to update and develop software and associated algorithms, validate with experimental data, use reliable databases (e.g., PDB), and use algorithms that give docking scores that accurately predict in vitro binding with comprehensive and fully retrospective coverage of the published literature [174,175,176]. For example, by September 2020, the Cambridge Structural Dataset (CSD) acquired more than 1.8 million entries, which may help with future developments in small molecule structural modelling [177]. Consequently, the above-mentioned tools could help with future design of pharmacophores that possess the desired biological activity [178,179,180].
One of the main reasons for implementing in silico drug design is to predict the ligand–target binding in terms of binding site and binding strength. To predict potential ligands to treat NDs, novel target proteins must be identified and studied, and the resulting docking studies should be validated in vitro and eventually in the clinic [181,182,183].
In the meantime, there is no effective treatment to cure NDs, although many treatments are available that offer minor improvement of symptoms [2]. The development of effective treatments is further hindered by the BBB that excludes many molecules from the CNS parenchyma [184,185,186]. Accordingly, clinical effectiveness of a potential drug is not guaranteed even with positive data in silico, in vitro, and in vivo [187,188,189,190].
New experimental approaches including genome-wide association studies (GWAS) [188,191,192], CRISPR-Cas9 technology [193,194,195], high throughput screening (HTS) [196], organ-on-chip technologies [197,198], functional MRI (fMRI) techniques [199,200], and positron emission tomography (PET) [201] may lead to new drug targets for NDs, which can feed into future CADD projects.
Being incurable, the NDs are major challenges to healthcare providers and research scientists. The accelerating increase in the numbers of affected people adds more impetus to tackle NDs. Developing a better understanding of NDs and the underlying molecular pathophysiology will provide more opportunities to develop novel treatments in the near future. This may be achieved with the incorporation of computational tools. CADD can have a major impact on drug discovery by saving both time and money and reducing the risk of following up with the development of non-viable leads.

Supplementary Materials

Author Contributions

Conceptualization, M.M.S.; writing—original draft preparation, M.M.S., Z.A.-O., and P.K.; table preparation, Z.A.-O. and M.M.S.; writing—review and editing, M.M.S., P.K., A.L., R.M.B., and R.W.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maciejczyk, M.; Zalewska, A. Salivary redox biomarkers in selected neurodegenerative diseases. J. Clin. Med. 2020, 9, 497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rabanel, J.-M.; Perrotte, M.; Ramassamy, C. Nanotechnology at the Rescue of Neurodegenerative Diseases: Tools for Early Diagnostic. In Nanobiotechnology in Neurodegenerative Diseases; Springer: Berlin/Heidelberg, Germany, 2019; pp. 19–48. [Google Scholar]
  3. Sehgal, S.A.; Hammad, M.A.; Tahir, R.A.; Akram, H.N.; Ahmad, F. Current Therapeutic Molecules and Targets in Neurodegenerative Diseases Based on in silico Drug Design. Curr. Neuropharmacol. 2018, 16, 649–663. [Google Scholar] [CrossRef]
  4. Livingston, G.; Huntley, J.; Sommerlad, A.; Ames, D.; Ballard, C.; Banerjee, S.; Brayne, C.; Burns, A.; Cohen-Mansfield, J.; Cooper, C. Dementia prevention, intervention, and care: 2020 report of the Lancet Commission. Lancet 2020, 396, 413–446. [Google Scholar] [CrossRef]
  5. Prince, M.; Knapp, M.; Guerchet, M.; McCrone, P.; Prina, M.; Comas-Herrera, A.; Wittenberg, R.; Adelaja, B.; Hu, B.; King, D. Dementia UK: Update; Alzheimers Society: Belfast, UK, 2014; pp. 1–136. [Google Scholar]
  6. Prime Minister’s Challenge on Dementia. Available online: https://www.gov.uk/government/publications/prime-ministers-challenge-on-dementia-2020 (accessed on 14 April 2020).
  7. Shukla, R.; Singh, T.R. Virtual screening, pharmacokinetics, molecular dynamics and binding free energy analysis for small natural molecules against cyclin-dependent kinase 5 for Alzheimer’s disease. J. Biomol. Struct. Dyn. 2020, 38, 248–262. [Google Scholar] [CrossRef]
  8. Mouchlis, V.D.; Melagraki, G.; Zacharia, L.C.; Afantitis, A. Computer-Aided Drug Design of β-Secretase, γ-Secretase and Anti-Tau Inhibitors for the Discovery of Novel Alzheimer’s Therapeutics. Int. J. Mol. Sci. 2020, 21, 703. [Google Scholar] [CrossRef] [Green Version]
  9. Am Ende, D.J.; Am Ende, M.T. Chemical engineering in the pharmaceutical industry: An introduction. Chem. Eng. Pharm. Ind. Drug Prod. Des. Dev. Modeling 2019, 1–17. [Google Scholar] [CrossRef]
  10. Talele, T.T.; Khedkar, S.A.; Rigby, A.C. Successful applications of computer aided drug discovery: Moving drugs from concept to the clinic. Curr. Top. Med. Chem. 2010, 10, 127–141. [Google Scholar] [CrossRef]
  11. Kaushik, A.C.; Kumar, A.; Bharadwaj, S.; Chaudhary, R.; Sahi, S. Structure-Based Approach for In-silico Drug Designing. In Bioinformatics Techniques for Drug Discovery; Springer: Berlin/Heidelberg, Germany, 2018; pp. 21–25. [Google Scholar]
  12. Hamad, O.; Amran, S.; Sabbah, A. Drug Discovery-Yesterday and Tomorrow: The Common Approaches in Drug Design and Cancer. Cell Cell. Life Sci. J. 2018, 3, 000119. [Google Scholar]
  13. Lu, W.; Zhang, R.; Jiang, H.; Zhang, H.; Luo, C. Computer-Aided Drug Design in Epigenetics. Front. Chem. 2018, 6. [Google Scholar] [CrossRef]
  14. Kapetanovic, I. Computer-aided drug discovery and development (CADDD): In silico-chemico-biological approach. Chem. Biol. Interact. 2008, 171, 165–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Macalino, S.J.Y.; Gosu, V.; Hong, S.; Choi, S. Role of computer-aided drug design in modern drug discovery. Arch. Pharmacal. Res. 2015, 38, 1686–1701. [Google Scholar] [CrossRef] [PubMed]
  16. Xiang, M.; Cao, Y.; Fan, W.; Chen, L.; Mo, Y. Computer-aided drug design: Lead discovery and optimization. Comb. Chem. High Throughput Screen. 2012, 15, 328–337. [Google Scholar] [CrossRef] [PubMed]
  17. Anderson, A.; Blaney, J.; Blundell, T.; Clark, D.; Davis, A.M.; Ealick, S.; Kim, S.-H.; McCammon, J.A.; Verdonk, M.; Wijnand, M. Computational and Structural Approaches to Drug Discovery: Ligand-Protein Interactions; Royal Society of Chemistry: London, UK, 2007. [Google Scholar]
  18. Hauri, S.; Khakzad, H.; Happonen, L.; Teleman, J.; Malmström, J.; Malmström, L. Rapid determination of quaternary protein structures in complex biological samples. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Baig, M.H.; Ahmad, K.; Rabbani, G.; Danishuddin, M.; Choi, I. Computer aided drug design and its application to the development of potential drugs for neurodegenerative disorders. Curr. Neuropharmacol. 2018, 16, 740–748. [Google Scholar] [CrossRef] [PubMed]
  20. Makhouri, F.R.; Ghasemi, J.B. In Silico studies in drug research against neurodegenerative diseases. Curr. Neuropharmacol. 2018, 16, 664–725. [Google Scholar] [CrossRef] [PubMed]
  21. Fawzi, M.M.; Abdallah, H.H.; Suroowan, S.; Jugreet, S.; Zhang, Y.; Hu, X. In Silico Exploration of Bioactive Phytochemicals Against Neurodegenerative Diseases via Inhibition of Cholinesterases. Curr. Pharm. Des. 2020, 26, 4151–4162. [Google Scholar] [CrossRef]
  22. Schmiedel, J.M.; Lehner, B. Determining protein structures using deep mutagenesis. Nat. Genet. 2019, 51, 1177. [Google Scholar] [CrossRef] [Green Version]
  23. Kumar, J.; Ranjan, T.; Kumar, R.R.; Ansar, M.; Rajani, K.; Kumar, M.; Kumar, V.; Kumar, A. In silico Characterization and Homology Modelling of Potato Leaf Roll Virus (PLRV) Coat Protein. Curr. J. Appl. Sci. Technol. 2019, 1–8. [Google Scholar] [CrossRef]
  24. Morales-Navarro, S.; Prent-Peñaloza, L.; Rodríguez Núñez, Y.A.; Sánchez-Aros, L.; Forero-Doria, O.; González, W.; Campilllo, N.E.; Reyes-Parada, M.; Martínez, A.; Ramírez, D. Theoretical and Experimental Approaches AiMed. at Drug Design Targeting Neurodegenerative Diseases. Processes 2019, 7, 940. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, B.; Dai, P.; Ding, D.; Del Rosario, A.; Grant, R.A.; Pentelute, B.L.; Laub, M.T. Affinity-based capture and identification of protein effectors of the growth regulator ppGpp. Nat. Chem. Biol. 2019, 15, 141–150. [Google Scholar] [CrossRef]
  26. Chan, H.S.; Li, Y.; Dahoun, T.; Vogel, H.; Yuan, S. New binding sites, new opportunities for GPCR drug discovery. Trends Biochem. Sci. 2019, 44, 312–330. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, Y.; Qiao, S.; Ji, S.; Han, N.; Liu, D.; Zhou, J. Identification of DNA–protein binding sites by bootstrap multiple convolutional neural networks on sequence information. Eng. Appl. Artif. Intell. 2019, 79, 58–66. [Google Scholar] [CrossRef]
  28. Ye, W.; Wang, W.; Jiang, C.; Yu, Q.; Chen, H. Molecular dynamics simulations of amyloid fibrils: An in silico approach. Acta Biochim. Biophys. Sin. 2013, 45, 503–508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Makarasen, A.; Kuno, M.; Patnin, S.; Reukngam, N.; Khlaychan, P.; Deeyohe, S.; Intachote, P.; Saimanee, B.; Sengsai, S.; Boonsri, P. Molecular Docking Studies and Synthesis of Amino-oxy-diarylquinoline Derivatives as Potent Non-nucleoside HIV-1 Reverse Transcriptase Inhibitors. Drug Res. 2019, 69, 671–682. [Google Scholar] [CrossRef]
  30. Vilar, S.; Sobarzo-Sánchez, E.; Uriarte, E. In Silico Prediction of P-glycoprotein Binding: Insights from Molecular Docking Studies. Curr. Med. Chem. 2019, 26, 1746–1760. [Google Scholar] [CrossRef] [PubMed]
  31. Nunes, R.R.; Fonseca, A.L.D.; Pinto, A.C.D.S.; Maia, E.H.B.; Silva, A.M.D.; Varotti, F.D.P.; Taranto, A.G. Brazilian malaria molecular targets (BraMMT): Selected receptors for virtual high-throughput screening experiments. Memórias Do Inst. Oswaldo Cruz 2019, 114, e180465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Zerroug, A.; Belaidi, S.; BenBrahim, I.; Sinha, L.; Chtita, S. Virtual screening in drug-likeness and structure/activity relationship of pyridazine derivatives as Anti-Alzheimer drugs. J. King Saud Univ. Sci. 2019, 31, 595–601. [Google Scholar] [CrossRef]
  33. Vieira, T.; Magalhaes, R.; Sousa, S. Tailoring specialized scoring functions for more efficient virtual screening. Frontiers 2019, 2, 1–4. [Google Scholar]
  34. Ray, R. Understanding the Structural Importance of the Non-Binding and Binding Parts of Bedaquiline and Its Analogues with ATP Synthase Subunit C Using Molecular Docking, Molecular Dynamics Simulation and 3D-QSAR Techniques. In Proceedings of the International Conference on Drug Discovery (ICDD) 2020, Hyderabad, India, 19 February 2020. [Google Scholar]
  35. Kotzabasaki, M.I.; Sotiropoulos, I.; Sarimveis, H. QSAR modeling of the toxicity classification of superparamagnetic iron oxide nanoparticles (SPIONs) in stem-cell monitoring applications: An integrated study from data curation to model development. RSC Adv. 2020, 10, 5385–5391. [Google Scholar] [CrossRef] [Green Version]
  36. Gbeddy, G.; Egodawatta, P.; Goonetilleke, A.; Ayoko, G.; Chen, L. Application of quantitative structure-activity relationship (QSAR) model in comprehensive human health risk assessment of PAHs, and alkyl-, nitro-, carbonyl-, and hydroxyl-PAHs laden in urban road dust. J. Hazard. Mater. 2020, 383, 121154. [Google Scholar] [CrossRef]
  37. Du, M.; Qiu, Y.; Li, Q.; Li, Y. Efficacy coefficient method assisted quadruple-activities 3D-QSAR pharmacophore model for application in environmentally friendly PAE molecular modification. Environ. Sci. Pollut. Res. Int. 2020, 27, 24103–24114. [Google Scholar] [CrossRef]
  38. Fan, F.; Warshaviak, D.T.; Hamadeh, H.K.; Dunn, R.T. The integration of pharmacophore-based 3D QSAR modeling and virtual screening in safety profiling: A case study to identify antagonistic activities against adenosine receptor, A2A, using 1,897 known drugs. PLoS ONE 2019, 14, e0204378. [Google Scholar] [CrossRef] [Green Version]
  39. Hou, Y.; Dan, X.; Babbar, M.; Wei, Y.; Hasselbalch, S.G.; Croteau, D.L.; Bohr, V.A. Ageing as a risk factor for neurodegenerative disease. Nat. Rev. Neurol. 2019, 15, 565–581. [Google Scholar] [CrossRef] [PubMed]
  40. Espinoza-Moraga, M.; Caballero, J.; Gaube, F.; Winckler, T.; Santos, L.S. 1-Benzyl-1,2,3,4-tetrahydro-β-carboline as channel blocker of N-methyl-D-aspartate receptors. Chem. Biol. Drug Des. 2012, 79, 594–599. [Google Scholar] [CrossRef]
  41. Mishra, C.B.; Kumari, S.; Manral, A.; Prakash, A.; Saini, V.; Lynn, A.M.; Tiwari, M. Design, synthesis, in-silico and biological evaluation of novel donepezil derivatives as multi-target-directed ligands for the treatment of Alzheimer’s disease. Eur. J. Med. Chem. 2017, 125, 736–750. [Google Scholar] [CrossRef]
  42. Samadi, A.; de la Fuente Revenga, M.; Pérez, C.; Iriepa, I.; Moraleda, I.; Rodríguez-Franco, M.I.; Marco-Contelles, J. Synthesis, pharmacological assessment, and molecular modeling of 6-chloro-pyridonepezils: New dual AChE inhibitors as potential drugs for the treatment of Alzheimer’s disease. Eur. J. Med. Chem. 2013, 67, 64–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ha, C.H.H.; Fatima, A.; Gaurav, A. In Silico investigation of flavonoids as potential trypanosomal nucleoside hydrolase inhibitors. Adv. Bioinform. 2015, 2015, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. De Andrade Teles, R.B.; Diniz, T.C.; Costa Pinto, T.C.; de Oliveira Júnior, R.G.; Gama E Silva, M.; de Lavor, É.M.; Fernandes, A.W.C.; de Oliveira, A.P.; de Almeida Ribeiro, F.P.R.; da Silva, A.A.M.; et al. Flavonoids as Therapeutic Agents in Alzheimer’s and Parkinson’s Diseases: A Systematic Review of Preclinical Evidences. Oxidative Med. Cell. Longev. 2018, 2018, 1–21. [Google Scholar] [CrossRef]
  45. Wei, H.; Wu, G.; Chen, J.; Zhang, X.; Xiong, C.; Lei, Y.; Chen, W.; Ruan, J. (2S)-5, 2′, 5′-trihydroxy-7-methoxyflavanone, a natural product from abacopteris penangiana, presents neuroprotective effects in vitro and in vivo. Neurochem. Res. 2013, 38, 1686–1694. [Google Scholar] [CrossRef]
  46. Fjelldal, M.F.; Freyd, T.; Evenseth, L.M.; Sylte, I.; Ring, A.; Paulsen, R.E. Exploring the overlapping binding sites of ifenprodil and EVT -101 in GluN2B-containing NMDA receptors using novel chicken embryo forebrain cultures and molecular modeling. Pharmacol. Res. Perspect. 2019, 7, e00480. [Google Scholar] [CrossRef] [Green Version]
  47. Kumar, S.; Chowdhury, S.; Kumar, S. In silico repurposing of antipsychotic drugs for Alzheimer’s disease. Bmc Neurosci. 2017, 18, 76. [Google Scholar] [CrossRef]
  48. Thomas, S.J.; Grossberg, G.T. Memantine: A review of studies into its safety and efficacy in treating Alzheimer’s disease and other dementias. Clin. Interv. Aging 2009, 4, 367. [Google Scholar]
  49. Memantine for Treatment of Cognitive Impairment in Patients with Parkinson’s Disease and Dementia. Available online: https://clinicaltrials.gov/ct2/show/NCT00294554 (accessed on 14 April 2021).
  50. Remya, C.; Dileep, K.V.; Tintu, I.; Variyar, E.J.; Sadasivan, C. Design of potent inhibitors of acetylcholinesterase using morin as the starting compound. Front. Life Sci. 2012, 6, 107–117. [Google Scholar] [CrossRef]
  51. Du, Y.; Qu, J.; Zhang, W.; Bai, M.; Zhou, Q.; Zhang, Z.; Li, Z.; Miao, J. Morin reverses neuropathological and cognitive impairments in APPswe/PS1dE9 mice by targeting multiple pathogenic mechanisms. Neuropharmacology 2016, 108, 1–13. [Google Scholar] [CrossRef] [PubMed]
  52. Tadayon, M.; Garkani-Nejad, Z. In silico study combining QSAR, docking and molecular dynamics simulation on 2,4-disubstituted pyridopyrimidine derivatives. J. Recept. Signal Transduct. 2019, 39, 167–174. [Google Scholar] [CrossRef] [PubMed]
  53. Roy, K.K.; Tota, S.; Tripathi, T.; Chander, S.; Nath, C.; Saxena, A.K. Lead optimization studies towards the discovery of novel carbamates as potent AChE inhibitors for the potential treatment of Alzheimer’s disease. Bioorganic. Med. Chem. 2012, 20, 6313–6320. [Google Scholar] [CrossRef]
  54. Samadi, A.; Estrada, M.; Pérez, C.; Rodríguez-Franco, M.I.; Iriepa, I.; Moraleda, I.; Chioua, M.; Marco-Contelles, J. Pyridonepezils, new dual AChE inhibitors as potential drugs for the treatment of Alzheimer’s disease: Synthesis, biological assessment, and molecular modeling. Eur. J. Med. Chem. 2012, 57, 296–301. [Google Scholar] [CrossRef]
  55. Popugaeva, E.; Chernyuk, D.; Zhang, H.; Postnikova, T.Y.; Pats, K.; Fedorova, E.; Poroikov, V.; Zaitsev, A.V.; Bezprozvanny, I. Derivatives of Piperazines as potential therapeutic agents for Alzheimer’s disease. Mol. Pharmacol. 2019, 95, 337–348. [Google Scholar] [CrossRef] [Green Version]
  56. Varadaraju, K.R.; Kumar, J.R.; Mallesha, L.; Muruli, A.; Mohana, K.N.S.; Mukunda, C.K.; Sharanaiah, U. Virtual Screening and Biological Evaluation of Piperazine Derivatives as Human Acetylcholinesterase Inhibitors. Int. J. Alzheimers Dis. 2013, 2013, 653962. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, Y.; Lv, Y.; Jin, L.; Liang, G. Revealing the mechanism of EGCG, Genistein, Rutin, Quercetin, and Silibinin against hIAPP aggregation via computational simulations. Interdiscip. Sci. Comput. Life Sci. 2020, 12, 59–68. [Google Scholar] [CrossRef] [PubMed]
  58. Ramalingayya, G.V.; Cheruku, S.P.; Nayak, P.G.; Kishore, A.; Shenoy, R.; Rao, C.M.; Krishnadas, N. Rutin protects against neuronal damage in vitro and ameliorates doxorubicin-induced memory deficits in vivo in Wistar rats. Drug Des. Dev. Ther. 2017, 11, 1011. [Google Scholar] [CrossRef] [Green Version]
  59. Vancraenenbroeck, R.; De Raeymaecker, J.; Lobbestael, E.; Gao, F.; De Maeyer, M.; Voet, A.; Baekelandt, V.; Taymans, J.M. In silico, in vitro and cellular analysis with a kinome-wide inhibitor panel correlates cellular LRRK2 dephosphorylation to inhibitor activity on LRRK2. Front. Mol. Neurosci. 2014, 7, 51. [Google Scholar] [CrossRef]
  60. West, A.B. Achieving neuroprotection with LRRK2 kinase inhibitors in Parkinson disease. Exp. Neurol. 2017, 298, 236–245. [Google Scholar] [CrossRef]
  61. Padhi, A.K.; Banerjee, K.; Gomes, J.; Banerjee, M. Computational and Functional Characterization of Angiogenin Mutations, and Correlation with Amyotrophic Lateral Sclerosis. PLoS ONE 2014, 9, e111963. [Google Scholar] [CrossRef] [PubMed]
  62. Huang, H.-J.; Chang, T.-T.; Chen, H.-Y.; Chen, C.Y.-C. Finding inhibitors of mutant superoxide dismutase-1 for amyotrophic lateral sclerosis therapy from traditional chinese medicine. Evid. Based Complement. Altern. Med. 2014, 2014, 156276. [Google Scholar] [CrossRef]
  63. Bello, O.S.; Gonzalez, J.; Capani, F.; Barreto, G.E. In silico docking reveals possible Riluzole binding sites on Nav1. 6 sodium channel: Implications for amyotrophic lateral sclerosis therapy. J. Theor. Biol. 2012, 315, 53–63. [Google Scholar] [CrossRef]
  64. Benavides-Serrato, A.; Saunders, J.T.; Holmes, B.; Nishimura, R.N.; Lichtenstein, A.; Gera, J. Repurposing Potential of Riluzole as an ITAF Inhibitor in mTOR Therapy Resistant Glioblastoma. Int. J. Mol. Sci. 2020, 21, 344. [Google Scholar] [CrossRef] [Green Version]
  65. Huang, N.K.; Lin, J.H.; Lin, J.T.; Lin, C.I.; Liu, E.M.; Lin, C.J.; Chen, W.P.; Shen, Y.C.; Chen, H.M.; Chen, J.B.; et al. A new drug design targeting the adenosinergic system for Huntington’s disease. PLoS ONE 2011, 6, e20934. [Google Scholar] [CrossRef]
  66. Frydryskova, K.; Masek, T.; Pospisek, M. Changing faces of stress: Impact of heat and arsenite treatment on the composition of stress granules. Wiley Interdiscip. Rev. RNA 2020, e1596. [Google Scholar] [CrossRef]
  67. Hebert, L.E.; Weuve, J.; Scherr, P.A.; Evans, D.A. Alzheimer disease in the United States (2010–2050) estimated using the 2010 census. Neurology 2013, 80, 1778–1783. [Google Scholar] [CrossRef] [Green Version]
  68. Prince, M.; Knapp, M.; Guerchet, M.; McCrone, P.; Prina, M.; Comas-Herrera, A.; Wittenberg, R.; Adelaja, B.; Hu, B.; King, D. Dementia UK: Overview; Alzheimers Society: Belfast, UK, 2014. [Google Scholar]
  69. Mayrhofer, A.; Shora, S. Psychosocial Interventions for Younger People Diagnosed with Dementia: A Focus on Communities. In Proceedings of the Young Dementia Annual Conference 2019, London, UK, 20 November 2019; Available online: https://careinfo.org/event/young-dementia-2019/ (accessed on 28 April 2021).
  70. Nelson, P.T.; Alafuzoff, I.; Bigio, E.H.; Bouras, C.; Braak, H.; Cairns, N.J.; Castellani, R.J.; Crain, B.J.; Davies, P.; Tredici, K.D. Correlation of Alzheimer disease neuropathologic changes with cognitive status: A review of the literature. J. Neuropathol. Exp. Neurol. 2012, 71, 362–381. [Google Scholar] [CrossRef]
  71. Bellenguez, C.; Grenier-Boley, B.; Lambert, J.-C. Genetics of Alzheimer’s disease: Where we are, and where we are going. Curr. Opin. Neurobiol. 2020, 61, 40–48. [Google Scholar] [CrossRef]
  72. Bordone, M.P.; Salman, M.M.; Titus, H.E.; Amini, E.; Andersen, J.V.; Chakraborti, B.; Diuba, A.V.; Dubouskaya, T.G.; Ehrke, E.; Espindola de Freitas, A. The energetic brain–A review from students to students. J. Neurochem. 2019, 151, 139–165. [Google Scholar] [CrossRef]
  73. Nedergaard, M.; Goldman, S.A. Glymphatic failure as a final common pathway to dementia. Science 2020, 370, 50–56. [Google Scholar] [CrossRef]
  74. Kitchen, P.; Salman, M.M.; Halsey, A.M.; Clarke-Bland, C.; MacDonald, J.A.; Ishida, H.; Vogel, H.J.; Almutiri, S.; Logan, A.; Kreida, S. Targeting Aquaporin-4 Subcellular Localization to Treat Central Nervous System Edema. Cell 2020, 181, 784–799.e719. [Google Scholar] [CrossRef]
  75. Salman, M.M.; Sheilabi, M.A.; Bhattacharyya, D.; Kitchen, P.; Conner, A.C.; Bill, R.M.; Woodroofe, M.N.; Conner, M.T.; Princivalle, A.P. Transcriptome analysis suggests a role for the differential expression of cerebral aquaporins and the MAPK signalling pathway in human temporal lobe epilepsy. Eur. J. Neurosci. 2017, 46, 2121–2132. [Google Scholar] [CrossRef]
  76. Sylvain, N.J.; Salman, M.M.; Pushie, M.J.; Hou, H.; Meher, V.; Herlo, R.; Peeling, L.; Kelly, M.E. The effects of trifluoperazine on brain edema, aquaporin-4 expression, and metabolic markers during the acute phase of stroke using photothrombotic mouse model. Biochim. Et Biophys. Acta (BBA) Biomembr. 2021, 1863, 183573. [Google Scholar] [CrossRef] [PubMed]
  77. Abir-Awan, M.; Kitchen, P.; Salman, M.M.; Conner, M.T.; Conner, A.C.; Bill, R.M. Inhibitors of mammalian aquaporin water channels. Int. J. Mol. Sci. 2019, 20, 1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kitchen, P.; Day, R.E.; Taylor, L.H.; Salman, M.M.; Bill, R.M.; Conner, M.T.; Conner, A.C. Identification and molecular mechanisms of the rapid tonicity-induced relocalization of the aquaporin 4 channel. J. Biol. Chem. 2015, 290, 16873–16881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Salman, M.M.; Kitchen, P.; Woodroofe, M.N.; Brown, J.E.; Bill, R.M.; Conner, A.C.; Conner, M.T. Hypothermia increases aquaporin 4 (AQP4) plasma membrane abundance in human primary cortical astrocytes via a calcium/transient receptor potential vanilloid 4 (TRPV4)- and calmodulin-mediated mechanism. Eur. J. Neurosci. 2017, 46, 2542–2547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Akincioglu, H.; Gulcin, I. Potent Acetylcholinesterase Inhibitors: Potential Drugs for Alzheimer’s Disease. Mini Rev. Med. Chem. 2020. [Google Scholar] [CrossRef]
  81. Tang, H.; Song, P.; Li, J.; Zhao, D. Effect of Salvia miltiorrhiza on acetylcholinesterase: Enzyme kinetics and interaction mechanism merging with molecular docking analysis. Int. J. Biol. Macromol. 2019, 135, 303–313. [Google Scholar] [CrossRef] [PubMed]
  82. Ram, H.; Jaipal, N.; Kumar, P.; Deka, P.; Kumar, S.; Kashyap, P.; Kumar, S.; Singh, B.P.; Alqarawi, A.A.; Hashem, A.; et al. Dual Inhibition of DPP-4 and Cholinesterase Enzymes by the Phytoconstituents of the Ethanolic Extract of Prosopis cineraria Pods: Therapeutic Implications for the Treatment of Diabetes-associated Neurological Impairments. Curr. Alzheimer Res. 2019, 16, 1230–1244. [Google Scholar] [CrossRef] [PubMed]
  83. Mazumder, M.K.; Choudhury, S. Tea polyphenols as multi-target therapeutics for Alzheimer’s disease: An in silico study. Med. Hypotheses 2019, 125, 94–99. [Google Scholar] [CrossRef] [PubMed]
  84. Ali, M.A.; Vuree, S.; Goud, H.; Hussain, T.; Nayarisseri, A.; Singh, S.K. Identification of High-affinity Small Molecules Targeting Gamma Secretase for the Treatment of Alzheimer’s Disease. Curr. Top. Med. Chem. 2019, 19, 1173–1187. [Google Scholar] [CrossRef] [PubMed]
  85. D’Amelio, M.; Sheng, M.; Cecconi, F. Caspase-3 in the central nervous system: Beyond apoptosis. Trends Neurosci. 2012, 35, 700–709. [Google Scholar] [CrossRef] [PubMed]
  86. Imbriani, P.; Tassone, A.; Meringolo, M.; Ponterio, G.; Madeo, G.; Pisani, A.; Bonsi, P.; Martella, G. Loss of Non-Apoptotic Role of Caspase-3 in the PINK1 Mouse Model of Parkinson’s Disease. Int. J. Mol. Sci. 2019, 20, 3407. [Google Scholar] [CrossRef] [Green Version]
  87. Kavanagh, E.; Rodhe, J.; Burguillos, M.A.; Venero, J.L.; Joseph, B. Regulation of caspase-3 processing by cIAP2 controls the switch between pro-inflammatory activation and cell death in microglia. Cell Death Dis. 2014, 5, e1565. [Google Scholar] [CrossRef] [Green Version]
  88. Snigdha, S.; Smith, E.D.; Prieto, G.A.; Cotman, C.W. Caspase-3 activation as a bifurcation poInt. between plasticity and cell death. Neurosci. Bull. 2012, 28, 14–24. [Google Scholar] [CrossRef]
  89. Cancela, S.; Canclini, L.; Mourglia-Ettlin, G.; Hernández, P.; Merlino, A. Neuroprotective effects of novel nitrones: In vitro and in silico studies. Eur. J. Pharmacol. 2020, 871, 172926. [Google Scholar] [CrossRef]
  90. Goubau, C.; Jaeken, J.; Levtchenko, E.N.; Thys, C.; Di Michele, M.; Martens, G.A.; Gerlo, E.; De Vos, R.; Buyse, G.M.; Goemans, N.; et al. Homozygosity for aquaporin 7 G264V in three unrelated children with hyperglyceroluria and a mild platelet secretion defect. Genet Med. 2013, 15, 55–63. [Google Scholar] [CrossRef] [Green Version]
  91. Greig, N.H.; Reale, M.; Tata, A.M. New pharmacological approaches to the cholinergic system: An overview on muscarinic receptor ligands and cholinesterase inhibitors. Recent Pat. CNS Drug Discov. 2013, 8, 123–141. [Google Scholar] [CrossRef] [PubMed]
  92. Volpato, D.; Holzgrabe, U. Designing hybrids targeting the cholinergic system by modulating the muscarinic and nicotinic receptors: A concept to treat Alzheimer’s disease. Molecules 2018, 23, 3230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Koch, A.; Bonus, M.; Gohlke, H.; Klöcker, N. Isoform-specific Inhibition of N-methyl-D-aspartate Receptors by Bile Salts. Sci. Rep. 2019, 9, 1–17. [Google Scholar] [CrossRef] [PubMed]
  94. Song, X.; Jensen, M.O.; Jogini, V.; Stein, R.A.; Lee, C.H.; McHaourab, H.S.; Shaw, D.E.; Gouaux, E. Mechanism of NMDA receptor channel block by MK-801 and memantine. Nature 2018, 556, 515–519. [Google Scholar] [CrossRef]
  95. Geerts, H.; Grossberg, G.T. Pharmacology of acetylcholinesterase inhibitors and N-methyl-D-aspartate receptors for combination therapy in the treatment of Alzheimer’s disease. J. Clin. Pharmacol. 2006, 46, 8s–16s. [Google Scholar] [CrossRef]
  96. Singh, R.; Ganeshpurkar, A.; Kumar, D.; Kumar, D.; Kumar, A.; Singh, S.K. Identifying potential GluN2B subunit containing N-Methyl-D-aspartate receptor inhibitors: An integrative in silico and molecular modeling approach. J. Biomol. Struct. Dyn. 2019, 1–13. [Google Scholar] [CrossRef] [PubMed]
  97. Waqar, M.; Batool, S. In silico analysis of binding interaction of conantokins with NMDA receptors for potential therapeutic use in Alzheimer’s disease. J. Venom. Anim. Toxins Incl. Trop. Dis. 2017, 23, 42. [Google Scholar] [CrossRef] [Green Version]
  98. Alokam, R.; Singhal, S.; Srivathsav, G.S.; Garigipati, S.; Puppala, S.; Sriram, D.; Perumal, Y. Design of dual inhibitors of ROCK-I and NOX2 as potential leads for the treatment of neuroinflammation associated with various neurological diseases including autism spectrum disorder. Mol. Biosyst. 2015, 11, 607–617. [Google Scholar] [CrossRef]
  99. Simpson, D.S.; Oliver, P.L. ROS generation in microglia: Understanding oxidative stress and inflammation in neurodegenerative disease. Antioxidants 2020, 9, 743. [Google Scholar] [CrossRef]
  100. Sturchio, A.; Marsili, L.; Mahajan, A.; Grimberg, B.; Kauffman, M.A.; Espay, A.J. How Have Advances in Genetic Technology Modified Movement Disorders Nosology? Eur. J. Neurol. 2020. [Google Scholar] [CrossRef]
  101. Paul, D.A.; Qureshi, A.R.M.; Rana, A.Q. Peripheral neuropathy in Parkinson’s disease. Neurol. Sci. 2020. [Google Scholar] [CrossRef]
  102. Coundouris, S.P.; Terrett, G.; Laakso, L.; Schweitzer, D.; Kneebone, A.; Rendell, P.G.; Henry, J.D. A meta-analytic review of prospection deficits in Parkinson’s disease. Neurosci. Biobehav. Rev. 2020, 108, 34–47. [Google Scholar] [CrossRef]
  103. Kelly, J.; Moyeed, R.; Carroll, C.; Albani, D.; Li, X. Gene expression meta-analysis of Parkinson’s disease and its relationship with Alzheimer’s disease. Mol. Brain 2019, 12, 16. [Google Scholar] [CrossRef] [PubMed]
  104. Henderson, E.J.; Lyell, V.; Bhimjiyani, A.; Amin, J.; Kobylecki, C.; Gregson, C.L. Management of fracture risk in Parkinson’s: A revised algorithm and focused review of treatments. Parkinsonism Relat. Disord. 2019, 64, 181–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Haunton, V.J. Movement disorders: A themed collection. Age Ageing 2020, 49, 12–15. [Google Scholar] [CrossRef]
  106. Félix, J.P.; Vieira, F.H.T.; Cardoso, Á.A.; Ferreira, M.V.G.; Franco, R.A.P.; Ribeiro, M.A.; Araújo, S.G.; Corrêa, H.P.; Carneiro, M.L. A Parkinson’s Disease Classification Method: An Approach Using Gait Dynamics and Detrended Fluctuation Analysis. In Proceedings of the 2019 IEEE Canadian Conference of Electrical and Computer Engineering (CCECE), Edmonton, AB, Canada, 5–8 May 2019; pp. 1–4. [Google Scholar]
  107. Chakraborty, A.; Brauer, S.; Diwan, A. A review of possible therapies for Parkinson’s disease. J. Clin. Neurosci. 2020. [Google Scholar] [CrossRef]
  108. Ma, Z.; Liu, H.; Wu, B. Structure-based drug design of catechol-O-methyltransferase inhibitors for CNS disorders. Br. J. Clin. Pharmacol. 2014, 77, 410–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Borovac, J.A. Focus: The Aging Brain: Side effects of a dopamine agonist therapy for Parkinson’s disease: A mini-review of clinical pharmacology. Yale J. Biol. Med. 2016, 89, 37. [Google Scholar]
  110. Thobois, S. Proposed dose equivalence for rapid switch between dopamine receptor agonists in Parkinson’s disease: A review of the literature. Clin. Ther. 2006, 28, 1–12. [Google Scholar] [CrossRef]
  111. Missale, C.; Nash, S.R.; Robinson, S.W.; Jaber, M.; Caron, M.G. Dopamine receptors: From structure to function. Physiol. Rev. 1998, 78, 189–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Le Foll, B.; Collo, G.; Rabiner, E.A.; Boileau, I.; Pich, E.M.; Sokoloff, P. Dopamine D3 receptor ligands for drug addiction treatment: Update on recent findings. In Progress in Brain Research; Elsevier: Amsterdam, The Netherlands, 2014; Volume 211, pp. 255–275. [Google Scholar]
  113. Maggio, R.; Aloisi, G.; Silvano, E.; Rossi, M.; Millan, M.J. Heterodimerization of dopamine receptors: New insights into functional and therapeutic significance. Parkinsonism Relat. Disord. 2009, 15, S2–S7. [Google Scholar] [CrossRef]
  114. Connor-Robson, N.; Booth, H.; Martin, J.G.; Gao, B.; Li, K.; Doig, N.; Vowles, J.; Browne, C.; Klinger, L.; Juhasz, P. An integrated transcriptomics and proteomics analysis reveals functional endocytic dysregulation caused by mutations in LRRK2. Neurobiol. Dis. 2019, 127, 512–526. [Google Scholar] [CrossRef] [PubMed]
  115. Jin, W. Novel Insights into PARK7 (DJ-1), a Potential Anti-Cancer Therapeutic Target, and Implications for Cancer Progression. J. Clin. Med. 2020, 9, 1256. [Google Scholar] [CrossRef] [PubMed]
  116. Bartonikova, T.; Mensikova, K.; Kolarikova, K.; Vodicka, R.; Vrtel, R.; Otruba, P.; Kaiserova, M.; Vastik, M.; Mikulicova, L.; Ovecka, J.; et al. New endemic familial parkinsonism in south Moravia, Czech Republic and its genetical background. Medicine 2018, 97, e12313. [Google Scholar] [CrossRef]
  117. Schmidt, S.H.; Knape, M.J.; Boassa, D.; Mumdey, N.; Kornev, A.P.; Ellisman, M.H.; Taylor, S.S.; Herberg, F.W. The dynamic switch mechanism that leads to activation of LRRK2 is embedded in the DFGpsi motif in the kinase domain. Proc. Natl. Acad. Sci. USA 2019, 116, 14979–14988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Nagarajan, N.; Chellam, J.; Kannan, R.R. Exploring the functional impact of mutational drift in LRRK2 gene and identification of specific inhibitors for the treatment of Parkinson disease. J. Cell. Biochem. 2018, 119, 4878–4889. [Google Scholar] [CrossRef] [PubMed]
  119. Flor, P.J.; Battaglia, G.; Nicoletti, F.; Gasparini, F.; Bruno, V. Neuroprotective activity of metabotropic glutamate receptor ligands. In Molecular and Cellular Biology of Neuroprotection in the CNS; Springer: Berlin/Heidelberg, Germany, 2003; pp. 197–223. [Google Scholar]
  120. Montastruc, J.; Rascol, O.; Senard, J. Glutamate antagonists and Parkinson’s disease: A review of clinical data. Neurosci. Biobehav. Rev. 1997, 21, 477–480. [Google Scholar] [CrossRef]
  121. Marzo, A.; Dal Bo, L.; Monti, N.C.; Crivelli, F.; Ismaili, S.; Caccia, C.; Cattaneo, C.; Fariello, R.G. Pharmacokinetics and pharmacodynamics of safinamide, a neuroprotectant with antiparkinsonian and anticonvulsant activity. Pharmacol. Res. 2004, 50, 77–85. [Google Scholar] [CrossRef]
  122. Stocchi, F.; Arnold, G.; Onofrj, M.; Kwiecinski, H.; Szczudlik, A.; Thomas, A.; Bonuccelli, U.; Van Dijk, A.; Cattaneo, C.; Sala, P. Improvement of motor function in early Parkinson disease by safinamide. Neurology 2004, 63, 746–748. [Google Scholar] [CrossRef]
  123. Wijesekera, L.C.; Leigh, P.N. Amyotrophic lateral sclerosis. Orphanet J. Rare Dis. 2009, 4, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Brownlee, A.; Palovcak, M. The role of augmentative communication devices in the medical management of ALS. NeuroRehabilitation 2007, 22, 445–450. [Google Scholar] [CrossRef]
  125. Martin, L.J.; Liu, Z. Opportunities for neuroprotection in ALS using cell death mechanism rationales. Drug Discov. Today Dis. Models 2004, 1, 135–143. [Google Scholar] [CrossRef]
  126. Aebischer, P.; Kato, A.C. Playing defense against Lou Gehrig’s disease. Sci. Am. 2007, 297, 86–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Furukawa, Y.; Torres, A.S.; O’Halloran, T.V. Oxygen-induced maturation of SOD1: A key role for disulfide formation by the copper chaperone CCS. EMBO J. 2004, 23, 2872–2881. [Google Scholar] [CrossRef] [PubMed]
  128. Hirayama, K.; Fujiwara, Y.; Terada, T.; Shimizu, K.; Wada, K.; Kabuta, T. Virtual screening identification of novel chemical inhibitors for aberrant interactions between pathogenic mutant SOD1 and tubulin. Neurochem. Int. 2019, 126, 19–26. [Google Scholar] [CrossRef] [PubMed]
  129. Srinivasan, E.; Rajasekaran, R. Rational design of linear tripeptides against the aggregation of human mutant SOD1 protein causing amyotrophic lateral sclerosis. J. Neurol. Sci. 2019, 405, 116425. [Google Scholar] [CrossRef] [PubMed]
  130. Srinivasan, E.; Rajasekaran, R. Quantum chemical and molecular mechanics studies on the assessment of interactions between resveratrol and mutant SOD1 (G93A) protein. J. Comput. Aided Mol. Des. 2018, 32, 1347–1361. [Google Scholar] [CrossRef] [PubMed]
  131. Srinivasan, E.; Rajasekaran, R. Computational investigation of curcumin, a natural polyphenol that inhibits the destabilization and the aggregation of human SOD1 mutant (Ala4Val). RSC Adv. 2016, 6, 102744–102753. [Google Scholar] [CrossRef]
  132. Srinivasan, E.; Rajasekaran, R. Comparative binding of kaempferol and kaempferide on inhibiting the aggregate formation of mutant (G85R) SOD1 protein in familial amyotrophic lateral sclerosis: A quantum chemical and molecular mechanics study. Biofactors 2018, 44, 431–442. [Google Scholar] [CrossRef]
  133. Zhang, Y.Y.; Mei, Z.Q.; Wu, J.W.; Wang, Z.X. Enzymatic activity and substrate specificity of mitogen-activated protein kinase p38alpha in different phosphorylation states. J. Biol. Chem. 2008, 283, 26591–26601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Thalhamer, T.; McGrath, M.; Harnett, M. MAPKs and their relevance to arthritis and inflammation. Rheumatology 2008, 47, 409–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Chico, L.K.; Van Eldik, L.J.; Watterson, D.M. Targeting protein kinases in central nervous system disorders. Nat. Rev. Drug Discov. 2009, 8, 892–909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Fernandes, V.; Sharma, D.; Kalia, K.; Tiwari, V. Neuroprotective effects of silibinin: An in silico and in vitro study. Int. J. Neurosci. 2018, 128, 935–945. [Google Scholar] [CrossRef]
  137. Laev, S.S.; Salakhutdinov, N.F.; Lavrik, O.I. Tyrosyl-DNA phosphodiesterase inhibitors: Progress and potential. Bioorganic. Med. Chem. 2016, 24, 5017–5027. [Google Scholar] [CrossRef]
  138. Salado, I.G.; Redondo, M.; Bello, M.L.; Perez, C.; Liachko, N.F.; Kraemer, B.C.; Miguel, L.; Lecourtois, M.; Gil, C.; Martinez, A.; et al. Protein Kinase CK-1 Inhibitors As New Potential Drugs for Amyotrophic Lateral Sclerosis. J. Med. Chem. 2014, 57, 2755–2772. [Google Scholar] [CrossRef]
  139. Bissaro, M.; Moro, S. Rethinking to riluzole mechanism of action: The molecular link among protein kinase CK1δ activity, TDP-43 phosphorylation, and amyotrophic lateral sclerosis pharmacological treatment. Neural. Regen. Res. 2019, 14, 2083–2085. [Google Scholar] [CrossRef] [PubMed]
  140. Cozza, G.; Gianoncelli, A.; Montopoli, M.; Caparrotta, L.; Venerando, A.; Meggio, F.; Pinna, L.A.; Zagotto, G.; Moro, S. Identification of novel protein kinase CK1 delta (CK1δ) inhibitors through structure-based virtual screening. Bioorganic. Med. Chem. Lett. 2008, 18, 5672–5675. [Google Scholar] [CrossRef]
  141. Cozza, G.; Pinna, L.A.; Moro, S. Kinase CK2 inhibition: An update. Curr. Med. Chem. 2013, 20, 671–693. [Google Scholar] [CrossRef]
  142. Kearney, J.A.; Buchner, D.A.; De Haan, G.; Adamska, M.; Levin, S.I.; Furay, A.R.; Albin, R.L.; Jones, J.M.; Montal, M.; Stevens, M.J. Molecular and pathological effects of a modifier gene on deficiency of the sodium channel Scn8a (Nav1. 6). Hum. Mol. Genet. 2002, 11, 2765–2775. [Google Scholar] [CrossRef] [Green Version]
  143. Caldwell, J.H.; Schaller, K.L.; Lasher, R.S.; Peles, E.; Levinson, S.R. Sodium channel Nav1. 6 is localized at nodes of Ranvier, dendrites, and synapses. Proc. Natl. Acad. Sci. USA 2000, 97, 5616–5620. [Google Scholar] [CrossRef] [Green Version]
  144. Gunasekaran, R.; Narayani, R.S.; Vijayalakshmi, K.; Alladi, P.A.; Shobha, K.; Nalini, A.; Sathyaprabha, T.; Raju, T. Exposure to cerebrospinal fluid of sporadic amyotrophic lateral sclerosis patients alters Nav1. 6 and Kv1. 6 channel expression in rat spinal motor neurons. Brain Res. 2009, 1255, 170–179. [Google Scholar] [CrossRef] [PubMed]
  145. Saba, L.; Viscomi, M.T.; Martini, A.; Caioli, S.; Mercuri, N.B.; Guatteo, E.; Zona, C. Modified age-dependent expression of NaV1. 6 in an ALS model correlates with motor cortex excitability alterations. Neurobiol. Dis. 2019, 130, 104532. [Google Scholar] [CrossRef]
  146. Seki, S.; Yamamoto, T.; Quinn, K.; Spigelman, I.; Pantazis, A.; Olcese, R.; Wiedau-Pazos, M.; Chandler, S.H.; Venugopal, S. Circuit-Specific Early Impairment of Proprioceptive Sensory Neurons in the SOD1G93A Mouse Model for ALS. J. Neurosci. 2019, 39, 8798–8815. [Google Scholar] [CrossRef] [PubMed]
  147. Grant, P.; Song, J.Y.; Swedo, S.E. Review of the use of the glutamate antagonist riluzole in psychiatric disorders and a description of recent use in childhood obsessive-compulsive disorder. J. Child Adolesc. Psychopharmacol. 2010, 20, 309–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Anzai, I.; Toichi, K.; Tokuda, E.; Mukaiyama, A.; Akiyama, S.; Furukawa, Y. Screening of drugs inhibiting in vitro oligomerization of Cu/Zn-superoxide dismutase with a mutation causing amyotrophic lateral sclerosis. Front. Mol. Biosci. 2016, 3, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Martin, J.B.; Gusella, J.F. Huntingtons disease. N. Engl. J. Med. 1986, 315, 1267–1276. [Google Scholar] [CrossRef] [PubMed]
  150. Novak, M.J.; Tabrizi, S.J. Huntington’s disease. BMJ 2010, 340, c3109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Wexler, N.S.; Collett, L.; Wexler, A.R.; Rawlins, M.D.; Tabrizi, S.J.; Douglas, I.; Smeeth, L.; Evans, S.J. Incidence of adult Huntington’s disease in the UK: A UK-based primary care study and a systematic review. BMJ Open 2016, 6, e009070. [Google Scholar] [CrossRef]
  152. Kavanaugh, M.S.; Cho, C.; Maeda, H.; Swope, C. “I am no longer alone”: Evaluation of the first North American camp for youth living in families with Huntington’s disease. Child. Youth Serv. Rev. 2017, 79, 325–332. [Google Scholar] [CrossRef]
  153. Ayala, K.M. Role of the SLP in management of Huntington’s disease: A literature review. 2017. Available online: https://repositories.lib.utexas.edu/handle/2152/62346 (accessed on 14 April 2021). [CrossRef]
  154. Travessa, A.M.; Rodrigues, F.B.; Mestre, T.A.; Ferreira, J.J. Fifteen years of clinical trials in Huntington’s disease: A very low clinical drug development success rate. J. Huntingt. Dis. 2017, 6, 157–163. [Google Scholar] [CrossRef] [PubMed]
  155. Tabrizi, S. Targeting Huntingtin Expression in Patients with Huntington’s Disease. N. Engl. J. Med. 2019, 380, 2307–2316. [Google Scholar] [CrossRef]
  156. Iftikhar, H.; Batool, S.; Deep, A.; Narasimhan, B.; Sharma, P.C.; Malhotra, M. In silico analysis of the inhibitory activities of GABA derivatives on 4-aminobutyrate transaminase. Arab. J. Chem. 2017, 10, S1267–S1275. [Google Scholar] [CrossRef] [Green Version]
  157. Boonstra, E.; de Kleijn, R.; Colzato, L.S.; Alkemade, A.; Forstmann, B.U.; Nieuwenhuis, S. Neurotransmitters as food supplements: The effects of GABA on brain and behavior. Front. Psychol. 2015, 6, 1520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Vijayakumar, S.; Kasthuri, G.; Prabhu, S.; Manogar, P.; Parameswari, N. Screening and identification of novel inhibitors against human 4-aminobutyrate-aminotransferase: A computational approach. Egypt J. Basic Appl. Sci. 2018, 5, 210–219. [Google Scholar] [CrossRef]
  159. Schneider, G. Virtual screening: An endless staircase? Nat. Rev. Drug Discov. 2010, 9, 273–276. [Google Scholar] [CrossRef]
  160. Verkhivker, G.M.; Bouzida, D.; Gehlhaar, D.K.; Rejto, P.A.; Arthurs, S.; Colson, A.B.; Freer, S.T.; Larson, V.; Luty, B.A.; Marrone, T. Deciphering common failures in molecular docking of ligand-protein complexes. J. Comput. Aided Mol. Des. 2000, 14, 731–751. [Google Scholar] [CrossRef]
  161. Josephs, D.; Spicer, J.; O’Doherty, M. Molecular imaging in clinical trials. Target. Oncol. 2009, 4, 151–168. [Google Scholar] [CrossRef]
  162. Cheatham, T.E., III; Young, M.A. Molecular dynamics simulation of nucleic acids: Successes, limitations, and promise. Biopolym. Orig. Res. Biomol. 2000, 56, 232–256. [Google Scholar] [CrossRef]
  163. Klebe, G. Virtual ligand screening: Strategies, perspectives and limitations. Drug Discov. Today 2006, 11, 580–594. [Google Scholar] [CrossRef]
  164. Ferreira, L.G.; Dos Santos, R.N.; Oliva, G.; Andricopulo, A.D. Molecular docking and structure-based drug design strategies. Molecules 2015, 20, 13384–13421. [Google Scholar] [CrossRef] [PubMed]
  165. Korb, O.; Olsson, T.S.; Bowden, S.J.; Hall, R.J.; Verdonk, M.L.; Liebeschuetz, J.W.; Cole, J.C. Potential and limitations of ensemble docking. J. Chem. Inf. Modeling 2012, 52, 1262–1274. [Google Scholar] [CrossRef]
  166. MacDonald, D.; Breton, R.; Sutcliffe, R.; Walker, J. Uses and limitations of quantitative structure-activity relationships (QSARs) to categorize substances on the Canadian Domestic Substance List as persistent and/or bioaccumulative, and inherently toxic to non-human organisms. Sar Qsar Environ. Res. 2002, 13, 43–55. [Google Scholar] [CrossRef] [PubMed]
  167. Papakonstantinou, E.; Megalooikonomou, V.; Vlachakis, D. Dark Suite: A comprehensive toolbox for computer-aided drug design. Embnet. J. 2020, 25, e928. [Google Scholar] [CrossRef]
  168. Berdigaliyev, N.; Aljofan, M. An overview of drug discovery and development. Future Med. Chem. 2020, 12, 939–947. [Google Scholar] [CrossRef]
  169. Shi, J.; Zha, W. Predicting Human Pharmacokinetics: Physiologically Based Pharmacokinetic Modeling and In Silico ADME Prediction in Early Drug Discovery. Eur. J. Drug Metab. Pharmacokinet. 2019, 44, 135–137. [Google Scholar] [CrossRef]
  170. Van De Waterbeemd, H.; Gifford, E. ADMET in silico modelling: Towards prediction paradise? Nat. Rev. Drug Discov. 2003, 2, 192–204. [Google Scholar] [CrossRef]
  171. Faller, B.; Winiwarter, S.; Chang, G.; Desai, P.; Menzel, C.; Rieko, A.; Keefer, C.; Broccatelli, F. Prediction of fraction unbound in microsomal and hepatocyte incubations–a comparison of methods across industry data sets (by the IQ in silico ADME working group). Mol. Pharm. 2019, 16, 4077–4085. [Google Scholar]
  172. Blomme, E.A.; Will, Y. Toxicology strategies for drug discovery: Present and future. Chem. Res. Toxicol. 2016, 29, 473–504. [Google Scholar] [CrossRef] [PubMed]
  173. Pruss, R.M. Phenotypic screening strategies for neurodegenerative diseases: A pathway to discover novel drug candidates and potential disease targets or mechanisms. CNS Neurol. Disord. Drug Targets Former. Curr. Drug Targets CNS Neurol. Disord. 2010, 9, 693–700. [Google Scholar] [CrossRef]
  174. Ren, J.; Wen, L.; Gao, X.; Jin, C.; Xue, Y.; Yao, X. CSS-Palm 2.0: An updated software for palmitoylation sites prediction. Protein Eng. Des. Sel. 2008, 21, 639–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Nikzad, H.; Karimian, M.; Sareban, K.; Khoshsokhan, M.; Colagar, A.H. MTHFR-Ala222Val and male infertility: A study in Iranian men, an updated meta-analysis and an in silico-analysis. Reprod. Biomed. Online 2015, 31, 668–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Glykos, N.M. Software news and updates carma: A molecular dynamics analysis program. J. Comput. Chem. 2006, 27, 1765–1768. [Google Scholar] [CrossRef] [PubMed]
  177. CCDC. What’s New. Available online: https://www.ccdc.cam.ac.uk/solutions/whats-new/ (accessed on 14 April 2021).
  178. Zuo, Z.; MacMillan, D.W. Decarboxylative arylation of α-amino acids via photoredox catalysis: A one-step conversion of biomass to drug pharmacophore. J. Am. Chem. Soc. 2014, 136, 5257–5260. [Google Scholar] [CrossRef]
  179. Vázquez, K.; Espinosa-Bustos, C.; Soto-Delgado, J.; Tapia, R.A.; Varela, J.; Birriel, E.; Segura, R.; Pizarro, J.; Cerecetto, H.; González, M. New aryloxy-quinone derivatives as potential anti-Chagasic agents: Synthesis, trypanosomicidal activity, electrochemical properties, pharmacophore elucidation and 3D-QSAR analysis. RSC Adv. 2015, 5, 65153–65166. [Google Scholar] [CrossRef]
  180. Bennett, B.C.; Wan, Q.; Ahmad, M.F.; Langan, P.; Dealwis, C.G. X-ray structure of the ternary MTX NADPH complex of the anthrax dihydrofolate reductase: A pharmacophore for dual-site inhibitor design. J. Struct. Biol. 2009, 166, 162–171. [Google Scholar] [CrossRef]
  181. François, P.; Hakim, V. Design of genetic networks with specified functions by evolution in silico. Proc. Natl. Acad. Sci. USA 2004, 101, 580–585. [Google Scholar] [CrossRef] [Green Version]
  182. Fischer, B.; Fukuzawa, K.; Wenzel, W. Receptor-specific scoring functions derived from quantum chemical models improve affinity estimates for in-silico drug discovery. Proteins Struct. Funct. Bioinform. 2008, 70, 1264–1273. [Google Scholar] [CrossRef]
  183. Khatami, S.G.; Mubeen, S.; Hofmann-Apitius, M. Data science in neurodegenerative disease: Its capabilities, limitations, and perspectives. Curr. Opin. Neurol. 2020, 33, 249. [Google Scholar] [CrossRef]
  184. Ferro, M.P.; Heilshorn, S.C.; Owens, R.M. Materials for blood brain barrier modeling in vitro. Mater. Sci. Eng. R Rep. 2020, 140, 100522. [Google Scholar] [CrossRef]
  185. May, J.-N.; Golombek, S.K.; Baues, M.; Dasgupta, A.; Drude, N.; Rix, A.; Rommel, D.; von Stillfried, S.; Appold, L.; Pola, R. Multimodal and multiscale optical imaging of nanomedicine delivery across the blood-brain barrier upon sonopermeation. Theranostics 2020, 10, 1948. [Google Scholar] [CrossRef]
  186. Juthani, R.; Madajewski, B.; Yoo, B.; Zhang, L.; Chen, P.-M.; Chen, F.; Turker, M.Z.; Ma, K.; Overholtzer, M.; Longo, V.A. Ultrasmall Core-Shell Silica Nanoparticles for Precision Drug Delivery in a High-Grade Malignant Brain Tumor Model. Clin. Cancer Res. 2020, 26, 147–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Blauwendraat, C.; Heilbron, K.; Vallerga, C.L.; Bandres-Ciga, S.; von Coelln, R.; Pihlstrøm, L.; Simón-Sánchez, J.; Schulte, C.; Sharma, M.; Krohn, L. Parkinson’s disease age at onset genome-wide association study: Defining heritability, genetic loci, and α-synuclein mechanisms. Mov. Disord. 2019, 34, 866–875. [Google Scholar] [CrossRef] [Green Version]
  188. Nalls, M.A.; Blauwendraat, C.; Vallerga, C.L.; Heilbron, K.; Bandres-Ciga, S.; Chang, D.; Tan, M.; Kia, D.A.; Noyce, A.J.; Xue, A. Identification of novel risk loci, causal insights, and heritable risk for Parkinson’s disease: A meta-analysis of genome-wide association studies. Lancet Neurol. 2019, 18, 1091–1102. [Google Scholar] [CrossRef]
  189. Ciani, M.; Benussi, L.; Bonvicini, C.; Ghidoni, R. Genome Wide Association Study and Next Generation Sequencing: A glimmer of light towards new possible horizons in Frontotemporal Dementia research. Front. Neurosci. 2019, 13, 506. [Google Scholar] [CrossRef] [PubMed]
  190. Ibanez, L.; Farias, F.H.; Dube, U.; Mihindukulasuriya, K.A.; Harari, O. Polygenic risk scores in neurodegenerative diseases: A review. Curr. Genet. Med. Rep. 2019, 7, 22–29. [Google Scholar] [CrossRef] [Green Version]
  191. Adams, H.H.; Evans, T.E.; Terzikhan, N. The Uncovering Neurodegenerative Insights Through Ethnic Diversity consortium. Lancet Neurol. 2019, 18, 915. [Google Scholar] [CrossRef] [Green Version]
  192. Mancuso, R.; Fryatt, G.; Cleal, M.E.; Obst, J.; Pipi, E.; Monzon-Sandoval, J.; Winchester, L.; Ribe, E.; Webber, C.; Nevado, A. CSF1R inhibition by JNJ-40346527 alters microglial proliferation and phenotype and results in attenuation of neurodegeneration in P301S mice. Brain 2019, 142, 3243–3264. [Google Scholar] [CrossRef]
  193. Rahman, S.; Datta, M.; Kim, J.; Jan, A.T. CRISPR/Cas: An intriguing genomic editing tool with prospects in treating neurodegenerative diseases. Semin. Cell Dev. Biol. 2019, 96, 22–31. [Google Scholar] [CrossRef]
  194. Offen, D.; Perets, N.; Betzer, O.; Popovtzer, R.; Shapira, R.; Ashery, U. Targeting damages in the brain: Exosomes derived from MSC present migration and homing abilities to different neurodegenerative and neuropsychiatric locations. Cytotherapy 2019, 21, e6. [Google Scholar] [CrossRef]
  195. Raikwar, S.P.; Kikkeri, N.S.; Sakuru, R.; Saeed, D.; Zahoor, H.; Premkumar, K.; Mentor, S.; Thangavel, R.; Dubova, I.; Ahmed, M.E. Next generation precision medicine: CRISPR-mediated genome editing for the treatment of neurodegenerative disorders. J. Neuroimmune Pharmacol. 2019, 14, 608–641. [Google Scholar] [CrossRef]
  196. Aldewachi, H.; Al-Zidan, R.N.; Conner, M.T.; Salman, M.M. High-Throughput Screening Platforms in the Discovery of Novel Drugs for Neurodegenerative Diseases. Bioengineering 2021, 8, 30. [Google Scholar] [CrossRef] [PubMed]
  197. Mittal, R.; Woo, F.W.; Castro, C.S.; Cohen, M.A.; Karanxha, J.; Mittal, J.; Chhibber, T.; Jhaveri, V.M. Organ-on-chip models: Implications in drug discovery and clinical applications. J. Cell. Physiol. 2019, 234, 8352–8380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Salman, M.M.; Marsh, G.; Kusters, I.; Delincé, M.; Di Caprio, G.; Upadhyayula, S.; de Nola, G.; Hunt, R.; Ohashi, K.G.; Gray, T.; et al. Design and Validation of a Human Brain Endothelial Microvessel-on-a-Chip Open Microfluidic Model Enabling Advanced Optical Imaging. Front. Bioeng. Biotechnol. 2020, 8. [Google Scholar] [CrossRef] [PubMed]
  199. Chen, J.J. Functional MRI of brain physiology in aging and neurodegenerative diseases. Neuroimage 2019, 187, 209–225. [Google Scholar] [CrossRef] [PubMed]
  200. Muir, E.R.; Biju, K.; Cong, L.; Rogers, W.E.; Hernandez, E.T.; Duong, T.Q.; Clark, R.A. Functional MRI of the mouse olfactory system. Neurosci. Lett. 2019, 704, 57–61. [Google Scholar] [CrossRef] [PubMed]
  201. Yan, S.; Zheng, C.; Cui, B.; Qi, Z.; Zhao, Z.; An, Y.; Qiao, L.; Han, Y.; Zhou, Y.; Lu, J. Multiparametric imaging hippocampal neurodegeneration and functional connectivity with simultaneous PET/MRI in Alzheimer’s disease. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 2440–2452. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (A) Schematic representation of CADD process. (B) Comparison of traditional and computer-aided drug development in terms of time and cost investments.
Figure 1. (A) Schematic representation of CADD process. (B) Comparison of traditional and computer-aided drug development in terms of time and cost investments.
Ijms 22 04688 g001
Figure 2. Overview of molecular mechanisms and drug targets (red text) in Alzheimer’s, Parkinson’s, ALS, and Huntington’s diseases. Figure made using Biorender.com, accessed on 28 April 2021.
Figure 2. Overview of molecular mechanisms and drug targets (red text) in Alzheimer’s, Parkinson’s, ALS, and Huntington’s diseases. Figure made using Biorender.com, accessed on 28 April 2021.
Ijms 22 04688 g002
Table 1. The 30 most highly-scored pieces of software for CADD. Software was ranked according to the equation Si = log(106.Ci/Ctotal) where Si is the score for tool i, Ci is the number of citations to tool i, and Ctotal is the number of citations to all tools considered. Number of citations was obtained using Google Scholar, last accessed on 14 April 2021.
Table 1. The 30 most highly-scored pieces of software for CADD. Software was ranked according to the equation Si = log(106.Ci/Ctotal) where Si is the score for tool i, Ci is the number of citations to tool i, and Ctotal is the number of citations to all tools considered. Number of citations was obtained using Google Scholar, last accessed on 14 April 2021.
No.SoftwareNo. of Citations to Published StudiesScoreFeaturesAccessibility Website
1HADDOCK26,4904.7323Docks protein−protein based on biochemical or biophysical informationFreehttps://wenmr.science.uu.nl/haddock2.4/
2AutoDock Autodock 1 Autodock 2.4 Autodock 3 Autodock 4 Autodock 4.2 Autodock Vina AutoDockFR AutoDockTools22,4224.6599Automated docking toolsFree http://autodock.scripps.edu/
3Glide Glide 1.8 Glide 2 Glide 2.522,0914.6535Rapid, accurate docking and scoring approachSubscriptionhttps://www.schrodinger.com/glide
4FlexX19,9874.6100Predicts the geometry of the protein–ligand complex and estimates the binding affinityFreehttps://www.biosolveit.de/FlexX/
5LigandFit19,8904.6079Presents a shape-based approach for docking ligands into the active site of the protein Subscription https://www.phenix-online.org/documentation/reference/ligandfit.html
6AmberTools14,5724.4728A suite of biomolecular simulation programsSubscriptionhttps://ambermd.org/
7ENCoM13,1454.4280 A coarse-grained normal mode analysis method utilized for different residues in proteins or nucleotides in RNAFree http://biophys.umontreal.ca/nrg/resources.html
8PROCHECK-NMR10,7834.3420Checks the stereochemical quality of a protein structure solved by NMRFreehttps://www.ebi.ac.uk/thornton-srv/software/PROCHECK/
9MCDOCK10,6034.3347Allows for a full flexibility of ligands in the docking calculationsFreeDOI: 10.1021/jm990129n
10ICM ICM 2.8 ICM-Dock10,2714.3209A new method for protein modelling and design applications to docking and structure predictionSubscriptionhttp://www.molsoft.com/docking.html
11Dock Dock2 Dock3 Dock4 Dock5 Dock6 Dock7 Dock8 Dock981814.2221Based on a geometric matching algorithmFreehttp://dock.compbio.ucsf.edu/
12SOFT Docking74744.1828Predicts the sites of interaction between two cognate molecules based on their 3D structuresSubscriptionhttps://0-doi-org.brum.beds.ac.uk/10.1016/0022-2836(91)90859-5
13FDS71884.1659Cluster analysis based on distance similaritiesFree http://www.scfbio-iitd.res.in/dock/fds.jsp
14DockVision69504.1512Increases capability to generate laudable resultsFreehttp://dockvision.sness.net/overview/overview.html
15PRODOCK64424.1183Renders the programming easier and the definition of molecular flexibility more straightforwardSubscriptionhttps://0-doi-org.brum.beds.ac.uk/10.1002/(SICI)1096-987X(199903)20:4<412::AID-JCC3>3.0.CO;2-N
16YASARA YASARA Dynamics YASARA Model YASARA NMR Module YASARA Structure YASARA View YASARA Virtual Reality Workstation YASARA/WHAT IF Twinset 58704.0779A molecular-graphics, -modelling, and -simulation programFreehttp://www.yasara.org/products.htm
17KBDOCK58204.0742A program that proposes structural templates for protein dockingFree http://kbdock.loria.fr/
18TreeDock57964.0724A docking tool that is able to explore all clash-free orientations at very fine resolution in a reasonable timeSubscriptionhttps://0-doi-org.brum.beds.ac.uk/10.1021/ja011240x
19LePro56394.0605Generates a docking input file for LeDock with refined protein atoms within 0.4 nm of any atom of the ligandFreehttp://www.lephar.com/download.htm
20DockoMatic55944.0570A software that docks secondary ligands, used to assist inverse virtual screening Free https://0-doi-org.brum.beds.ac.uk/10.1186/1756-0500-3-289
21SYBYL_ChemScore SYBYL_D-Score SYBYL_F-Score SYBYL_G-Score54864.0485A conformational sampling and scoring functionSubscriptionhttps://0-doi-org.brum.beds.ac.uk/10.1021/jm0203783
22ZDOCK ZDOCKpro54154.0429A new scoring function for the initial stage of unbound dockingSubscriptionhttp://zdock.umassmed.edu/
23AADS50874.0157An automated active site identification, docking, and scoring protocolFree http://www.scfbio-iitd.res.in/dock/ActiveSite_new.jsp
24Surflex Dock48963.9991An automatic and flexible molecular docking algorithm for rapid in silico drug-screening applicationsSubscriptionhttps://0-doi-org.brum.beds.ac.uk/10.1007/s10822-007-9114-2
25PyMOL PyMOL 1.4.1 PyMOL 2.1.1 PyMOL 2.448053.9910An open-source, user-sponsored, molecular visualization systemSubscriptionhttp://www.pymol.org
26FlipDock46143.9733Allows the automated docking of flexible ligand molecules into active sites of flexible receptor moleculesFreehttp://flipdock.scripps.edu/
27SymmDock45453.9668A flexible induced-fit backbone refinement in molecular dockingFree http://bioinfo3d.cs.tau.ac.il/FiberDock/php.php
28ClusPro43603.9487A widely used tool for protein–protein dockingFree http://nrc.bu.edu/cluster
29Surflex41803.9304A robust screening toolSubscriptionhttps://pubmed.ncbi.nlm.nih.gov/12570372/
30ConsDock40013.9114A pose within 2 Ao RMSD of the X-ray structure can be performed with this softwareSubscriptionhttps://0-doi-org.brum.beds.ac.uk/10.1002/prot.10119
Table 2. NDs with specified molecular targets and selected examples of drugs that have been identified with the aid of in silico drug design. The assay format used to validate each drug is indicated and drugs that progressed to clinical trials are highlighted in bold.
Table 2. NDs with specified molecular targets and selected examples of drugs that have been identified with the aid of in silico drug design. The assay format used to validate each drug is indicated and drugs that progressed to clinical trials are highlighted in bold.
NDsMolecular Docking TargetsMoleculeSoftwareAssay Type
Alzheimer’s diseaseAcetylcholinesterase, Beta-secretase enzymes, Muscarinic and nicotinic ACh receptors, N-methyl-D-aspartate receptor, Tau proteins1-benzy-l1,2,3,4-tetrahydro- b-carboline), 3-substituted-1H-indoles, 6-triazolyl amidine derivatives [40]ICMcell-based assay [40]
Chloropyridonepezil [41]Autodock VinaIn vitro blood–brain barrier model [42]
Flavone, 5-hydroxyflavone, 7-hydroxyflavone, chrysin, apigenin, kaempferol, fisetin, and quercetin [43]AutoDockMice and rats models [44,45]
Ifenprodil [46]Schrödinger SuitePrimary cultures from chicken embryo forebrain (E10) [46]
Memantine [47,48]GlideHuman clinical trial [49]
Morin [50]GlideIn APPswe/PS1dE9 mice [51]
Pyridopyrimidine derivatives [52]Auto grid and auto dockIn vitro enzyme inhibitory model [53]
Pyridonepezil [54]Autodock VinaIn vitro blood–brain barrier model [42]
Piperazine derivatives [55]PASS softwareTested on AChE in vitro by using Ellman’s method [56]
Rutin [57]AutoDock and Autodock VinaDoxorubicin (DOX)-treated neuroblastoma cells (IMR32) and doxorubic-induced cognitive dysfunction in Wistar rats [58]
Parkinson’s diseaseDopamine receptors, expression and mitochondrial localization, Mutant LRRK2, Mutated, PINK1, PARK2, DJ1 SNCA MotifLRRK2 kinase inhibitors (9-methyl-N-phenylpurine-2,8-diamine, N-phenylquinazolin-4-amine, and 1,3-dihydroindol-2-one) [59]MOEBoth in vitro and in vivo studies were established [60]
Amyotrophic lateral sclerosisMutant SODI, SODI oligomerization, CASP-3, CASP-8, TDP-43, p38 MAPK Nav1.6 sodium channelAngiogenin [61]AmberTools20HeLa cells (Nuclear translocation assay) [61])
Hesperidin and THSG [62])(Molecular Dynamics (MD) SimulationHigh affinity to mutant SOD1 [62]
Riluzole [63]PROCHECK programFDA-approved drug for ALS [64]
Huntington’s diseaseFIP-2 Specificity protein, 1HTT Interacting proteins Mutant HTT, Infant Testing Nuclear receptor corepressor, Postsynaptic density-95T1–11 (synthesized in a high yield by the substitution reaction) [65]AutoDockToolsPC12 cells [65]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Salman, M.M.; Al-Obaidi, Z.; Kitchen, P.; Loreto, A.; Bill, R.M.; Wade-Martins, R. Advances in Applying Computer-Aided Drug Design for Neurodegenerative Diseases. Int. J. Mol. Sci. 2021, 22, 4688. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22094688

AMA Style

Salman MM, Al-Obaidi Z, Kitchen P, Loreto A, Bill RM, Wade-Martins R. Advances in Applying Computer-Aided Drug Design for Neurodegenerative Diseases. International Journal of Molecular Sciences. 2021; 22(9):4688. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22094688

Chicago/Turabian Style

Salman, Mootaz M., Zaid Al-Obaidi, Philip Kitchen, Andrea Loreto, Roslyn M. Bill, and Richard Wade-Martins. 2021. "Advances in Applying Computer-Aided Drug Design for Neurodegenerative Diseases" International Journal of Molecular Sciences 22, no. 9: 4688. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22094688

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

Article Metrics

Back to TopTop