Open access peer-reviewed chapter

Cholinesterases and Their Inhibitors

Written By

Mesut Işık

Reviewed: 10 January 2022 Published: 03 March 2022

DOI: 10.5772/intechopen.102585

From the Edited Volume

Hydrolases

Edited by Sajjad Haider, Adnan Haider and Angel Catalá

Chapter metrics overview

238 Chapter Downloads

View Full Metrics

Abstract

The main focus of this section is to review the available information on ChEs (ChEs) and their inhibitors. The ChE enzymes cause damage to the cholinergic system by hydrolyzing the neurotransmitter acetylcholine (ACh). ChE inhibitors, playing an important role in the cholinergic system, are used in the treatment of Alzheimer’s disease (AD) because of their effects on maintaining ACh levels in brain regions and preventing Aβ accumulation by inhibiting ChE. In this context, it is important to develop many synthetic and natural origin ChE inhibitors for the treatment of abnormalities in the cholinergic system and disorders with neuropsychiatric symptoms. In this section, firstly, general information about ACh and its synthesis in the cholinergic system is given, then ChEs and their catalytic properties, their roles in AD, and their molecular forms are explained. In the following section, the active site of Cantis was defined. The anti-ChE activity of the developed inhibitors was discussed, and then the mechanism of their binding to the ChE active site was explained by molecular docking. In the final section, many types of ChE inhibitors are described and discussed in detail in this section, and the properties and binding mechanism of these inhibitors are summarized.

Keywords

  • Alzheimer’s disease
  • acetylcholine
  • acetylcholinesterase
  • butyrylcholinesterase
  • cholinesterase inhibitors
  • molecular docking

1. Introduction

Acetylcholine (ACh), one of the most important neurotransmitters, is the primary substrate for cholinesterases (ChEs). First identified in autonomic ganglia, neuromuscular junctions, and many synapses in the central nervous system (CNS), the ACh is also the primary neurotransmitter in preganglionic sympathetic and parasympathetic neurons and in the adrenal medulla. In the CNS, ACh is mainly found in interneurons and long axon cholinergic pathways [1, 2]. The ACh is formed as a result of esterification of acetic acid with quaternary ammonium alcohol choline. This ACh found in cholinergic neurons and other cell types is biosynthesized by the transfer of acetyl group from acetyl-coenzyme A to choline catalyzed by choline acetyl transferase (ChAT, EC 2.3.1.6). The synthesis of ACh by ChAT is presented in the following reaction.

AcetylCoA+CholineChATAcetylcholine+CoASHE1

The synthesis of acetyl-CoA, which is the precursor for ACh synthesis from various metabolic pathways with carbohydrate, protein, and fat metabolism, has been reported by many researchers [3].

The cholinergic system is based on ACh, which was first recognized by Loewi in the 1920s and is widely found in both the peripheral and central nervous systems. There are two types of receptors in the nervous system and neuromuscular junctions: ACh receptors (muscarinic and nicotinic). The cholinergic receptors are also known to be expressed in many cells, including immune system and endothelial cells [4]. The receptors are identified based on their response to specific antagonists and agonists [5]. Although three types (M1–M3) of muscarinic receptors have been identified pharmacologically, five types have been reported based on molecular cloning experiments [6]. Muscarinic receptors are metabotropic and use G proteins for signal transduction, while nicotinic receptors are ionotropic and use ligand-gated ion channels for signal transduction [7]. The M1 receptor, one of the muscarinic receptor types and most common in the cerebral cortex, has its highest concentrations in the anterior olfactory nucleus, cerebral cortex, olfactory tubercle, dentate gyrus, hippocampus, and nucleus accumbens. The M2 receptor, a presynaptic autoreceptor that directs cholinergic release, is located in brain regions with abundant cholinergic neurons [8]. M3 and M4 receptors are located in the mainly in the olfactory tubercle, diencephalic, striatum, and brainstem regions [9]. Two types of the nicotinic receptor have also been identified in the CNS. The receptors are commonly found in the thalamus, substantia nigra, and periaqueductal gray [6, 10].

Advertisement

2. Cholinesterases

The existence of ChEs, one of the serine hydrolase class enzymes, was first suggested by Henry Dale in 1914, and research on the subject has been continuing since the 1930s. Research on the determination of the sequence of amino acids in the active site of ChEs has been started since the early 1959 [11]. The ChEs are divided into two groups due to their substrate selectivity to ACh and butyrylcholine. AcetylChE (AChE, acetylcholine acetyl hydrolase, EC 3.1.1.7) is known as true ChE, while butyrylChE (BChE, acylcholine acyl hydrolase, EC 3.1.1.8) is known as pseudoChE, serum ChE, or nonspecific ChE. While AChE inhibits at high substrate concentrations, BChE is activated. These enzymes also differ from each other in terms of selective inhibition of enzymes. The enzymes, which show 65% similarity in their amino acid sequence contents, are encoded in the chromosomes (7 and 3) [12, 13, 14].

The AChE is found in erythrocyte membranes and many tissues such as central and peripheral tissues, nerve and muscles, motor and sensory fibers, and noncholinergic and cholinergic fibers [1]. Another enzyme, BChE, is mainly synthesized in the liver and released into the plasma and is known to show high activity in many tissues [12, 13]. AChE, one of the fastest known enzymes with catalytic activity, is responsible for the hydrolysis of ACh in cholinergic synapses [11, 14]. Although the basic physiological function of BChE is not yet fully known, it is important both pharmacologically and toxicologically due to its ability to break down ester drugs such as carbamates, aspirin, succinylcholine, cocaine, antidepressant drugs such as sertraline, amitriptyline, pesticides, organophosphate, and chemical warfare agents. It has been reported that the expression of AChE and BChE is increased in many pathological conditions such as glioma, lung cancers, meningioma, leukemia, and ovarian tumors [12, 15].

The ChEs differ in their catalytic activities according to the artificial substrate preference. Since ACh is the more selective substrate for AChE, it is more hydrolyzed by AChE than by BChE. As expressed in Table 1, the catalytic activity of the ChEs may vary depending on the type of substrates [16]. AChE and BChE are responsible for the hydrolysis of different types of choline esters.

EnzymeSubstrateProduct
AChEAcetyl-β-methyl-thiocholineβ-metyl-thiocholine + acetate
AChEAcetyl-β-methyl-cholineβ-methyl-choline + acetate
AChE > BChEAcetylcholineacetate + choline
AChE > BChEAcetylthiocholineacetate + thiocholine
BChE > AChEButyrylcholinebutyrate + choline
BChE > AChEButyrylthiocholinebutyrate + thiocholine
BChESuccinylcholinesuccinate + choline
BChEBenzoylcholinebenzoiate + choline

Table 1.

Selected substrates and products of AChE and BChE hydrolysis [16].

AChE, playing an important role in impulse transmission by hydrolyzing ACh in the central and peripheral nervous system, catalyzes the hydrolysis of the neurotransmitter ACh to choline and acetate. BChE is responsible for the hydrolysis of its more specific substrate (butyrylcholine) to butyrate and choline. Hydrolysis of ACh/butyrylcholine by the ChEs is presented in the following reaction.

E2
E3

The ChE activity method is based on the formation of 5-thio-2-nitrobenzoic acid, a yellow compound, with 5,5-ditiyobis(2-nitrobenzoik) asit (DTNB) of thiocholine, which is formed by the hydrolysis of ACh and BCh as substrate. The ChE activities were determined using the Ellman method. Reaction mixture contains acetylthiocholine iodide (10 mM) for AChE and butyrylthiocholine iodide (0,5 mM) for BChE as substrate, 5,5-ditiyobis(2-nitrobenzoik) asit (DTNB, 10 mM), and Tris–HCl buffer (pH 8, 1 M) [17, 18, 19].

2.1 The role of cholinesterases in Alzheimer’s disease

Alzheimer’s disease (AD) is the most common type of dementia seen in the elderly population. Its etiology is not known exactly. This disease has been expressed as a neurodegenerative disease characterized by irreversible loss of nerve cells, difficulties in cholinergic nerve conduction, memory and mental functions, thinking and interpretation, and personality and behavioral disorders. It has been reported that amyloid β-peptide (Aβ) aggregation, neurofibrillary network formation originating from hyperphosphorylated tau proteins, oxidative stress, and low ACh levels are effective in the pathology of AD [20, 21, 22]. Some hypotheses have been proposed to explain the pathogenesis of AD.

Cholinergic hypothesis: It is one of the first hypotheses. In this hypothesis, it is thought that AD occurs due to the reduction of acetylcholine, an important neurotransmitter. Degeneration of cholinergic neurons in the basal forebrain of AD patients, as well as a significant decrease in cholinergic receptors and choline acetyltransferase (ChAT) levels in the cerebral cortex, was detected. Although many of the previous treatment approaches are based on this hypothesis, clinical studies have revealed that treatment strategies to increase ACh levels provide only symptomatic relief but are not effective in the development and treatment of the disease [23, 24]. However, there is evidence that the use of ChE inhibitors in new approaches based on this hypothesis may affect Aβ aggregation [25].

Amyloid cascade hypothesis: The hypothesis proposed in 1991 reports that Aβ deposits are the primary factor in AD. In AD, there is a series of neurodegenerative events triggered by the production, aggregation, storage, and toxicity of derivatives that occur after the production of amyloid precursor protein (APP) in the brain [26], and this study has been supported by molecular genetic, neuropathological, and biochemical studies. In this hypothesis, it is suggested that Aβ proteins, which accumulate in certain parts of the brain and form insoluble fibrils, later damage nerve cells by forming senile plaques; these cells break their connections with each other and reduce the amount of neurotransmitters [27].

Dysfunction of cholinergic neurotransmission in the brain causes an increase in AD symptoms. In the brain, choline-ester-based neurotransmitters are catalyzed by ChEs such as AChE and BChE. Therefore, ACh level decreases in the perisynaptic region of AD. Inhibition of AChE to prevent this reduction in ACh level is considered an important therapeutic target for the disease. The role of BChE in the progression of AD has been stated, and inhibition of BChE gains importance for treatment purposes. In addition, the presence of two types of the ChEs was detected between neuritic plaques and neurofibrillary tangles in AD brain. Thus, it has been suggested that both AChE and BChE may be involved in the formation of aggregates of the Aβ peptide in the AD brain [28]. Toxicity in AD brains varies depending on the amount of AChE and BChE, which have the potential to form complexes with Aβ aggregates [29, 30]. Depending on the disease, changes are seen in the amounts of AChE forms. The tetrameric form of AChE (G4), which predominates in the healthy brain, decreased at the onset of AD, while its monomeric and dimeric forms (G1 and G2) remained unchanged. In addition, studies have reported that the level of G1 and G2 forms is elevated in the plasma of AD patients [28, 31, 32].

2.2 Molecular forms of cholinesterases

The ChEs exist as amphiphilic or soluble molecular forms in tissues and body fluids [1, 12]. Monomeric, dimeric, and tetrameric molecular forms of ChE arise from the posttranslational modification of the expressed protein. The soluble monomeric and tetrameric membrane-bound forms are the predominant enzyme species in humans [10, 33] The forms are summarized below:

Amphiphilic dimers (Type 1): The dimeric form of AChE covalently bound to the membrane by a glycophosphatidylinositol extension, soluble only by detergents and aggregates in the absence of detergent. This form is found in erythrocytes, muscle, and lymphocytes of mammals.

Amphiphilic monomers and dimers (Type 2): Unlike Type 1, the form does not contain glycolipids anchor, does not aggregate in the absence of detergent, and can only be dissolved in salt solutions. The forms of the ChEs are frequently encountered in the muscle, brain, and intestine.

Hydrophobic-tailed tetramers (Type 3): This form of AChE, hydrophobically attached to the plasma membrane by a 20 kDa polypeptide anchor, is widely found in the CNS of mammals.

Collagen-like tailed or asymmetrical forms (Type 4): These forms can be identified by the presence of a collagen-like tail that allows them to attach to the basal lamina. This tail consists of collagen 3 helical subunits, each associated with a ChE tetramer. It finds it more common for AChE rather than BChE in neuromuscular junctions.

Soluble tetrameric form (G4): This form consists of four identical monomers and is stabilized by the interactions of hydrophobic amino acids at the C-terminus of the monomers. These forms are common in mammalian body fluids and soluble fractions of tissue homogenates. The form of ChE in the mammalian brain, mammalian body fluids, and soluble fractions of tissue homogenates is the tetrameric form (G4) [1].

2.3 Active site of cholinesterases

Homo sapiens AChE belonging to the serine hydrolases is known to have a very high catalytic activity, with each molecule of AChE degrading about 25,000 molecules of ACh per second [34, 35]. The enzyme’s active site contains the esteratic subsite and anionic subsite in catalytic center. There is a glutamate residue in the anionic region and a serine residue with a functional -OH group in the esteratic region (ES) and an imidazole ring. In addition to the ES, there is an acyl region and a choline-binding subsite at the catalytic active site (CAS) [34, 36]. The enzyme also contains the peripheral anionic site (PAS). The sites serve as sites for binding AChE and other quaternary ligands and involved in substrate-based inhibition [37]. The anionic site contains many amino acid residues such as Phe 330, Trp 84, and Tyr 121 for the electric eel AChE or Phe 338, Trp 86, and Tyr 337 for murine AChE. The PAS, localized on the AChE surface around the cavity entrance, was recognized as a target for multiple AChE activity modulators. Asp70 and Tyr332 residues in the PAS are involved in the binding of positively charged substrate/ligands and in the activity control of the enzyme. The site contains the most significant amino acids residues such as Asp 72, Tyr 121, Tyr 70, Trp 279, and Tyr 334 [16].

How the substrate binds to the active site of AChE (PDB code: 4EY5) via different types of amino acid residues is predicted by molecular docking. The quaternary nitrogen in the ACh interacts with the anionic site formed by the amino acid tryptophan (Trp), and through the aspartic acid and tyrosine located at the entrance of the active center, the substrates are directed to interact with the active center [38]. The molecular docking results show that ACh has Pi-cation interaction with Tyr337 and Trp86 in the active site of human AChE. In addition, nitrogen in the ACh formed a salt bridge with Glu202 (Figure 1). The AChE is composed of a hydrophobic cloud with Ala203, Hıs447, Gly120-Gly122, Gly448,Tyr133, Tyr124, and Phe297.

Figure 1.

2D and 3D interaction diagrams of ACh in the active site of human AChE (PDB ID: 4EY5).

Advertisement

3. Molecular docking

Molecular docking has gained importance with the increase in studies on drug development and design. Such studies have gained momentum with the creation of protein data banks. Molecular docking is a computer-generated tool for finding the optimal configuration and energy between the two if the protein and ligand structures are known. The basis of the method is to identify possible conformational states between the protein-ligand complex and to choose the complex with the lowest free energy. There are many applications such as Schrodinger, AutoDock, DOCK, and MolDock developed to simulate interactions (hydrogen bonding, hydrophobic interactions, Van der Waals) between two or more molecules [39]. Computational docking of test ligands to predict binding sites to the ChE active site becomes one approach to predict potential inhibitors against the target [40]. The docking studies were carried out using the GLIDE software [41].

Previous studies reported interactions with the AChE amino acid residues with various antipsychotic drugs. Recent studies reported interactions with critical residues such as Leu288, Ser292, Thr237, Val238, Gln368, Pi-cation interaction with Arg295, Pi-alkyl and alkyl interactions with Pro289, Val299, Pro367, 234 of AChE [42, 43] Donepezil and Pimozide ligands showed the lowest energy and the best favorable interactions. The Galanthamine (−7.3 kcal/mol) and Risvagtimine (−7.95 kcal/mol) showed higher binding energies than Donepezil (−8.5 kcal/mol) ligands [44].

In addition, in another study, binding free energy analysis was performed to understand that some drugs are potential candidates. It is a more suitable drug candidate with lower binding energies. Brexpiprazole and pimavanserin have binding free energies of −212.690 kcal/mol and − 108.626 kcal/mol, respectively, whereas Donepezil has binding energy of −180.517. The values, providing the highest contribution in terms of energy, are obtained from Van Der Waals forces. With significantly lower binding energy, these results indicate that Brexipyrazole with lower binding energy binds to the active site and other binding pockets in the protein with higher affinity than Donepezil and Pimavanserin [13]. In the study, it was stated that reference tacrine has higher binding affinity to AChE than to the natural substrate ACh [40]. In the present study, the tacrine was predicted to interact with AChE-binding site residues Trp84, Phe330, Tyr334, Ile 439, Trp432, Tyr442, and Gly441.The molecular docking results show that tacrine has Pi-stacking interaction with Phe 330 and Trp84 in the active site of human AChE. In addition, the AChE is composed of a hydrophobic cloud with Tyr334, Ile 439, Trp432, Tyr442, and Gly441 (Figure 2).

Figure 2.

2D and 3D interaction diagrams of tacrine in the active site of human AChE (PDB ID: 4EY5, Schrodinger 2017).

Advertisement

4. Cholinesterase inhibitors

ChE inhibitors constitute a wide group of compounds with different physicochemical properties, including drugs used in the treatment of many diseases, natural toxins, pesticides, and chemical warfare agents. The ChEs, which play an important role in the hydrolysis of pesticides, are also of great importance in food and agriculture research. Specific AChE inhibitors play an important role in the regulation of ACh metabolism due to the physiological importance of AChE, especially in the brain and blood. Since ACh is a more selective substrate for AChE, it takes a more active role in this metabolism than BChE. Specific BChE inhibitors such as tetraisopropyl pyrophosphoramide (iso-OMPA) are mainly of diagnostic importance [45, 46]. AChE inhibitors, one of the ChE inhibitors used in the treatment of many diseases, are more widely known than BChE. Drugs that suppress the development of AD through action on the cholinergic system are predominantly selective inhibitors of AChE. It is stated that these inhibitors also suppress Aβ aggregate formation and oxidative stress. Selective inhibitors of BChE for this disease have also been investigated as potential drugs for AD, but to a much lesser extent than AChE [47, 48, 49, 50]. Inhibition of AChE also plays an important role in nerve agent toxicology. However, BChE can sometimes replace AChE temporarily inhibited by these agents and slowly hydrolyze accumulated ACh [51].

Compounds with AChE inhibitory potential can be divided into three main groups as follows [52]:

  1. Compounds binding at the active site interact with either anionic site (e.g., tacrine) or esteratic (e.g., nerve agents).

  2. Compounds interacting with the aromatic gorge (e.g., decamethonium).

  3. Compounds bound at the PAS (e.g., propidium, huperzine).

In vitro inhibition studies: The inhibition effects of compounds against AChE are determined with at least five different inhibitor concentrations. The IC50 values of the compounds are calculated from activity (%)-[compound] graphs for each compound. The inhibition types and KI constants are found by Lineweaver and Burk’s curves according to previous works [53, 54, 55].

4.1 ChE inhibitors with therapeutic potential

Many compounds such as donepezil, galantamine, metrifonate, physostigmine, rivastigmine and tacrine, and memantine have been developed as ChE inhibitors. ChE inhibitors, having a regulatory effect on ACh metabolism, have been developed for the treatment of many diseases. For example, donepezil, galantamine, rivastigmine, tacrine, and memantine are drugs used in the treatment of AD [56]. However, the efficacy of these drugs is limited, and they are known to have various dose-related side effects, especially at higher doses. Galantamine and donepezil have AChE inhibitory potential [57, 58], whereas rivastigmine has a reversible inhibitory effect for both AChE and BChE. In particular, donepezil is a highly selective inhibitor for AChE rather than BChE. In a study, the inhibitory potentials (IC50 values) of donepezil, physostigmine, tacrine, and rivastigmine for AChE were found to be 6.7, 0.67, 77, and 4.3 nM, respectively [58].

Donepezil: Donepezil is a drug approved for the mild to moderate treatment of AD (Figure 3A). However, donepezil has many side effects such as muscle weakness loss of appetite, diarrhea, muscle cramps, nausea, and insomnia. In addition, when high doses of the drug are given to patients, many symptoms such as muscle weakness, low blood pressure, severe nausea, severe vomiting, and respiratory problems occur [59]. In addition to being a selective inhibitor of ChE, donepezil may have multiple mechanisms of action. They act at molecular and cellular levels in many pathways, including inhibition of various aspects of glutamate-induced excitotoxicity, stimulation of the neuroprotective isoform of AChE and reduction of oxidative-stress-dependent effects, and in the pathogenesis of AD [60]. The donepezil has a molecular structure that causes simultaneous inhibition by binding to the active and PAS of AChE. However, it has been stated that it does not interact directly with the catalytic triad or the oxyanion hole [61].

Figure 3.

The molecular structures of ChE inhibitors such as donepezil (a), galantamine (B), metrifonate (C), physostigmine (D), rivastigmine (E), and tacrine (F).

Galantamine: Galantamine, one of the alkaloid groups, is found in many plants and has been used as a medicine for decades in Russia and Eastern European countries for many purposes such as the treatment of myasthenia, myopathy, and CNS-related sensory and motor defects. The galantamine has the appropriate molecular structure to bind to nicotinic receptors in the cholinergic system (Figure 3B). As its effectiveness against ChE was revealed in the 1950s, it started to be used in the treatment of various neurological diseases [56]. It was approved for use after it was found to be effective in the treatment of cognitive and many symptoms related to AD. However, it also has many side effects such as severe nausea, convulsions, vomiting, stomach cramps, irregular breathing, and muscle weakness.

Metrifonate: Metrophonate, a prodrug of dichlorvos (2,2-dichlorovinyl dimethyl phosphate), is an irreversible AChE inhibitor with biphasic action (Figure 3C). Initially, it interacts competitively with the enzyme and then turns into a noncompetitive type of inhibition by phosphorylation of the enzyme esteratic site. In other words, metrifonate from the organophosphate group interacts with the esteratic site of the enzyme (Table 2). Having close inhibition effects on both AChE and BChE, metrifonate can be defined as a pseudo-irreversible ChE inhibitor [62].

InhibitorsClassEnzymeEnzymatic active site
DonepezilPiperidineBChE > AChEAnionic
GalantaminePhenanthrene
alkaloid
AChE > BChE
MetrifonateOrganophosphateBChE > AChEEsteratic
PhysostigmineCarbamateBChE > AChEEsteratic
RivastigmineCarbamateAChE > BChEEsteratic
TacrineAcridineBChE > AChEAnionic

Table 2.

Pharmacological properties of some selective ChE inhibitors [10].

Physostigmine: Physostigmine (also known as Eserine) with AChE inhibitory effect is one of the compounds first isolated from Calabar bean in 1864. Although this compound is able to cross the blood–brain barrier (BBB), its use is limited due to its short half-life and many side effects such as stomach cramps, diarrhea, increased saliva production, and excessive sweating. Therefore, physostigmine is not approved for the treatment of AD. Physostigmine from the organophosphate group has a molecular structure to interact with the ES of the enzyme (Figure 3D). Based on this structure, many of its derivatives have been designed and synthesized. Among the derivatives, eseroline, tolserine, and phenserine have ChE inhibitory potential [56, 63].

Rivastigmine: Rivastigmine has been used for the treatment of moderate AD and dementia associated with Parkinson’s disease [64]. Although it is stated that rivastigmine may exert its pharmacological effect through the cholinergic system, its mechanism of action has still not been clarified. Rivastigmine binds reversibly to both AChE and BChE and can cause inhibition. Also, rivastigmine is a carbamate class compound that is converted to various phenolic derivatives that are rapidly excreted from the body. The compound has the potential to bind with high affinity to the ES of AChE during ACh hydrolysis [56, 65], and has the appropriate molecular structure to bind to ChE in the cholinergic system (Figure 3E). The rivastigmine can cause many side effects such as weight loss, diarrhea, stomach pain, and loss of appetite, and in overdose, it can cause numerous symptoms such as irregular heartbeat and chest pain [66].

Tacrine: Tacrine, synthesized in the 1930s, was first used as a muscle relaxant antagonist and respiratory stimulant, moreover, due to its therapeutic effect on the cholinergic system in AD patients, and was approved by the FDA in 1993. The amine group of tacrine, whose molecular structure is presented in Figure 3F, interacts with amino acid residues Phe330 and Trp84 located in the “anionic region” of AChE (Table 2). As a result of the interaction, it has been stated that it is an effective inhibitor developed for the ChEs [67]. However, the use of tacrine has been limited due to liver toxicity, short half-life, and many side effects such as loss of appetite, nausea, vomiting, and diarrhea [56, 68].

4.2 Naturally derived inhibitors

ChE inhibitors used in the treatment of Alzheimer’s, Parkinson’s, and many diseases have short half-lives and many side effects. For this reason, there has been an increased interest in studies on the determination of natural origin inhibitors. Although these types of compounds have less ChE inhibitory activity than synthetics, they have much less side effects due to their natural origin.

Phenolic compounds: A number of flavonoid compounds, which have free radical scavenging properties and important roles in the prevention of oxidative stress, are natural ChE inhibitors in vitro. Galangin, a flavonol group from Alpiniae officinarum, showed a strong inhibitory effect against AChE [69]. Many studies have reported that phenolic compounds, reducing oxidative stress due to antioxidant properties, have inhibitory effects on AChE. The inhibitory effect on the AChE of phenolic compounds strongly depends on the structure of a particular compound, especially the position and/or number of the C=O and OH groups [70]. Chlorogenic acid has an inhibition effect on AChE in the hippocampus and frontal cortex (IC50: 98.17 μg/ml). In vitro, caffeic acid has an activation effect in the cerebral cortex, cerebellum, hypothalamus, whole blood, and lymphocytes, while it has an inhibition effect at the concentrations of 0.5, 1.0, 1.5, and 2 mM in the muscles [71]. Hydroquinone, chlorogenic acid, and 4-hydroxybenzoic acid have inhibitory potential against the AChE with IC50 and KI values in the range of 0.26 ± 0.01–36.34 ± 2.72 mM and 0.72 ± 0.00–29.23 ± 2.62 mM, respectively. The effectiveness of the compounds has been associated with its structure [49]. Consequently, as phenolic compounds have both AChE inhibitory effect and antioxidant properties, they can be considered as alternative drugs in the treatment of AD.

Cardanol: In 2009, various non-isoprenoid phenolic lipids obtained from Anacardium occidentale were investigated for their inhibitory activity against AChE. In particular, cardanol, a phenolic lipid, has shown promising results [56]. In one study, a novel series of cardanol derivatives was designed as AChE inhibitors and tested for their inhibitory effects against ChEs. The derivatives showed the highest inhibitory activity against AChE, with IC50 of 6.6 μM [72].

Huperzine (Hup): Hup, obtained from the Huperzia serrata, is available in two types, Hup-A and Hup-B (Figure 4A and B, respectively). Hup-A is used in the treatment of AD and age-related memory loss and in improving cognitive functions by regulating ACh level. It is more effective than tacrine, galantamine, and rivastigmine and is a highly selective and potent inhibitor of AChE. However, it is less active against BChE compared with AChE. Many hybrids of Hup-A also have an AChE inhibitory effect [56, 73].

Figure 4.

The molecular structures of Hup-A and Hup-B.

4.3 Synthetic analogs

In order to prevent or reduce the side effects and toxicity of ChE inhibitors, which are known to be used as drugs, their synthetic analogs have been developed as ChE inhibitors [74]. However, the main problem of synthetic analogs is their inability to penetrate the blood–brain barrier (BBB), and their inhibitory effect may be lower compared with reference inhibitors [75]. The derivatives of tacrine have shown improved AChE-inhibitory activities compared with the tacrine used as drug [76]. Various analog compounds were synthesized and tested by Ali et al. in 2009. Most of them showed moderate AChE-inhibitory effects. Ali et al. suggested that the presence of methoxy groups on the phenyl ring significantly improved the inhibition of AChE [75]. Sulfonamide compounds, which have a wide range of biological applications such as antimicrobial, antiviral, diuretic, and anticancer agents, are found as active ingredients in many drugs. In recent years, many studies have been carried out on the design and synthesis of sulfonamide-derived compounds with ChE inhibitory potential. Many of the compounds have been reported to be selective inhibitors for AChE and BChE [19, 77, 78, 79, 80].

Advertisement

5. Conclusions

In conclusion, many synthetic and natural ChE inhibitors have been discovered recently. Due to the side effects of ChE inhibitors used in the treatment of many neurodegenerative diseases such as AD, studies conducted within the scope of the discovery and design of alternative inhibitors have gained importance. Although the side effects of natural origin inhibitors are much less than those of synthetic compounds, their effect is low. In some studies, derivatives of known inhibitors have been synthesized to increase the effectiveness of inhibitors and reduce their toxicity. The binding affinities of these derivatives to the enzyme active site vary depending on the structure. Therefore, the design of targeted inhibitors suitable for the enzyme active site by molecular docking method sheds light on the development of alternative drugs for treatment. The design of compounds with the ChE inhibitor potential by molecular docking method provides a significant advantage in terms of financial and workload since ineffective syntheses are not made.

Advertisement

Acknowledgments

I would like to thank Dr. Cüneyt Türkeş for his kindly helpful suggestions during the preparation of the book chapter.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Appendices and nomenclature

ADAlzheimer disease
AChacetylcholine
AChEacetylcholinesterase
beta‐amyloid
BBBblood-brain-barrier
BChEbutyrylcholinesterase
CAScatalytic active site
CNScentral nervous system
ChEcholinesterases
ChATcholine acetyl transferase
DTNB5,5-ditiyobis(2-nitrobenzoik) asit
ESesteratic region
iso-OMPAtetraisopropyl pyrophosphoramide
PASperipheral anionic site

References

  1. 1. Wang R, Xi CT. Neuroprotective effects of huperzine a: A natural cholinesterase inhibitor for the treatment of Alzheimer’s disease. Neuro-Signals. 2005;14:71-82
  2. 2. Perry E, Walker M, Grace J, Perry R. Acetylcholine in mind: A neurotransmitter correlate of consciousness? Trends in Neurosciences. 1999;22:273-280
  3. 3. Jankowska-Kulawy A, Bielarczyk H, Pawelczyk T, Wróblewska M, Szutowicz A. Acetyl-CoA and acetylcholine metabolism in nerve terminal compartment of thiamine deficient rat brain. Journal of Neurochemistry. 2010;115:333-342
  4. 4. Wessler I, Kirkpatrick CJ. Acetylcholine beyond neurons: The non-neuronal cholinergic system in humans. British Journal of Pharmacology. 2008;154:1558-1571
  5. 5. BL S, SI D, NR S, SM R, CH B, RB R. The effects of prolonged administration of an alpha7 nicotinic cholinergic agonist intervention on neurocognitive function in schizophrenia. Neuropsychopharmacology. 2011;36:2774-2781
  6. 6. Hurst R, Rollema H, Bertrand D. Nicotinic acetylcholine receptors: From basic science to therapeutics. Pharmacology and Therapeutics. 2013;137:22-54
  7. 7. Turner AJ. Proteins, transmitters and synapses. Biochemical Education. 1995:223-223
  8. 8. Flynn DD, Ferrari-DiLeo G, Mash DC, Levey AI. Differential regulation of molecular subtypes of muscarinic receptors in Alzheimer’s disease. Journal of Neurochemistry. 1995;64:1888-1891
  9. 9. Schliebs R, Roßner S. Distribution of muscarinic acetylcholine receptors in the CNS. CNS Neurotransmitters and Neuromodulators. 1995. pp. 67-83
  10. 10. Cummings JL. Cholinesterase inhibitors: A new class of psychotropic compounds. The American Journal of Psychiatry. 2000;157:4-15
  11. 11. Massoulié J, Pezzementi L, Bon S, Krejci E, Vallette FM. Molecular and cellular biology of cholinesterases. Progress in Neurobiology. 1993;41:31-91
  12. 12. Cokugras AN. Butyrylcholinesterase: Structure and physiological importance. Turkish Journal of Biochemistry. 2003;28:54-61
  13. 13. Kundu D, Dubey VK. Potential alternatives to current cholinesterase inhibitors: An in silico drug repurposing approach. Drug Development and Industrial Pharmacy. 2021;47:919-930
  14. 14. Işık M, Beydemir Ş, Yılmaz A, Naldan ME, Aslan HE, Gülçin İ. Oxidative stress and mRNA expression of acetylcholinesterase in the leukocytes of ischemic patients. Biomedicine & Pharmacotherapy. 2017;87:561-567
  15. 15. Özturan Özer E, Unsal Tan O, Ozadali K, Küçükkilinç T, Balkan A, Uçar G. Synthesis, molecular modeling and evaluation of novel N′-2-(4- benzylpiperidin−/piperazin-1-yl)acylhydrazone derivatives as dual inhibitors for cholinesterases and Aβ aggregation. Bioorganic Med Chem Lett. 2013;23:440-443
  16. 16. Pohanka M. Cholinesterases, a target of pharmacology and toxicology. Biomed Pap. 2011:155
  17. 17. Ellman GL, Courtney KD, Andres V, Featherstone RM. A new and rapid colorimetric determination of acetylcholinesterase activity. Biochemical Pharmacology. 1961;7:88-95
  18. 18. Türkeş C, Arslan M, Demir Y, Çoçaj L, Rifati Nixha A, Beydemir Ş. Synthesis, biological evaluation and in silico studies of novel N-substituted phthalazine sulfonamide compounds as potent carbonic anhydrase and acetylcholinesterase inhibitors. Bioorganic Chemistry. 2019;89:103004
  19. 19. Türkeş C, Akocak S, Işık M, Lolak N, Taslimi P, Durgun M, et al. Novel inhibitors with sulfamethazine backbone: Synthesis and biological study of multi-target cholinesterases and α-glucosidase inhibitors. Journal of Biomolecular Structure & Dynamics. 2021:1-13
  20. 20. Banks WA. Drug delivery to the brain in Alzheimer’s disease: Consideration of the blood-brain barrier. Advanced Drug Delivery Reviews. 2012;64:629-639
  21. 21. Maltsev AV, Bystryak S, Galzitskaya OV. The role of β-amyloid peptide in neurodegenerative diseases. Ageing Research Reviews. 2011;10:440-452
  22. 22. Işık M, Beydemir Ş. AChE mRNA expression as a possible novel biomarker for the diagnosis of coronary artery disease and Alzheimer’s disease, and its association with oxidative stress. Archives of Physiology and Biochemistry. 2019:1-8
  23. 23. Auld DS, Kornecook TJ, Bastianetto S, Quirion R. Alzheimer’s disease and the basal forebrain cholinergic system: Relations to β-amyloid peptides, cognition, and treatment strategies. Progress in Neurobiology. 2002;68:209-245
  24. 24. Bartus RT, Emerich DF, Davis KL, Marin D, Purohit DP, Perl D, et al. Cholinergic markers in Alzheimer disease. Journal of the American Medical Association. 1999;282:2208-2209
  25. 25. Shen ZX. Brain cholinesterases: III. Future perspectives of AD research and clinical practice. Medical Hypotheses. 2004;63:298-307
  26. 26. Hardy JA, Higgins GA. Alzheimer’s disease: The amyloid cascade hypothesis. Science. 1992;256:184-185
  27. 27. Robinson SR, Bishop GM. Aβ as a bioflocculant: Implications for the amyloid hypothesis of Alzheimer’s disease. Neurobiology of Aging. 2002;23:1051-1072
  28. 28. Kandasamy S, Loganathan C, Sakayanathan P, Karthikeyan S, Stephen AD, Marimuthu DK, et al. In silico, theoretical biointerface analysis and in vitro kinetic analysis of amine compounds interaction with acetylcholinesterase and butyrylcholinesterase. International Journal of Biological Macromolecules. 2021;185:750-760
  29. 29. Alvarez A, Alarcón R, Opazo C, Campos EO, Muñoz FJ, Calderón FH, et al. Stable complexes involving acetylcholinesterase and amyloid-β peptide change the biochemical properties of the enzyme and increase the neurotoxicity of Alzheimer’s fibrils. The Journal of Neuroscience. 1998;18:3213-3223
  30. 30. Mesulam -Marsel, Geula C. Butyrylcholinesterase reactivity differentiates the amyloid plaques of aging from those of dementia. Annals of Neurology. 1994;36:722-727
  31. 31. Atack JR, Perry EK, Bonham JR, Perry RH, Tomlinson BE, Blessed G, et al. Molecular forms of acetylcholinesterase in senile dementia of Alzheimer type: Selective loss of the intermediate (10S) form. Neuroscience Letters. 1983;40:199-204
  32. 32. García-Ayllón MS, Riba-Llena I, Serra-Basante C, Alom J, Boopathy R, Sáez-Valero J. Altered levels of acetylcholinesterase in Alzheimer plasma. PLoS One. 2010;5:e8701
  33. 33. Schegg KM, Harrington LS, Neilsen S, Zweig RM, Peacock JH. Soluble and membrane-bound forms of brain acetylcholinesterase in Alzheimer’s disease. Neurobiology of Aging. 1992;13:697-704
  34. 34. Colovic MB, Krstic DZ, Lazarevic-Pasti TD, Bondzic AM, Vasic VM. Acetylcholinesterase inhibitors: Pharmacology and toxicology. Current Neuropharmacology. 2013;11:315-335
  35. 35. Taylor P, Radić Z. The cholinesterases: From genes to proteins. Annual Review of Pharmacology and Toxicology. 1994;34:281-320
  36. 36. Tougu V. Acetylcholinesterase: mechanism of catalysis and inhibition. Current Medicinal Chemistry-Central Nervous System Agents. 2001;1:155-170
  37. 37. Lopes JPB, Silva L, da Costa FG, Antonio Ceschi M, Seibert Lüdtke D, Ferreira Dantas R, et al. Design, synthesis, cholinesterase inhibition and molecular modelling study of novel tacrine hybrids with carbohydrate derivatives. Bioorganic. Medicinal Chemistry. 2018;26:5566-5577
  38. 38. Masson P, Legrand P, Bartels CF, Froment MT, Schopfer LM, Lockridge O. Role of aspartate 70 and tryptophan 82 in binding of succinyldithiocholine to human butyrylcholinesterase. Biochemistry. 1997;36:2266-2277
  39. 39. Novikov FN, Chilov GG. Molecular docking: Theoretical background, practical applications and perspectives. Mendeleev Communications. 2009;19:237-242
  40. 40. Mohan V, Gibbs A, Cummings M, Jaeger E, DesJarlais R. Docking: Successes and challenges. Current Pharmaceutical Design. 2005;11:323-333
  41. 41. Release S. 2: Maestro. New York, NY: Schrödinger, LLC; 2017
  42. 42. Larik FA, Shah MS, Saeed A, Shah HS, Channar PA, Bolte M, et al. New cholinesterase inhibitors for Alzheimer’s disease: Structure activity relationship, kinetics and molecular docking studies of 1–butanoyl–3–arylthiourea derivatives. International Journal of Biological Macromolecules. 2018;116:144-150
  43. 43. Ujan R, Saeed A, Channar PA, Larik FA, Abbas Q, Alajmi MF, et al. Drug-1,3,4-thiadiazole conjugates as novel mixed-type inhibitors of acetylcholinesterase: Synthesis, molecular docking, pharmacokinetics, and ADMET evaluation. Molecules. 2019;24:860
  44. 44. Kumar S, Chowdhury S, Kumar S. In silico repurposing of antipsychotic drugs for Alzheimer’s disease. BMC Neuroscience. 2017;18:1-16
  45. 45. Adler M, Filbert MG. Role of butyrylcholinesterase in canine tracheal smooth muscle function. FEBS Letters. 1990;267:107-110
  46. 46. Akdag A, Işık M, Göktaş H. Conducting polymer-based electrochemical biosensor for the detection of acetylthiocholine and pesticide via acetylcholinesterase. Biotechnology and Applied Biochemistry. 2021;68:1113-1119
  47. 47. Villarroya M, García A, Marco J. New classes of AChE inhibitors with additional pharmacological effects of interest for the treatment of Alzheimers disease. Current Pharmaceutical Design. 2004;10:3177-3184
  48. 48. Giacobini E. Selective inhibitors of butyrylcholinesterase: A valid alternative for therapy of Alzheimer’s disease? Drugs and Aging. 2001;18:891-898
  49. 49. Işık M, Beydemir Ş. The impact of some phenolic compounds on serum acetylcholinesterase: Kinetic analysis of an enzyme/inhibitor interaction and molecular docking study. Journal of Biomolecular Structure & Dynamics. 2021;39:6515-6523
  50. 50. Işık M. The binding mechanisms and inhibitory effect of intravenous Anesthetics on AChE In vitro and In vivo: Kinetic analysis and molecular docking. Neurochemical Research. 2019;44:2147-2155
  51. 51. Giacobini E. Cholinesterase inhibitors: New roles and therapeutic alternatives. Pharmacological Research. 2004;50:433-440
  52. 52. Weiner L, Shnyrov VL, Konstantinovskii L, Roth E, Ashani Y, Silman I. Stabilization of torpedo californica acetylcholinesterase by reversible inhibitors. Biochemistry. 2009;48:563-574
  53. 53. Türkeş C, Beydemir Ş. Inhibition of human serum Paraoxonase-I with Antimycotic drugs: In vitro and In silico studies. Applied Biochemistry and Biotechnology. 2020;190:252-269
  54. 54. Demir Y, Işık M, Gülçin İ, Beydemir Ş. Phenolic compounds inhibit the aldose reductase enzyme from the sheep kidney. Journal of Biochemical and Molecular Toxicology. 2017;31:e21936
  55. 55. Lineweaver H, Burk D. The determination of enzyme dissociation constants. Journal of the American Chemical Society. 1934;56:658-666
  56. 56. Sharma K. Cholinesterase inhibitors as Alzheimer’s therapeutics (review). Molecular Medicine Reports. 2019;20:1479-1487
  57. 57. Weinstock M. Selectivity of cholinesterase inhibition. CNS Drugs. 1999;12:307-323
  58. 58. Ogura H, Kosasa T, Kuriya Y, Yamanishi Y. Comparison of inhibitory activities of donepezil and other cholinesterase inhibitors on acetylcholinesterase and butyrylcholinesterase in vitro. Methods and Findings in Experimental and Clinical Pharmacology. 2000;22:609-613
  59. 59. Rogers SL, Farlow MR, Doody RS, Mohs R, Friedhoff LT. A 24-week, double-blind, placebo-controlled trial of donepezil in patients with Alzheimer’s disease. Neurology. 1998;50:136-145
  60. 60. Jacobson SA, Sabbagh MN. Donepezil: Potential neuroprotective and disease-modifying effects. Expert Opinion on Drug Metabolism and Toxicology. 2008;4:1363-1369
  61. 61. Kryger G, Silman I, Sussman JL. Structure of acetylcholinesterase complexed with E2020 (Ariceptρ): Implications for the design of new anti-Alzheimer drugs. Structure. 1999;7:297-307
  62. 62. Nordberg A, Svensson AL. Cholinesterase inhibitors in the treatment of Alzheimer’s disease. A comparison of tolerability and pharmacology. Drug Safety. 1998;19:465-480
  63. 63. Coelho F, Birks JFC. Physostigmine for Alzheimer’s disease. Cochrane Database of Systematic Reviews. 2001:Cd001499
  64. 64. Inglis F. The tolerability and safety of cholinesterase inhibitors in the treatment of dementia. International Journal of Clinical Practice, Supplement. 2002;127:45-46
  65. 65. Onor ML, Trevisiol M, Aguglia E. Rivastigmine in the treatment of Alzheimer’s disease: An update. Clinical interventions in aging. 2007;2:17-32
  66. 66. Corey-Bloom J, Anand R, Veach J. A randomized trial evaluating the efficacy and safety of ENA 713 (rivastigmine tartrate), a new acetylcholinesterase inhibitor, in patients with mild to moderately severe Alzheimer’s disease. International Journal of Geriatric Psychopharmacology. 1998;1:55-65
  67. 67. Harel M, Schalk I, Ehret-Sabatier L, Bouet F, Goeldner M, Hirth C, et al. Quaternary ligand binding to aromatic residues in the active-site gorge of acetylcholinesterase. Proceedings of the National Academy of Sciences of the United States of America. 1993;90:9031-9035
  68. 68. Farlow M, Gracon SI, Hershey LA, Lewis KW, Sadowsky CH, Dolan UJ. A controlled trial of Tacrine in Alzheimer’s disease. JAMA J Am Med Assoc. 1992;268:2523-2529
  69. 69. Guo AJY, Xie HQ, Choi RCY, Zheng KYZ, Bi CWC, Xu SL, et al. Galangin, a flavonol derived from Rhizoma Alpiniae Officinarum, inhibits acetylcholinesterase activity in vitro. Chemico-Biological Interactions. 2010;187:246-248
  70. 70. Szwajgier D. Anticholinesterase activity of phenolic acids and their derivatives. Zeitschrift fur Naturforsch - sect C Journal of Biosciences. 2013;68:125-132
  71. 71. Anwar J, Spanevello RM, Thomé G, Stefanello N, Schmatz R, Gutierres J, et al. Effects of caffeic acid on behavioral parameters and on the activity of acetylcholinesterase in different tissues from adult rats. Pharmacology, Biochemistry, and Behavior. 2012;103:386-394
  72. 72. Lemes LFN, De Andrade RG, De Oliveira AS, Da Silva FMR, De Castro CG, Da Silva BM, et al. Cardanol-derived AChE inhibitors: Towards the development of dual binding derivatives for Alzheimer’s disease. European Journal of Medicinal Chemistry. 2016;108:687-700
  73. 73. Camps P, El Achab R, Morral J, Munoz-Torrero D, Badia A, Eladi Banos J, et al. New tacrine-huperzine a hybrids (huprines): Highly potent tight-binding acetylcholinesterase inhibitors of interest for the treatment of Alzheimer’s disease. Journal of Medicinal Chemistry. 2000;43:4657-4666
  74. 74. Korabecny J, Musilek K, Holas O, Binder J, Zemek F, Marek J, et al. Synthesis and in vitro evaluation of N-alkyl-7-methoxytacrine hydrochlorides as potential cholinesterase inhibitors in Alzheimer disease. Bioorganic Med Chem Lett. 2010;20:6093-6105
  75. 75. Ali MA, Yar MS, Hasan MZ, Ahsan MJ, Pandian S. Design, synthesis and evaluation of novel 5,6-dimethoxy-1-oxo-2,3-dihydro-1H-2-indenyl-3,4-substituted phenyl methanone analogues. Bioorganic Med Chem Lett. 2009;19:5075-5077
  76. 76. De La Torre P, Saavedra LA, Caballero J, Quiroga J, Alzate-Morales JH, Cabrera MG, et al. A novel class of selective acetylcholinesterase inhibitors: Synthesis and evaluation of (E)-2-(benzo[d]thiazol-2-yl)-3-heteroarylacrylonitriles. Molecules. 2012;17:12072-12085
  77. 77. Lolak N, Akocak S, Türkeş C, Taslimi P, Işık M, Beydemir Ş, et al. Synthesis, characterization, inhibition effects, and molecular docking studies as acetylcholinesterase, α-glycosidase, and carbonic anhydrase inhibitors of novel benzenesulfonamides incorporating 1,3,5-triazine structural motifs. Bioorganic Chemistry. 2020;100:103897
  78. 78. Taslimi P, Işık M, Türkan F, Durgun M, Türkeş C, Gülçin İ, et al. Benzenesulfonamide derivatives as potent acetylcholinesterase, α-glycosidase, and glutathione S-transferase inhibitors: Biological evaluation and molecular docking studies. Journal of Biomolecular Structure & Dynamics. 2021;39:5449-5460
  79. 79. Işık M, Demir Y, Durgun M, Türkeş C, Necip A, Beydemir Ş. Molecular docking and investigation of 4-(benzylideneamino)- and 4-(benzylamino)-benzenesulfonamide derivatives as potent AChE inhibitors. Chemical Papers. 2020;74:1395-1405
  80. 80. Akocak S, Taslimi P, Lolak N, Işık M, Durgun M, Budak Y, et al. Synthesis, characterization, and inhibition study of novel substituted Phenylureido Sulfaguanidine derivatives as α-glycosidase and cholinesterase inhibitors. Chemistry & Biodiversity. 2021;18:e2000958

Written By

Mesut Işık

Reviewed: 10 January 2022 Published: 03 March 2022