Next Article in Journal
Phytochemicals from Rhizophora mucronata Propagules, Its In Vitro Anti-Cancer and In Silico Drug-Likeness Potential
Next Article in Special Issue
Piperine Derivatives Enhance Fusion and Axonal Transport of Mitochondria by Activating Mitofusins
Previous Article in Journal
Synthesis and Self-Assembling Properties of Peracetylated β-1-Triazolyl Alkyl D-Glucosides and D-Galactosides
Previous Article in Special Issue
Carbon Monoxide Therapy Using Hybrid Carbon Monoxide-Releasing/Nrf2-Inducing Molecules through a Neuroprotective Lens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Questions in the Chemical Enzymology of MAO

1
Biomedical Sciences Research Complex, School of Biology, University of St Andrews, St Andrews KY16 9ST, UK
2
Laboratory for Food Chemistry, Department of Analytical Chemistry, National Institute of Chemistry, Hajdrihova 19, SI-1000 Ljubljana, Slovenia
*
Author to whom correspondence should be addressed.
Chemistry 2021, 3(3), 959-978; https://doi.org/10.3390/chemistry3030069
Submission received: 10 July 2021 / Revised: 25 August 2021 / Accepted: 28 August 2021 / Published: 31 August 2021

Abstract

:
We have structure, a wealth of kinetic data, thousands of chemical ligands and clinical information for the effects of a range of drugs on monoamine oxidase activity in vivo. We have comparative information from various species and mutations on kinetics and effects of inhibition. Nevertheless, there are what seem like simple questions still to be answered. This article presents a brief summary of existing experimental evidence the background and poses questions that remain intriguing for chemists and biochemists researching the chemical enzymology of and drug design for monoamine oxidases (FAD-containing EC 4.1.3.4).

1. Introduction

Monoamine oxidase (E.C. 1.4.3.4) enzymes MAO A and MAO B are FAD-containing proteins located on the outer face of the mitochondrial inner membrane, retained there by hydrophobic interactions and a transmembrane helix. The redox co-factor (FAD) is covalently attached to a cysteine and buried deep inside the protein [1]. Discovered as tyramine oxidase [2], an enzyme that metabolized pharmacologically active amines such as dopamine [3], MAO binds a range of small molecules, whose kinetic parameters are shown in Table 1.
MAO became an important drug target when a tuberculosis drug was found to improve the mood of patients. The large antidepressant market fueled the discovery of other irreversible inhibitors of MAO [4,5,6] that act by covalent attachment [7,8,9]. The common drugs such as selegiline (Table 1, used in the treatment of Parkinson’s disease) modify the flavin, but binding to an amino acid residue was also reported for some compounds [10,11]. The major side-effect of treatment with the irreversible inhibitors is the tyramine-induced pressor “cheese effect” due to the inhibition of intestinal MAO A that normally metabolizes dietary tyramine found at high levels in aged cheeses [12]. The treatment of depression then switched to re-uptake inhibitors and drug discovery for MAO inhibition switched to reversible inhibitors, although the irreversible inactivators of MAO still remain important for treatment of certain types of depression. Furthermore, the increasing prevalence of neurodegenerative diseases has renewed interest in MAO inhibitors (MAOI) that can conserve the decreased levels of monoamine neurotransmitters in a deteriorating brain.
Although it is almost 100 years after the identification of monoamine oxidase, 70 years after its adoption as a drug target and 20 years after its structure was solved, some interesting questions still remain to be explored. These are outlined in this article as a starting point for new investigations.

2. What Is the Chemical Mechanism of Amine Oxidation? Is It the Same for Both MAO A and MAO B?

It is clear that MAO A and MAO B have different substrate and inhibitor selectivities (for example, [13,19,20,21]) defined by the distinct hydrophobicity and binding site architectures, but the key region close to the FAD is the same [21]. The chemical mechanism was substantially explored by traditional chemical approaches in the 1990s, but only four mechanisms for oxidation of the amine by MAO have attracted major attention: single electron transfer (SET), the polar nucleophilic mechanism or (as found in some other FAD-dependent oxidases) hydride extraction, and a derivative thereof, two-step hydride transfer [22] (see Figure 1). The strong chemical evidence for the single electron transfer mechanism came from in depth studies on the inactivation of MAO by cyclopropylamines and remains a probable mechanism for conversion of a cyclopropylamine into an activated product capable of forming a covalent bond either to the FAD or to a cysteine [10,23]. However, inability to detect a flavin semiquinone radical by electron paramagnetic resonance (EPR) during substrate turnover cast doubt on SET as the general catalytic pathway. Quantum chemical calculations for a SET mechanism for para-substituted benzylamines gave inverse correlation with experimental data, confirming that this is not the mechanism of substrate oxidation [24].
In depth studies based on Hammett correlations were carried out in support of a polar nucleophilic mechanism for MAO A where a convincing positive correlation of oxidation rate with electronic effects of benzylamine substituents was obtained (ρ = 2) [25]. In this study with MAO A, the more electron withdrawing character of the substituent, the faster was the rate of oxidation, consistent with the polar nucleophilic mechanism whereby a transient C4a adduct was formed with the flavin. In contrast, for MAO B, the substituents that more efficiently stabilize the developing electron density in the transition state slowed the rate, prompting the suggestion that MAO A and MAO B followed different mechanisms [26]. However, although the slower rate of MAO B with electron withdrawing substituents could be consistent with the hydride mechanism, nitrogen kinetic isotope effects for the oxidation of benzylamine by MAO B showed that the rehybridization of the N(5) nitrogen atom required for reduction of the flavin and the cleavage of the CH bond for oxidation of the amine were not concerted [27]. A quantum mechanics/molecular mechanics (QM/MM) study of benzylamine oxidation by the MAO also suggested that the transfer of electrons and proton was asynchronous, supporting the polar nucleophilic mechanism [28].
Hydride transfer was established for D-amino acid oxidase after high-resolution structures with bound substrates were solved giving insight into the probable transition state [29]. On the other hand, X-ray structures of MAO are available only with inhibitors bound (see, for example, [30]), but since computational modeling based on these structures became practical, evidence for a hydride transfer reaction in both MAO A and MAO B has been mounting [31,32,33,34]. In a QM/MM simulation of the rate-limiting step for the serotonin oxidation via hydride transfer to the N5 of the FAD in MAO A, the calculated barrier was 14.8 ± 0.8 kcal mol−1, in good agreement with the experimental value of 16.0 kcal mol−1. The two-step hydride transfer mechanism shown for MAO B proposes a hydride abstraction from the substrate and relaxation of the resulting flavin-substrate adduct into a neutral amine and fully reduced flavin. The second part of the reaction is presumably facilitated by two water molecules present at the binding site [34]. The activation free energy for the hydride transfer in this computational study was shown to be roughly two-times smaller (24 kcal mol−1) than for any other oxidation mechanism by MAO B.
Whether further experimental or modeling work would give clearer evidence for the same hydride mechanism in both MAO A and MAO B is not certain. However, could two enzymes with such similar catalytic sites support non-identical chemical mechanisms? Using Hammett correlations as in the example above has proved invaluable in many chemical mechanistic studies. However, inductive and resonance effects of the substituents on the aryl rings of reactants/substrates are rarely the only (although might be the prevailing) factors governing the reaction kinetics. It has been demonstrated for MAO, for instance, that steric hindrance in 2′-aryl substituted analogs of MPTP had a significant effect on the determined Hammett values [35], and not only intramolecular but also intermolecular steric effects and hydrophobic/hydrophilic interactions between the aryl substituents of a reactant and its surrounding can be a source of deviation from linear Hammett plots (e.g., [36,37,38,39]). These contributions to non-linearity are especially relevant when studying mechanisms of enzymatic reactions since these take place within a tight binding site where the degree of interactions of a particular substituent with amino acid residues of the immediate surrounding can be considerable and distinct from that in the bulk solvent. In the most severe cases, the rate-determining step or even the reaction mechanism itself can change with the substituent [40,41]. Therefore, could the differing kinetic data (that provide indirect insight into the catalytic mechanism of MAO) mostly reflect the steric and/or other intermolecular interactions imparted by a distinct catalytic site (Figure 2) which tilts the apparent activity of MAO A or MAO B to one or the other kinetic pattern in an otherwise identical catalytic mechanism?

3. The Oxidative Half-Reaction: How Does Oxygen Access the Reduced Flavin, and Is It Blocked or Facilitated by Ligands?

The regeneration of oxidized MAO, which completes MAO catalytic cycle, has been less intensely studied. At a minimum, more experiments are needed to explain the very different KM values for oxygen (MAO A at 0.06 mM but MAO B at 0.28 mM [43]). The kinetics of oxidation of the amine substrate and the accompanying reduction of the FAD in MAO have been studied in detail [44,45,46,47,48], but very little is known about how oxygen accesses the reduced flavin to restore the oxidized FAD and produce hydrogen peroxide (H2O2), the product of the second half-reaction and a potential cause of degenerative effects [49].
In common with other oxidases, MAO has the FAD:N(5)-water-Lys (305 in MAO A) motif thought to be key for oxygen activation [50]. By positioning oxygen in high affinity sites, the likelihood of O2 reacting with the correct stereochemistry with FAD is increased and generation of undesirable reactive oxygen species is minimized. The reaction takes place between oxygen and the C4a position of the flavin. Key factors influencing the reaction include accessibility, the distribution of charged and polar groups to favor electron transfer between the two reactants, and the extent of stabilization of the C4a-hydroperoxyflavin intermediate [51,52].
How does the oxygen get to the deeply buried site? In other oxidases, multiple pathways have been identified [51]. Oxygen diffusion to the reduced co-factor is most likely via tunnels that allow regulation of oxygen access [51]. Studies on MAO similar to those that revealed multiple pathways in cholesterol oxidases [53] would be useful. Could similar pathways in MAO provide insight into the substrate acceleration of re-oxidation (see below)? Moreover, induced fit with substrate alters the O2 locations and opens alternative channels [54]. Where are the oxygen access channels in MAO and does substrate or product binding to reduced MAO change access or activation?

4. Kinetics: Can Simulation Explain the Alternative Pathways in Turnover; Why Substrate Accelerates Oxidation and the Influence of Two Forms for Binding Inhibitors?

4.1. Both the Oxidized and Reduced Forms of MAO Can Bind Ligands during Turnover

Early work varying both benzylamine and oxygen indicated a ping-pong mechanism for rat MAO [55,56], but later kinetic studies demonstrated that some substrates followed a ternary complex mechanism (Figure 3) [45,57,58]. Spontaneous hydrolysis of the imine product to the aldehyde plus ammonia, at least some of which may occur in the active site, complicates the picture [59]. However, hydrolysis of imine in the active site is probably not the prevailing mechanism, but rather, it is primarily non-enzymatically hydrolyzed outside the enzyme [60]. The number (and orientation) of water molecules in the active site depends on the particular isozyme (MAO A versus MAO B) and the nature of the substrate [49,61,62]. Many water molecules can be of structural type (e.g., involved in hydrogen bonding), making them very unlikely to take part in any type of reaction [63]. Nonetheless, imine hydrolysis in the active site can still be considered plausible. Aldehydes are very poor uncompetitive inhibitors [60] so can be considered to dissociate easily, but does ammonia dissociate more slowly? With a pKa of ammonia being 9.2, it is not difficult to imagine a proton abstraction by ammonia from an acidic amino acid residue in the active site, resulting in a protonated species which can be effectively “trapped” by favorable cation-π interactions with the tyrosine residues that constitute the “aromatic cage” [63,64]. From results using a benzylamine analog oxidized to a spectrally distinct imine product, it has been argued that the reduced enzyme-imine complex is reoxidized before the bound imine signal is lost [59]. If either ammonia or imine remained bound, free reduced enzyme would be minimal, and no new ligand could bind. However, the kinetics demonstrating binding of inhibitor to reduced enzyme indicate that free reduced enzyme is present during turnover [57,58,65].
Ultimately, whether a pool of free reduced enzyme builds up during turnover will depend on the rate constants for the half-reactions, the concentrations of the substrates (amine and oxygen), and the affinities of the oxidized (Eo) and reduced (Er) forms of the enzyme for their ligands (see Figure 3). Although reduction of the FAD by the amine substrate is the slowest step in the reaction catalyzed by MAO, there are three possible pathways for the reoxidation of the reduced FAD co-factor. The rate of reoxidation of the reduced MAO B without ligand is slow (kox, 1.8 s−1 at 0.28 mM oxygen) and whether Ered-P is oxidized before dissociation of the imine product is questioned. When some substrates are preincubated with reduced MAO, the rate of oxidation increases 2–100 times [45]; so for those substrates, the pathway in the green box dominates. Based on data from [45], 2-phenylethylamine (PEA) reduces bovine MAO B at 572 s−1 (kred), rapidly generating Ered, but the steady-state turnover rate is 3.6 s−1. The imbalance in rates means that the MAO is mostly reduced (Ered) during steady-state oxidation of PEA. Based on steady-state and half-reaction studies, oxidation of PEA is mainly via the binary pathway. With benzylamine as substrate, the reduction and reoxidation rates are similar (10.9 and 7.6 s−1, respectively), and reoxidation is mainly via the ternary complex pathway indicated by the green rectangle. It should be noted that reoxidation rates are influenced by the concentration of oxygen ([O2]). Even with these similar rates of reduction and oxidation, some intermediate Ered is available for binding of inhibitor as has been demonstrated in bovine MAO B [57] and for human MAO B [65]. That the affinities of the oxidized (Eo) and reduced (Er) forms of the enzyme for inhibitors can differ was demonstrated for the tricyclic antidepressant amitriptyline which gives a Ki of 0.5 mM with Eo but 2.5 mM with Er [66]. The question is important for the drug design process because IC50 measurements used to screen new compounds will vary with the proportion of reduced enzyme present during turnover which depends upon the assay and substrate used [65,66].
The occurrence of differential binding to Eo and Er was established by in-depth kinetic investigations in vitro [65]. In laboratory assays, the O2 concentration in buffer at 30 °C (0.22 mM) is below the KM for O2 with MAO B (0.28 mM) [43], and the oxygen content drops with temperature. Although the kinetic isotope effects in air-saturated buffer (21.1% or 101.3 kPa oxygen) of 5–10 show that amine oxidation is the slower step [46], the rate of reoxidation at sub-saturating oxygen conditions will influence turnover more as the oxygen concentration decreases below its KM, especially at low substrate (amine) levels. (See more information on the kinetics of oxidation below.) In cell culture at 37 °C and 21% oxygen, a dissolved extracellular concentration of 0.21 mM was measured, but depending on conditions, oxygen concentration intracellularly was only 0.175 mM [67]. However, this is now considered hyperoxia rather than physiological normoxia. Human tissues have oxygen concentrations varying around 6%, but inside the cell, it is 1.3–2.5% [67], one-tenth of that for in vitro experimental concentrations. Could simulations explore MAO kinetics and redox status (proportions of Eo and Er) in the brain?

4.2. Substrate Stimulation of Reoxidation-Thermodynamic or Kinetic?

That the slowest step in the MAO reaction is the reductive half-reaction (oxidation of substrate amine to imine) is clear from studies involving deuterated substrates where large kinetic isotope effects of 5–14 were obtained [25,46,48,68,69,70,71]. However, the oxidative half-reaction of MAO measured in isolation is slower than the observed steady-state turnover rate for many substrates [45]. By pre-incubating chemically reduced MAO with substrate before rapidly mixing it with oxygen, it was demonstrated that the substrate stimulates the rate of reoxidation, whereas, for MAO A, a product (MPP+) or an inhibitor (D-amphetamine) inhibit the rate. Moreover, different substrates give different increases in the slow rate of reoxidation of free reduced enzyme (1 s−1) found with both MAO A and MAO B. For MAO A with kynuramine, the rate of flavin re-oxidation was 120 s−1, with benzylamine 23 s−1 and with MPTP 40 s−1. For MAO B, the rate of re-oxidation of the flavin in enzyme saturated with kynuramine was 2.2 s−1, benzylamine 7.6 s−1, and MPTP 6.0 s−1 [45]. Therefore, is it the imine or ammonia still bound that facilitates the oxidation (although MPP+ clearly inhibits) or does the binding of new substrate to the reduced enzyme govern the kinetics of oxidative half-reaction? Inhibition kinetics [57] do show that free enzyme is available before re-oxidation of the reduced flavin, so perhaps substrate acceleration is the dominant pathway for the normal substrates, but how is the effect achieved?
Measurement of the redox potential of the reductive half-reaction for MAO A in the absence and presence of substrate suggested that substrate increased the redox potential [72] but similar experiments in MAO B did not [73]. A change in redox potential would indicate a change in the conformation or environment of the FAD, but it is more likely that the reported MAO A change was due to lack of equilibrium between the dye and FAD. However, the environment of the flavin has long been known to influence the redox potential [74], and substrate (or inhibitor) binding is likely to alter that environment. Indeed, the shifts in the circular dichroism spectrum indicated differences in the FAD region when substrate was added [75]. The differences between the oxidized and reduced circular dichroism spectra also indicate a change in the FAD environment or shape (see below).

5. Is There Potential for Applied Use of MAO?

MAO turnover is slow (3–10 s−1 [45]) but specific (the pro-R hydrogen from the α carbon next to the amine is removed [76]). Tighter substrate cavity of MAO B results in increased steric hindrance and in distinctive amino acid residue interactions with small molecules. This makes MAO B more sensitive to the absolute configuration at chiral centers of enantiomeric substrates and inhibitors [20] with potential for high value chemical synthesis.
Specificity is also governed by hydrophobic and hydrophilic regions of the cavity. The environments of both substrate cavities are generally hydrophobic, but MAO B is slightly more so. Consequently, by transforming some regular MAO B inhibitors into quaternary ammonium salts, selective MAO A inhibitors were obtained [77]. Within each cavity, docking of variants within the many classes of inhibitors has demonstrated the importance of specific interactions such as hydrogen bonding. Specially, MAO A Phe208-Ile335 and MAO B Ile199-Tyr326 pairs (see Figure 2) were suggested to be major determinants which govern the differential inhibitor specificities of the two isozymes [78]. Site-directed mutagenesis studies showed that Ile335 in MAO A and Tyr326 in MAO B are key amino acid residues determining substrate and inhibitor specificities in human and rat MAO A and B [79,80,81]. By tempering with these two residues, the spatial and chemical architecture of the substrate cavity is altered, and MAO B selectivity starts to mimic that of MAO A and vice versa. A change in selectivity was also induced by the single mutation Phe208 to Ile199 [82]. However, the wide range of effective inhibitors demonstrates that affinity for a ligand is determined by multiple interactions within the large cavity (for examples, see recent reviews: [83,84,85]).
Purified MAO is not stable enough for most applications, but MAO electrodes incorporating membrane-bound MAO have been devised for measurement of amines, for example, for serotonin using MAO A [86,87]. For electrode use, an alternative to oxygen as the electron acceptor in the oxidative half-reaction is desirable. Methylene Blue and similar dyes are known acceptors [88,89], but performance could be improved by rational engineering of the oxygen pathway or reaction site [90]. This has been demonstrated for MAO A by mutating Lys305: after reduction by serotonin, the K305M MAO A reoxidation by oxygen was much slower but electrons could pass instead to a quinone [91].
In contrast to the mitochondrial MAO, the Aspergillus niger enzyme, MAO N is more tractable for protein engineering and large-scale production although the FAD is not covalently bound to the enzyme [92,93,94]. MAO N is suitable for chiral catalysis and use of whole cells avoids the limitation of FAD supply [95,96]. Whole cell methodologies might be the best way forward for any use of mitochondrial MAO, but tuning MAO N to the job seems the better option.
Perhaps also of interest for applications in biological tissues or in vivo is the observation that geometric isomers of cis- and trans-1-propargyl-4-styrylpiperidine distinguish between MAO A and MAO B. The cis isomers are potent human MAO A irreversible inhibitors whereas the trans analogues inactivate only MAO B both in vitro and in mouse brain [97].
Overall, drug design remains the major applied opportunity for treatment of depression, neurodegeneration or cancer [98]. Identification of key interactions at the catalytic site fuels innovation towards new lines of MAO inhibiting drugs with enhanced selectivity and potency.

6. Importance of Flexibility

6.1. Is Flexibility Important in the Reaction?

Crystal structures present a frozen view of enzymes, but from this accurate foundation, rapidly advancing computational approaches reveal a much more dynamic nature of enzymes in their physiologic environment, including, for MAO, at the interface of the cytosol and the mitochondrial outer membrane. Experimentally, MAO flexibility is seen as changes in the circular dichroism spectra of both flavin and aromatic amino acid environments after ligand binding or redox change [75]. An early modeling test (on the MAO B structure from which the reversible inhibitor, isatin, had been removed) fixing the amino acids around the flavin indicated no hindrance to flattening of isoalloxazine ring from the 30° bend seen in the crystal structure of the enzyme-bound flavin [99]. The isoalloxazine ring in many flavoproteins changes from flat to a bent ring upon reduction, but all crystal structures of MAO whether oxidized (reversible inhibitor bound) or reduced (irreversible inhibitor attached to the FAD) show a 30° bend along the N(5)-N(10) axis [78]. It is hypothesized that this strain at the coenzyme binding site may have catalytic relevance, mainly increasing the nucleophilicity of N(5) and electrophilicity of C(4a) [100]. Any movement of FAD or protein residues could also alter the oxygen accessibility, crucial for MAO turnover. If structural perturbations could be used effectively to steer the electronic characteristics of the cofactor and thus regulate the catalytic activity, why does flavin in MAO A and MAO B adopt a bent position in crystals regardless of its redox state?
Turning to the flexibility of the protein, induced fit within the active site is evidenced by the slow onset of full inhibition by larger reversible inhibitors (even 1-methyl-4-phenyl-pyridinium (MPP+) for which perturbation of the flavin spectrum showed an on-rate of 108). Slow on (30 min) and off (6 washes) rates were also observed in reversibility experiments with safinamide [101]. Molecular dynamics now confirm considerable movement in active site in addition to alternate binding modes obtained by docking. Enzyme tertiary structure adjustments of the active site environment are also necessary to accommodate the irreversibly bound inhibitors, which cannot adopt the most stable structural conformation otherwise [102]. Therefore, what constitutes the line between MAO flexibility and rigidity allowing for a number of structurally distinct (but related) substrates to be oxidized by the enzyme while assuring that certain biological functions are properly regulated on account of substrate specificity?
Flexibility (and rigidity) could also be important for crosstalk between the monomers in the MAO dimer. Pulsed electron paramagnetic studies demonstrated that both MAOs from human and rat are dimeric in the membrane [103]. Does dimer crosstalk play a role in kinetics or in ligand binding? Although there was some speculation in older literature about occupancy of one site changing the activity of the other, this topic needs evidence.

6.2. The Influence of the Membrane on MAO Activity

Simple steady-state kinetic experiments showed that membrane-bound MAO gave a lower KM for some substrates compared to purified enzyme. For example, MAO A had a 10-fold lower reversible Ki for clorgyline [104]. In experiments studying solvent accessibility to the active site using spin-labeled derivatives of the inactivator, pargyline, no difference between detergent-extracted and membrane-bound MAO was found [105]. However, lower Ki values for the initial reversible inhibition of MAO by propargylamine inactivators in membrane preparations was demonstrated in kinetic studies. The close association of MAO with the membrane and the close proximity of the entrance to the active site to the membrane surface might be expected to influence the dynamics of substrate access or product exit due to steric hindrances and the flexibility of the substrate entrance loop (residues 99–112) [106,107]. When this question was examined using molecular dynamic simulations, effects were indeed found. The membrane influenced the fluctuations of two loops at the entrance to the active site cavity (residues 85–110 and 155–165 in MAO B) [108]. Membrane control of the active site access was reinforced by the differences in the temperature-dependence of benzylamine oxidation by MAO B between purified enzyme and the same enzyme attached to nanodiscs [109]. (However, it should be noted that the preparation used for these experiments had a completely different spectrum from MAO B purified in at least three other laboratories.) Exploring the influence of the membrane using long molecular dynamic simulations, the same authors looked for oxygen tunnels leading to the FAD and found two entrances, one on each side of the gating loop (residues 99–112). Residues at both entrances are closely associated with the membrane. Given the insight from these computational studies, new investigation of previously observed lipid effects on rat mitochondria MAO could provide further understanding of the importance in vivo [110,111,112,113,114]. Could alterations in the membrane composition (from diet, for example) have consequences for activity, raising or lowering the monoamine pool in the brain?

6.3. Does the Oxidation of MAO Thiols Have Any Importance In Vivo?

Continuing the theme of biological influence from chemical changes, do the solvent accessible cysteines influence activity? Both MAO A and MAO B form a covalent 8α-thioether linkage with the flavin (at Cys407 and Cys398, respectively) and contain 8 other cysteine residues in reduced (SH) form. Although 7 of the 9 cysteines are in conserved sequence locations, only MAO B has one that interacts with ligand in the active site (Cys172) [115,116]. Free thiol groups in proteins are subject to various modifications from reaction with reactive species generated by mitochondria [117,118] and thiol modification can both regulate and damage cell function [119]. Although located on the cytosolic surface of the mitochondrial outer membrane, the 8 modifiable cysteines in MAO [120,121,122] are close to the greatest cellular source of reactive oxygen species, the electron transport chain on the inner membrane. The initial study where each Cys was mutated to serine demonstrated individual effects of each on MAO activity. Most of the mutants behaved like the wild-types, but complete loss of activity was observed for MAO A Ser374 and MAO B Ser156 and Ser365 mutants [123]. Later, Cys365 was labeled specifically during inactivation of MAO B with N-cyclopropyl-N-α-methylbenzylamine, indicating its involvement in the inhibitory pathway [124]. The equivalent Cys374 in MAO A substituted with alanine gave kcat/KM values 30% lower than for the wild type [122]. Chemical modification of the thiols revealed different susceptibility: MAO A was more rapidly inactivated by N-ethylmaleimide (half-life about 3 min) than MAO B (half-life about 8 h) [121]. Moreover, Cys266 and Cys5 seem to be the only two thiols modified in clorgyline-inactivated MAO A and pargyline-inactivated MAO B, respectively, which demonstrates that inhibitors, as well as substrates, stabilize the enzyme structure and increase the rigidity [116]. Does occupancy of the active site by the amine substrate prevent oxidation of cysteine residues in mitochondrial MAO and to what extent?
The presence of reactive oxygen species around mitochondria in the cell can oxidize protein thiols in the vicinity, so thiol modification could be significant for catalysis under physiological conditions. Further development and application of chemical methods to detect, distinguish, and time modify cysteines in the cell or tissues would allow better understanding of any consequences from modifications of cysteines in MAO [119,125,126].

7. Biological Questions

7.1. Does Variation in the Calcium Ion Concentration Change MAO Activity?

In neurons, ion fluxes during the action potential result in transient changes in the calcium ion concentration in the cytosol, but uptake by the endoplasmic reticulum and mitochondria quickly restore the normal low micromolar level. Calcium (Ca2+) was found to increase MAO activity in mitochondria and cells [127,128,129,130]. When a short burst of Ca2+ was induced in mice by a calcium channel agonist, the increase in MAO A activity was transient [130]. In vitro, preincubation of membrane-bound hMAO with 1 mM Ca2+ increased the MAO A maximum velocity 2-fold but did not change MAO B [131]. No other divalent cation increased activity but zinc ion was strongly inhibitory. More physiologically relevant, a 20% increase in activity was observed when Ca2+ was increased from 10 nM to 10 uM, the normal range for calcium signaling. How does Ca2+ exert this influence on activity? Is the ion size the determining factor? Why is only MAO A activated? A calcium-sensitive increase in MAO A activity increased the production of peroxyradicals in hippocampal cultures, and this was postulated to have implications for neuronal damage in Alzheimer disease [132]. Is this calcium-sensitive stimulation of MAO A activity increase relevant in the cell and in pathology?

7.2. Does MAO A Moonliaght in the Cell?

MAO metabolizes neurotransmitter amines in nervous tissue and biogenic amines in all cells so that alterations in activity alter signaling amine levels. Sustained changes in monoamine levels trigger adaptations in levels of other proteins, including receptors. For MAO A itself, putative non-catalytic roles for the protein have been proposed, involving either direct protein-protein interaction or a signaling relay influencing gene expression. The association of some Alzheimer’s disease (AD)-related presenilin-1 variants with depression led to studies that demonstrated a direct interaction between presenilin-1 and MAO A proteins and a change in MAO A activity [133,134]. MAO A knockout mice revealed the importance of serotonin and MAO A in apoptosis during brain development [135]. MAO A levels are influenced by both genetic and environmental factors, and MAO A activity influences psychopathologies [136,137]. The anorexigenic hormone leptin that protects mitochondrial function regulates the expression and activity of MAO B [138]. Leptin also reduced the increase in MAO A and MAO B expression found in hippocampal neurons treated with the Alzheimer’s disease peptide Aβ. In Alzheimer’s disease, which leads to cell loss in the brain, an increase in MAO B was evident but MAO A also increased in some regions. Intriguingly, co-regulation of MAO A and MAO B levels observed in normal post-mortem brains was lost in the hippocampus but not in the cortex of Alzheimer brains [139].
When MAO A activity in human neuroblastoma cells was knocked down to about 30% that of control cells, ATP was increased, reactive oxygen species were decreased, and the cells were protected against apoptosis. The effects were linked to electron transport because although inhibition of electron transport up-regulated MAO A in control cells, the low MAO A activity in the knock-down cells protected mitochondria from inhibition [140].
Although a genomic study found MAO A to be down-regulated in some cancers [141], it is clearly elevated in aggressive prostate cancer and classical Hodgkin’s lymphoma [142,143]. A small Phase 2 clinical trial found that the irreversible MAO inhibitor, phenelzine, successfully decreased markers of prostate cancer, opening a new application for MAO A inhibitors [144].
Keeping in mind the location of MAO at the interface between the cytosol and mitochondria, untangling the many interactions for MAO and the network of signaling and gene expression will require considerably more cellular and genetic research.

7.3. Species Differences and Mutations?

Kinetic studies identified differences in inhibitor binding to MAOs from different species [145,146,147,148,149]. Although mammals all have two forms of MAO, fish have only one [150]. After the structures of MAOs were published, a detailed examination of the kinetics and structures of MAOs from rat and human provided better understanding of the differences [105,151]. With different types of cells used in many pharmacological and toxicology studies, researchers should be aware also of differential expression of MAO A or B in each cell type [152]. The question in this section is curiosity-driven: species differences are seen in the active sites of MAO A and MAO B, but human mutations altering active site amino acid residues to influence activity or ligand binding have not been identified from clinical studies. Known human variants identified by genetic studies are sequence repeats that alter expression levels with behavioral consequences [153,154,155,156,157].
One theoretical study has provided the basis for a lower activity based on a mutation in the middle of the MAO B active site. When tyrosine 326 in MAO B is changed to isoleucine (as found in MAO A-so the mutant oxidizes serotonin better), the activity (kcat) with phenylethylamine dropped to 27% of the wild-type, and the KM increased [79]. When the oxidation of the amine was simulated, reaction free energy profiles showed that the barrier in the mutant was 1.06 kcal mol−1 higher, due to loss of electrostatic interactions and dielectric shielding [158]. Are there similar variants in the human populations and could new experiments and clinical studies demonstrate consequences of the lower activity giving higher amine levels in the brain for human behavior?

7.4. The Pharma Debate-Reversible or Irreversible Inhibition?

The effect of a successful drug is governed by binding affinity and by pharmacokinetics. For a high affinity reversible inhibitor, side effects should be low, and the inhibition is easily reversed. Indeed, it might be desirable that a surge in the amine substrate concentration would out-compete the inhibitor. For irreversible inhibitors, reaching a steady-state level of inhibition might take longer if the dose is low, but the effect would persist until the inactivated enzyme was replaced by a new protein molecule.
To achieve the desired clinical effect in depression or Parkinson’s disease, MAO inhibition must be greater than 75% [159]. Despite the rapid inhibition of MAO that would occur in vitro (immediate for reversible inhibitors and minutes for irreversible inhibitors) and the detection of MAO using 11C-deprenyl in positron emission tomography (PET) scans recorded from 25 min after the infusion [160], clinical improvement is slow [161]. At the end of treatment, a “wash-out” period of three weeks is recommended for the irreversible inhibitors to allow the synthesis of new protein, so activity resumes at the slow pace of turnover-the half-life of MAO B in baboons was measured by PET using labeled L-deprenyl as 30 days [162]. For reversible inhibitors, clearance of the drug from circulation should result in rapid return of MAO activity, but a minimum of two weeks washout is still recommended.
The question of whether a reversible MAOI or an irreversible MAOI provides better treatment has mostly been investigated for Parkinson’s disease for which the irreversible selegiline (or the similar drug, rasagiline) and the reversible safinamide are commonly used to delay the need for levodopa (l-dopa) treatment. Selegiline has neuroprotective effects [163] and antidepressant action, but tyramine potentiation fears and dietary restrictions persist, except for the transdermal form. Safinamide has other effects beneficial against PD including blockage of voltage-dependent sodium channels and has low tyramine potentiation [164]. When reversible and irreversible inhibitors are compared, both have been found to be effective as adjuvant therapy to delay or reduce the need for l-dopa [165]. However, a meta-analysis comparing MAOI alone or with l-dopa showed that selegiline was the most effective drug and that both it and rasagiline were more effective than safinamide (selegiline > rasagiline > safinamide) [166].
Further study of clinical outcomes with irreversible versus reversible inhibitors in depression would also be useful. Reuptake inhibitors are first choice for depression, then reversible MAOI, while irreversible inhibitors are used when other treatments fail. This is partly due to the side effects that require dietary restrictions but probably more due to inexperience with these drugs [167,168]. Certainly for Parkinson’s disease, prescribing habits vary from country to country [165].
Much recent effort has gone into designing reversible MAOI, especially for MAO B and Parkinson’s disease. Many scaffolds are used, but particularly successful compounds have been found amongst coumarin derivatives, which are competitive inhibitors with inhibition constants as low as 0.25 nM [169,170,171,172]. However, what if irreversible inhibitors provide more stable inhibition of MAO, especially in contrast to drop in inhibition unless there is strict compliance with the dosing regimen for reversible drugs? Certainly, multi-target compounds designed for use in multi-factorial neurodegeneration such as Alzheimer’s disease have employed the propargylamine moiety to give irreversible MAO inhibition leaving more flexibility to adapt the rest of the molecule to give high affinity binding to other targets such as the cholinesterases. For example, the propargylamine moiety for MAO inactivation has been combined with inhibition of the cholinesterases and antagonism at histamine H3 receptors in one molecule with suitable drug-like properties [18,173]. However, do these multi-target species really “outperform” a drug mixture composed of individual inhibitors in clinical use?
Therefore, what should drug research for MAO seek: reversible or irreversible? It is still a debate, but the original irreversible classes on MAO inhibitor drugs are still an essential part of clinical practice, with the reversible drugs appearing somewhat less effective.

8. Conclusions

After 100 years of research on the MAO enzymes and 60 years on the MAOI, there remain still interesting questions to explore for neurologists, pharmacologists, enzymologists, computational experts and chemists. Have fun!

Funding

A.A. is supported by the Slovenian Research Agency (Research Program P1-0005).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Edmondson, D.E.; Binda, C.; Mattevi, A. The FAD binding sites of human monoamine oxidases A and B. NeuroToxicology 2004, 25, 63–72. [Google Scholar] [CrossRef]
  2. Hare, M.L.C. Tyramine oxidase: A new enzyme system in liver. Biochem. J. 1928, 62, 968–979. [Google Scholar] [CrossRef]
  3. Tipton, K.F. 90years of monoamine oxidase: Some progress and some confusion. J. Neural. Transm. 2018, 125, 1519–1551. [Google Scholar] [CrossRef]
  4. Kim, T.; Xu, C.; Amsterdam, J.D. Relative effectiveness of tricyclic antidepressant versus monoamine oxidase inhibitor monotherapy for treatment-resistant depression. J. Affect. Disord. 2019, 250, 199–203. [Google Scholar] [CrossRef]
  5. Mann, J.J.; Aarons, S.F.; Wilner, P.J.; Keilp, J.G.; Sweeney, J.A.; Pearlstein, T.; Frances, A.J.; Kocsis, J.H.; Brown, R.P. A controlled-study of the antidepressant efficacy and side-effects of (−)-deprenyl—A selective monoamine-oxidase inhibitor. Arch. Gen. Psychiatry 1989, 46, 45–50. [Google Scholar] [CrossRef]
  6. Quitkin, F.; Rifkin, A.; Klein, D.F. Mono-amine oxidase-inhibitors—Review of anti-depressant effectiveness. Arch. Gen. Psychiatry 1979, 36, 749–760. [Google Scholar] [CrossRef]
  7. Maycock, A.L.; Abeles, R.H.; Salach, J.I.; Singer, T.P. Structure of covalent adduct formed by interaction of 3-dimethylamino-1-propyne and flavin of mitochondrial amine oxidase. Biochemistry 1976, 15, 114–125. [Google Scholar] [CrossRef]
  8. Gartner, B.; Hemmerich, P.; Zeller, E.A. Structure of flavin adducts with acetylenic substrates—Chemistry of monoamine-oxidase and lactate oxidase Inhibition. Eur. J. Biochem. 1976, 63, 211–221. [Google Scholar] [CrossRef] [PubMed]
  9. Binda, C.; Newton-Vinson, P.; Hubalek, F.; Edmondson, D.E.; Mattevi, A. Structure of human monoamine oxidase B, a drug target for the treatment of neurological disorders. Nat. Struct. Biol. 2002, 9, 22–26. [Google Scholar] [CrossRef] [PubMed]
  10. Silverman, R.B. Radical ideas about monoamine-oxidase. Acc. Chem. Res. 1995, 28, 335–342. [Google Scholar] [CrossRef]
  11. Kalgutkar, A.S.; Castagnoli, N.; Testa, B. Selective inhibitors of monoamine-oxidase (MAO-A and MAO-B) as probes of its catalytic site and mechanism. Med. Res. Rev. 1995, 15, 325–388. [Google Scholar] [CrossRef] [PubMed]
  12. Finberg, J.P.M.; Gillman, K. Selective inhibitors of monoamine oxidase type B and the “cheese effect”. Int. Rev. Neurobiol. 2011, 100, 169–190. [Google Scholar] [CrossRef] [PubMed]
  13. Youdim, M.B.H.; Edmondson, D.; Tipton, K.F. The therapeutic potential of monoamine oxidase inhibitors. Nat. Rev. Neurosci. 2006, 7, 295–309. [Google Scholar] [CrossRef] [PubMed]
  14. Ramsay, R.R. Molecular aspects of the activity and inhibition of the FAD-containing monoamine oxidases. In Pharmaceutical Biocatalysis: Fundamentals, Enzyme Inhibitors, and Enzymes in Health and Diseases; Grunwald, P., Ed.; Pan Stanford Publishing Pte. Ltd.: Singapore, 2019; Volume 4. [Google Scholar]
  15. Ramsay, R.R.; Maniquet, A.; Hagenow, S.; Pappalardo, M.; Saija, M.C.; Bryant, S.D.; Albreht, A.; Guccione, S. Molecules Parameters for irreversible inactivation of monoamine oxidase. Molecules 2020, 25, 5908. [Google Scholar] [CrossRef] [PubMed]
  16. Binda, C.; Wang, J.; Li, M.; Hubalek, F.; Mattevi, A.; Edmondson, D.E. Structural and mechanistic studies of arylalkylhydrazine inhibition of human monoamine oxidases A and B. Biochemistry 2008, 47, 5616–5625. [Google Scholar] [CrossRef] [PubMed]
  17. Hubalek, F.; Binda, C.; Li, M.; Herzig, Y.; Sterling, J.; Youdim, M.B.H.; Mattevi, A.; Edmondson, D.E. Inactivation of purified human recombinant monoamine oxidases A and B by rasagiline and its analogues. J. Med. Chem. 2004, 47, 1760–1766. [Google Scholar] [CrossRef]
  18. Bautista-Aguilera, O.M.; Hagenow, S.; Palomino-Antolin, A.; Farre-Alins, V.; Ismaili, L.; Joffrin, P.L.; Jimeno, M.L.; Soukup, O.; Janockova, J.; Kalinowsky, L.; et al. Multitarget-Directed Ligands Combining Cholinesterase and Monoamine Oxidase Inhibition with Histamine H3R Antagonism for Neurodegenerative Diseases. Angew. Chem. Int. Ed. 2017, 56, 12765–12769. [Google Scholar] [CrossRef] [Green Version]
  19. Sablin, S.O.; Krueger, M.J.; Singer, T.P.; Bachurin, S.O.; Khare, A.B.; Efange, S.M.N.; Tkachenko, S.E. Interaction of Tetrahydrostilbazoles with Monoamine-Oxidase-a and Monoamine-Oxidase-B. J. Med. Chem. 1994, 37, 151–157. [Google Scholar] [CrossRef] [PubMed]
  20. Bocchinfuso, R.; Robinson, J.B. The stereoselectivity of inhibition of rat liver mitochondrial MAO-A and MAO-B by the enantiomers of 2-phenylpropylamine and their derivatives. Eur. J. Med. Chem. 1999, 34, 293–300. [Google Scholar] [CrossRef]
  21. Edmondson, D.E.; DeColibus, L.; Binda, C.; Li, M.; Mattevi, A. New insights into the structures and functions of human monoamine oxidases A and B. J. Neural. Transm. 2007, 114, 703–705. [Google Scholar] [CrossRef]
  22. Chajkowski-Scarry, S.; Rimoldi, J.M. Monoamine oxidase A and B substrates: Probing the pathway for drug development. Future Med. Chem. 2014, 6, 697–717. [Google Scholar] [CrossRef]
  23. Lu, X.; Rodriguez, M.; Ji, H.; Silverman, R.B.; Vintém, A.-P.B.; Ramsay, R.R. Irreversible inactivation of mitochondrial monoamine oxidases. In Flavins and Flavoproteins 2002; Chapman, S.K., Perham, R.N., Scrutton, N.S., Eds.; Rudolf Weber: Berlin, Germany, 2002; pp. 817–830. [Google Scholar]
  24. Erdem, S.S.; Buyukmenekse, B. Computational investigation on the structure-activity relationship of the biradical mechanism for monoamine oxidase. J. Neural. Transm. 2011, 118, 1021–1029. [Google Scholar] [CrossRef]
  25. Miller, J.R.; Edmondson, D.E. Structure-activity relationships in the oxidation of para- substituted benzylamine analogues by recombinant human liver monoamine oxidase A. Biochemistry 1999, 38, 13670–13683. [Google Scholar] [CrossRef]
  26. Orru, R.; Aldeco, M.; Edmondson, D.E. Do MAO A and MAO B utilize the same mechanism for the C–H bond cleavage step in catalysis? Evidence suggesting differing mechanisms. J. Neural. Transm. 2013, 120, 847–851. [Google Scholar] [CrossRef] [Green Version]
  27. MacMillar, S.; Edmondson, D.E.; Matsson, O. Nitrogen kinetic isotope effects for the Monoamine Oxidase B-catalyzed oxidation of benzylamine and (1,1-(2)H(2))benzylamine: Nitrogen rehybridization and CH bond cleavage are not concerted. J. Am. Chem. Soc. 2011, 133, 12319–12321. [Google Scholar] [CrossRef] [PubMed]
  28. Zenn, R.K.; Abad, E.; Kastner, J. Influence of the environment on the oxidative deamination of p-substituted benzylamines in monoamine oxidase. J. Phys. Chem. B 2015, 119, 3678–3686. [Google Scholar] [CrossRef] [PubMed]
  29. Umhau, S.; Pollegioni, L.; Molla, G.; Diederichs, K.; Welte, W.; Pilone, M.S.; Ghisla, S. The x-ray structure of D-amino acid oxidase at very high resolution identifies the chemical mechanism of flavin-dependent substrate dehydrogenation. Proc. Natl. Acad. Sci. USA 2000, 97, 12463–12468. [Google Scholar] [CrossRef] [Green Version]
  30. Edmondson, D.E.; Binda, C.; Mattevi, A. Structural insights into the mechanism of amine oxidation by monoamine oxidases A and B. Arch. Biochem. Biophys. 2007, 464, 269–276. [Google Scholar] [CrossRef] [Green Version]
  31. Akyuz, M.A.; Erdem, S.S. Computational modeling of the direct hydride transfer mechanism for the MAO catalyzed oxidation of phenethylamine and benzylamine: ONIOM (QM/QM) calculations. J. Neural. Transm. 2013, 120, 937–945. [Google Scholar] [CrossRef] [PubMed]
  32. Oanca, G.; Stare, J.; Vianello, R.; Mavri, J. Multiscale simulation of monoamine oxidase catalyzed decomposition of phenylethylamine analogs. Eur. J. Pharmacol. 2017, 817, 46–50. [Google Scholar] [CrossRef]
  33. Prah, A.; Purg, M.; Stare, J.; Vianello, R.; Mavri, J. How Monoamine Oxidase A Decomposes Serotonin: An Empirical Valence Bond Simulation of the Reactive Step. J. Phys. Chem. B 2020, 124, 8259–8265. [Google Scholar] [CrossRef]
  34. Vianello, R.; Repic, M.; Mavri, J. How are Biogenic Amines Metabolized by Monoamine Oxidases? Eur. J. Org. Chem. 2012, 36, 7057–7065. [Google Scholar] [CrossRef]
  35. Youngster, S.K.; McKeown, K.A.; Jin, Y.Z.; Ramsay, R.R.; Heikkila, R.E.; Singer, T.P. Oxidation of analogs of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine by monoamine oxidase-A and oxidase-B and the inhibition of monoamine oxidases by the oxidation products. J. Neurochem. 1989, 53, 1837–1842. [Google Scholar] [CrossRef] [PubMed]
  36. Taft, R.W. Linear Steric Energy Relationships. J. Am. Chem. Soc. 1953, 75, 4538–4539. [Google Scholar] [CrossRef]
  37. Dunford, H.B.; Adeniran, A.J. Hammett ϱσ correlation for reactions of horseradish peroxidase compound II with phenols. Arch. Biochem. Biophys. 1986, 251, 536–542. [Google Scholar] [CrossRef]
  38. Zhang, H.; Dunford, H.B. Hammett pσ correlation for reactions of lactoperoxidase compound II with phenols. Can. J. Chem. 1993, 71, 1990–1994. [Google Scholar] [CrossRef]
  39. Santiago, C.B.; Milo, A.; Sigman, M.S. Developing a Modern Approach to Account for Steric Effects in Hammett-Type Correlations. J. Am. Chem. Soc. 2016, 138, 13424–13430. [Google Scholar] [CrossRef]
  40. Hart, H.; Sedor, E.A. Mechanism of cyclodehydration of 2-phenyltriarylcarbinols. J. Am. Chem. Soc. 1967, 89, 2342–2347. [Google Scholar] [CrossRef]
  41. Stein, A.R.; Tencer, M.; Moffatt, E.A.; Dawe, R.; Sweet, J. Nonlinearity of Hammett.sigma..rho. correlations for benzylic systems: Activation parameters and their mechanistic implications. J. Org. Chem. 1980, 45, 3539–3540. [Google Scholar] [CrossRef]
  42. Juarez-Jimenez, J.; Mendes, E.; Galdeano, C.; Martins, C.; Silva, D.B.; Marco-Contelles, J.; Carreiras, M.d.C.; Javier Luque, F.; Ramsay, R.R. Exploring the structural basis of the selective inhibition of Monoamine Oxidase A by dicarbonitrile aminoheterocycles: Role of Asn181 and Ile335 validated by spectroscopic and computational studies. Biochim. Biophys. Acta-Proteins Proteom. 2014, 1844, 389–397. [Google Scholar] [CrossRef] [Green Version]
  43. Ramsay, R.R.; Tipton, K.F. Assessment of Enzyme Inhibition: A Review with Examples from the Development of Monoamine Oxidase and Cholinesterase Inhibitory Drugs. Molecules 2017, 22, 1192. [Google Scholar] [CrossRef] [Green Version]
  44. Husain, M.; Edmondson, D.E.; Singer, T.P. Kinetic-studies on the catalytic mechanism of liver Monoamine- Oxidase. Biochemistry 1982, 21, 595–600. [Google Scholar] [CrossRef] [PubMed]
  45. Tan, A.K.; Ramsay, R.R. Substrate-specific enhancement of the oxidative half-reaction of Monoamine-Oxidase. Biochemistry 1993, 32, 2137–2143. [Google Scholar] [CrossRef]
  46. Walker, M.C.; Edmondson, D.E. Structure-Activity-Relationships in the Oxidation of Benzylamine Analogs by Bovine Liver Mitochondrial Monoamine- Oxidase-B. Biochemistry 1994, 33, 7088–7098. [Google Scholar] [CrossRef] [PubMed]
  47. Nandigama, R.K.; Edmondson, D.E. Influence of FAD structure on its binding and activity with the C406A mutant of recombinant human liver monoamine oxidase A. J. Biol. Chem. 2000, 275, 20527–20532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Dunn, R.V.; Marshall, K.R.; Munro, A.W.; Scrutton, N.S. The pH dependence of kinetic isotope effects in monoamine oxidase A indicates stabilization of the neutral amine in the enzyme-substrate complex. FEBS J. 2008, 275, 3850–3858. [Google Scholar] [CrossRef] [PubMed]
  49. Edmondson, D. Hydrogen peroxide produced by mitochondrial monoamine oxidase catalysis: Biological implications. Curr. Pharm. Des. 2014, 20, 155–160. [Google Scholar] [CrossRef]
  50. Romero, E.; Castellanos, J.R.G.; Gadda, G.; Fraaije, M.W.; Mattevi, A. Same Substrate, Many Reactions: Oxygen Activation in Flavoenzymes. Chem. Rev. 2018, 118, 1742–1769. [Google Scholar] [CrossRef] [Green Version]
  51. Baron, R.; Riley, C.; Chenprakhon, P.; Thotsaporn, K.; Winter, R.T.; Alfieri, A.; Forneris, F.; van Berkel, W.J.H.; Chaiyen, P.; Fraaije, M.W.; et al. Multiple pathways guide oxygen diffusion into flavoenzyme active sites. Proc. Natl. Acad. Sci. USA 2009, 106, 10603–10608. [Google Scholar] [CrossRef] [Green Version]
  52. Gadda, G. Oxygen Activation in Flavoprotein Oxidases: The Importance of Being Positive. Biochemistry 2012, 51, 2662–2669. [Google Scholar] [CrossRef]
  53. Chen, L.; Lyubimov, A.Y.; Brammer, L.; Vrielink, A.; Sampson, N.S. The binding and release of oxygen and hydrogen peroxide are directed by a hydrophobic tunnel in cholesterol oxidase. Biochemistry 2008, 47, 5368–5377. [Google Scholar] [CrossRef] [Green Version]
  54. Vrielink, A.; Ghisla, S. Cholesterol oxidase: Biochemistry and structural features. FEBS J. 2009, 276, 6826–6843. [Google Scholar] [CrossRef]
  55. Houslay, M.D.; Tipton, K.F. Reaction Pathway of Membrane-Bound Rat-Liver Mitochondrial Monoamine-Oxidase. Biochem. J. 1973, 135, 735–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Houslay, M.D.; Tipton, K.F. Rat-Liver Mitochondrial Monoamine-Oxidase—Change in Reaction-Mechanism on Solubilization. Biochem. J. 1975, 145, 311–321. [Google Scholar] [CrossRef] [Green Version]
  57. Pearce, L.B.; Roth, J.A. Human-Brain Monoamine-Oxidase Type-B—Mechanism of Deamination as Probed by Steady-State Methods. Biochemistry 1985, 24, 1821–1826. [Google Scholar] [CrossRef] [PubMed]
  58. Ramsay, R.R. Kinetic mechanism of Monoamine Oxidase-A. Biochemistry 1991, 30, 4624–4629. [Google Scholar] [CrossRef] [PubMed]
  59. Edmondson, D.E.; Bhattacharyya, A.K.; Walker, M.C. Spectral and kinetic studies of imine product formation in the oxidation of p-(N,N-dimethylamino)benzylamine analogs by monoamine oxidase-B. Biochemistry 1993, 32, 5196–5202. [Google Scholar] [CrossRef] [PubMed]
  60. Woo, J.C.G.; Silverman, R.B. Monoamine oxidase B catalysis in low aqueous medium. Direct evidence for an imine product. J. Am. Chem. Soc. 1995, 117, 1663–1664. [Google Scholar] [CrossRef]
  61. Binda, C.; Mattevi, A.; Edmondson, D.E. Structural properties of human monoamine oxidases A and B. Int. Rev. Neurobiol. 2011, 100, 1–11. [Google Scholar]
  62. Dasgupta, S.; Mukherjee, S.; Mukhopadhyay, B.P.; Banerjee, A.; Mishra, D.K. Recognition dynamics of dopamine to human Monoamine oxidase B: Role of Leu171/Gln206 and conserved water molecules in the active site cavity. J. Biomol. Struct. Dyn. 2018, 36, 1439–1462. [Google Scholar] [CrossRef]
  63. Binda, C.; Hubalek, F.; Li, M.; Herzig, Y.; Sterling, J.; Edmondson, D.E.; Mattevi, A. Crystal structures of monoamine oxidase B in complex with four inhibitors of the N-propargylaminoindan class. J. Med. Chem. 2004, 47, 1767–1774. [Google Scholar] [CrossRef] [PubMed]
  64. Borštnar, R.; Repič, M.; Kamerlin, S.C.L.; Vianello, R.; Mavri, J. Computational Study of the pKa Values of Potential Catalytic Residues in the Active Site of Monoamine Oxidase B. J. Chem. Theory Comput. 2012, 8, 3864–3870. [Google Scholar] [CrossRef] [Green Version]
  65. Ramsay, R.; Olivieri, A.; Holt, A. An improved approach to steady-state analysis of monoamine oxidases. J. Neural. Transm. 2011, 118, 1003–1019. [Google Scholar] [CrossRef] [PubMed]
  66. Holt, A. On the practical aspects of characterising monoamine oxidase inhibition in vitro. J. Neural. Transm. 2018, 125, 1685–1705. [Google Scholar] [CrossRef] [PubMed]
  67. Carreau, A.; El Hafny-Rahbi, B.; Matejuk, A.; Grillon, C.; Kieda, C. Why is the partial oxygen pressure of human tissues a crucial parameter? Small molecules and hypoxia. J. Cell. Mol. Med. 2011, 15, 1239–1253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Nandigama, R.K.; Edmondson, D.E. Structure-activity relations in the oxidation of phenethylamine analogues by recombinant human liver monoamine oxidase A. Biochemistry 2000, 39, 15258–15265. [Google Scholar] [CrossRef] [PubMed]
  69. Fitzpatrick, P.F. Insights into the mechanisms of flavoprotein oxidases from kinetic isotope effects. J. Label. Compd. Radiopharm. 2007, 50, 1016–1025. [Google Scholar] [CrossRef] [Green Version]
  70. Wang, J.; Edmondson, D.E. Topological Probes of Monoamine Oxidases A and B in Rat Liver Mitochondria: Inhibition by TEMPO-Substituted Pargyline Analogues and Inactivation by Proteolysis. Biochemistry 2011, 50, 2499–2505. [Google Scholar] [CrossRef] [Green Version]
  71. Prah, A.; Ogrin, P.; Mavri, J.; Stare, J. Nuclear quantum effects in enzymatic reactions: Simulation of the kinetic isotope effect of phenylethylamine oxidation catalyzed by monoamine oxidase A. Phys. Chem. Chem. Phys. 2020, 22, 6838–6847. [Google Scholar] [CrossRef]
  72. Sablin, S.O.; Ramsay, R.R. Substrates but not inhibitors alter the redox potentials of monoamine oxidases. Antioxid. Redox Signal. 2001, 3, 723–729. [Google Scholar] [CrossRef]
  73. Edmondson, D.E.; Newton-Vinson, P. The covalent FAD of monoamine oxidase: Structural and functional role and mechanism of the flavinylation reaction. Antioxid. Redox Signal. 2001, 3, 789–806. [Google Scholar] [CrossRef] [PubMed]
  74. Walsh, J.D.; Miller, A.F. Flavin reduction potential tuning by substitution and bending. J. Mol. Struct. 2003, 623, 185–195. [Google Scholar] [CrossRef]
  75. Hynson, R.; Kelly, S.; Price, N.; Ramsay, R. Conformational changes in monoamine oxidase A in response to ligand binding or reduction. Biochim. Biophys. Acta-Gen. Subj. 2004, 1672, 60–66. [Google Scholar] [CrossRef] [PubMed]
  76. Yu, P.H.; Bailey, B.A.; Durden, D.A.; Boulton, A.A. Stereospecific deuterium substitution at the alpha-carbon position of dopamine and its effect on oxidative deamination catalyzed by MAO-A and MAO-B from different tissues. Biochem. Pharm. 1986, 35, 1027–1036. [Google Scholar] [CrossRef]
  77. Yu, P.H.; Davis, B.A. Inversion of selectivity of N-substituted propargylamine monoamine oxidase inhibitors following structural modifications to quaternary salts. Int. J. Biochem. Cell Biol. 1999, 31, 1391–1397. [Google Scholar] [CrossRef]
  78. Edmondson, D.E.; Binda, C.; Wang, J.; Upadhyay, A.K.; Mattevi, A. Molecular and mechanistic properties of the membrane-bound mitochondrial monoamine oxidases. Biochemistry 2009, 48, 4220–4230. [Google Scholar] [CrossRef] [Green Version]
  79. Geha, R.M.; Rebrin, I.; Chen, K.; Shih, J.C. Substrate and inhibitor specificities for human monoamine oxidase A and B are influenced by a single amino acid. J. Biol. Chem. 2001, 276, 9877–9882. [Google Scholar] [CrossRef] [Green Version]
  80. Ma, J.C.; Yoshimura, M.; Yamashita, E.; Nakagawa, A.; Ito, A.; Tsukihara, T. Structure of rat monoamine oxidase A and its specific recognitions for substrates and inhibitors. J. Mol. Biol. 2004, 338, 103–114. [Google Scholar] [CrossRef]
  81. Milczek, E.M.; Binda, C.; Rovida, S.; Mattevi, A.; Edmondson, D.E. The ‘gating’ residues Ile199 and Tyr326 in human monoamine oxidase B function in substrate and inhibitor recognition. FEBS J. 2011, 278, 4860–4869. [Google Scholar] [CrossRef] [Green Version]
  82. Geha, R.M.; Chen, K.; Shih, J.C. Phe(208) and Ile(199) in human Monoamine Oxidase A and B do not determine substrate and inhibitor specificities as in rat. J. Neurochem. 2000, 75, 1304–1309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Hong, R.; Li, X. Discovery of monoamine oxidase inhibitors by medicinal chemistry approaches. MedChemComm 2019, 10, 10–25. [Google Scholar] [CrossRef] [PubMed]
  84. Rehuman, N.A.; Mathew, B.; Jat, R.K.; Nicolotti, O.; Kim, H. A Comprehensive Review of Monoamine Oxidase-A Inhibitors in their Syntheses and Poteneies. Comb. Chem. High Throughput Screen. 2020, 23, 898–914. [Google Scholar] [CrossRef]
  85. Wu, S.-M.; Qiu, X.-Y.; Liu, S.-J.; Sun, J. Single Heterocyclic Compounds as Monoamine Oxidase Inhibitors: From Past to Present. Mini-Rev. Med. Chem. 2020, 20, 908–920. [Google Scholar] [CrossRef]
  86. Becerra-Hernandez, A.; Galindo-de-la-Rosa, J.; Martinez-Pimentel, Y.; Ledesma-Garcia, J.; Alvarez-Contreras, L.; Guerra-Balcazar, M.; Aguilar-Elguezabal, A.; Alvarez, A.; Chavez-Ramirez, A.U.; Vallejo-Becerra, V. Novel biomaterial based on monoamine oxidase-A and multi-walled carbon nanotubes for serotonin detection. Biochem. Eng. J. 2019, 149, 107240. [Google Scholar] [CrossRef]
  87. Kacar, C.; Erden, P.E.; Dalkiran, B.; Inal, E.K.; Kilic, E. Amperometric biogenic amine biosensors based on Prussian blue, indium tin oxide nanoparticles and diamine oxidase- or monoamine oxidase-modified electrodes. Anal. Bioanal. Chem. 2020, 412, 1933–1946. [Google Scholar] [CrossRef]
  88. Ramsay, R.R.; Dunford, C.; Gillman, P.K. Methylene blue and serotonin toxicity: Inhibition of monoamine oxidase A (MAO A) confirms a theoretical prediction. Br. J. Pharmacol. 2007, 152, 946–951. [Google Scholar] [CrossRef]
  89. Delport, A.; Harvey, B.H.; Petzer, A.; Petzer, J.P. The monoamine oxidase inhibition properties of selected structural analogues of methylene blue. Toxicol. Appl. Pharmacol. 2017, 325, 1–8. [Google Scholar] [CrossRef]
  90. Hiraka, K.; Tsugawa, W.; Sode, K. Alteration of Electron Acceptor Preferences in the Oxidative Half-Reaction of Flavin-Dependent Oxidases and Dehydrogenases. Int. J. Mol. Sci. 2020, 21, 3797. [Google Scholar] [CrossRef] [PubMed]
  91. Iacovino, L.G.; Manzella, N.; Resta, J.; Vanoni, M.A.; Rotilio, L.; Pisani, L.; Edmondson, D.E.; Parini, A.; Mattevi, A.; Mialet-Perez, J.; et al. Rational Redesign of Monoamine Oxidase A into a Dehydrogenase to Probe ROS in Cardiac Aging. ACS Chem. Biol. 2020, 15, 1795–1800. [Google Scholar] [CrossRef]
  92. Zajkoska, P.; Rosenberg, M.; Heath, R.; Malone, K.J.; Stloukal, R.; Turner, N.J.; Rebros, M. Immobilised whole-cell recombinant monoamine oxidase biocatalysis. Appl. Microbiol. Biotechnol. 2015, 99, 1229–1236. [Google Scholar] [CrossRef] [PubMed]
  93. Batista, V.F.; Galman, J.L.; Pinto, D.C.G.A.; Silva, A.M.S.; Turner, N.J. Monoamine Oxidase: Tunable Activity for Amine Resolution and Functionalization. ACS Catal. 2018, 8, 11889–11907. [Google Scholar] [CrossRef]
  94. Markosova, K.; Camattari, A.; Rosenberg, M.; Glieder, A.; Turner, N.J.; Rebros, M. Cloning and upscale production of monoamine oxidase N (MAO-N D5) by Pichia pastoris. Biotechnol. Lett. 2018, 40, 127–133. [Google Scholar] [CrossRef] [Green Version]
  95. Toscani, A.; Risi, C.; Black, G.W.; Brown, N.L.; Shaaban, A.; Turner, N.J.; Castagnolo, D. Monoamine Oxidase (MAO-N) Whole Cell Biocatalyzed Aromatization of 1,2,5,6-Tetrahydropyridines into Pyridines. ACS Catal. 2018, 8, 8781–8787. [Google Scholar] [CrossRef]
  96. Lauder, K.; Masci, D.; Toscani, A.; Al Mekdad, A.; Black, G.W.; Brown, N.L.; Turner, N.J.; Luisi, R.; Castagnolo, D. A facile and regioselective multicomponent synthesis of chiral aryl-1,2-mercaptoamines in water followed by monoamine oxidase (MAO-N) enzymatic resolution. Org. Biomol. Chem. 2019, 17, 8982–8986. [Google Scholar] [CrossRef] [PubMed]
  97. Knez, D.; Colettis, N.; Iacovino, L.G.; Sova, M.; Pišlar, A.; Konc, J.; Lešnik, S.; Higgs, J.; Kamecki, F.; Mangialavori, I.; et al. Stereoselective Activity of 1-Propargyl-4-styrylpiperidine-like Analogues That Can Discriminate between Monoamine Oxidase Isoforms A and B. J. Med. Chem. 2020, 63, 1361–1387. [Google Scholar] [CrossRef] [PubMed]
  98. Carradori, S.; Secci, D.; Petzer, J.P. MAO inhibitors and their wider applications: A patent review. Expert Opin. Ther. Pat. 2018, 28, 211–226. [Google Scholar] [CrossRef] [PubMed]
  99. Ramsay, R.R.; Bradley, R.J.E.; Giurato, L.; Guccione, S. The shape of flavin in monoamine oxidase. In Flavins and Flavoproteins 2008; Frago, S., Gomez-Moreno, C., Medina, M., Eds.; Prensas Universitarias de Zaragoza: Zaragoza, Spain, 2008; pp. 183–186. [Google Scholar]
  100. Eisenreich, W.; Kemter, K.; Bacher, A.; Mulrooney, S.B.; Williams, C.H., Jr.; Müller, F. 13C-, 15N- and 31P-NMR studies of oxidized and reduced low molecular mass thioredoxin reductase and some mutant proteins. Eur. J. Biochem. 2004, 271, 1437–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Rojas, R.J.; Edmondson, D.E.; Almos, T.; Scott, R.; Massari, M.E. Reversible and irreversible small molecule inhibitors of monoamine oxidase B (MAO-B) investigated by biophysical techniques. Bioorganic Med. Chem. 2015, 23, 770–778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Albreht, A.; Vovk, I.; Mavri, J.; Marco-Contelles, J.; Ramsay, R.R. Evidence for a Cyanine Link Between Propargylamine Drugs and Monoamine Oxidase Clarifies the Inactivation Mechanism. Front. Chem. 2018, 6, 169. [Google Scholar] [CrossRef] [Green Version]
  103. Upadhyay, A.K.; Borbat, P.P.; Wang, J.; Freed, J.H.; Edmondson, D.E. Determination of the oligomeric states of human and rat monoamine oxidases in the outer mitochondrial membrane and octyl beta-D-glucopyranoside micelles using pulsed dipolar electron spin resonance spectroscopy. Biochemistry 2008, 47, 1554–1566. [Google Scholar] [CrossRef] [PubMed]
  104. Esteban, G.; Allan, J.; Samadi, A.; Mattevi, A.; Unzeta, M.; Marco-Contelles, J.; Binda, C.; Ramsay, R.R. Kinetic and structural analysis of the irreversible inhibition of human monoamine oxidases by ASS234, a multi-target compound designed for use in Alzheimer’s disease. Biochim. Biophys. Acta-Proteins Proteom. 2014, 1844, 1104–1110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Upadhyay, A.K.; Wang, J.; Edmondson, D.E. Comparison of the structural properties of the active site cavities of human and rat monoamine oxidase A and B in their soluble and membrane-bound forms. Biochemistry 2008, 47, 526–536. [Google Scholar] [CrossRef] [PubMed]
  106. Son, S.Y.; Ma, A.; Kondou, Y.; Yoshimura, M.; Yamashita, E.; Tsukihara, T. Structure of human Monoamine Oxidase A at 2.2-angstrom resolution: The control of opening the entry for substrates/inhibitors. Proc. Natl. Acad. Sci. USA 2008, 105, 5739–5744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Apostolov, R.; Yonezawa, Y.; Standley, D.M.; Kikugawa, G.; Takano, Y.; Nakamura, H. Membrane Attachment Facilitates Ligand Access to the Active Site in Monoamine Oxidase A. Biochemistry 2009, 48, 5864–5873. [Google Scholar] [CrossRef]
  108. Allen, W.J.; Bevan, D.R. Steered Molecular Dynamics Simulations Reveal Important Mechanisms in Reversible Monoamine Oxidase B Inhibition. Biochemistry 2011, 50, 6441–6454. [Google Scholar] [CrossRef]
  109. Jones, H.B.L.; Crean, R.M.; Mullen, A.; Kendrick, E.G.; Bull, S.D.; Wells, S.A.; Carbery, D.R.; MacMillan, F.; van der Kamp, M.W.; Pudney, C.R. Exposing the Interplay Between Enzyme Turnover, Protein Dynamics, and the Membrane Environment in Monoamine Oxidase B. Biochemistry 2019, 58, 2362–2372. [Google Scholar] [CrossRef]
  110. Burlakova, E.B.; Kayrane, C.B.; Molochkina, E.M.; Khokhlova, A.P. Modifications of external mitochondrial-membrane lipids in mice liver and kinetic patterns of membrane-bound monoamine-oxidase in vivo and in vitro. Vopr. Meditsinskoi Khimii 1984, 30, 66–72. [Google Scholar]
  111. Ekstedt, B.; Oreland, L. Effect of lipid-depletion on different forms of monoamine-oxidase in rat-liver mitochondria. Biochem. Pharmacol. 1976, 25, 119–124. [Google Scholar] [CrossRef]
  112. Houslay, M.D. Lipid substitution of mitochondrial monoamine-oxidase can lead to the abolition of clorgyline selective-inhibition without alteration in the A-B ratio assessed by substrate utilization. Biochem. Pharmacol. 1980, 29, 3211–3213. [Google Scholar] [CrossRef]
  113. Medvedev, A.; Kirkel, A.; Kamyshanskaya, N.; Gorkin, V. Lipid-peroxidation affects catalytic properties of rat-liver mitochondrial Monoamine Oxidases and their sensitivity to proteolysis. Int. J. Biochem. 1993, 25, 1791–1799. [Google Scholar] [CrossRef]
  114. Huang, R.H. Lipid-protein interactions in the multiple forms of monoamine-oxidase—Enzymatic and electron-spin-resonance studies with purified intact rat-brain mitochondria. Mol. Pharmacol. 1980, 17, 192–198. [Google Scholar]
  115. Kearney, E.B.; Salach, J.I.; Walker, W.H.; Seng, R.; Singer, T.P. Structure of the covalently bound flavin of monoamine oxidase. Biochem. Biophys. Res. Commun. 1971, 42, 490–496. [Google Scholar] [CrossRef]
  116. Hubalek, F.; Pohl, J.; Edmondson, D.E. Structural comparison of human monoamine oxidases A and B—Mass spectrometry monitoring of cysteine reactivities. J. Biol. Chem. 2003, 278, 28612–28618. [Google Scholar] [CrossRef] [Green Version]
  117. Paulsen, C.E.; Carroll, K.S. Cysteine-Mediated Redox Signaling: Chemistry, Biology, and Tools for Discovery. Chem. Rev. 2013, 113, 4633–4679. [Google Scholar] [CrossRef] [PubMed]
  118. Nietzel, T.; Mostertz, J.; Hochgraefe, F.; Schwarzlaender, M. Redox regulation of mitochondrial proteins and proteomes by cysteine thiol switches. Mitochondrion 2017, 33, 72–83. [Google Scholar] [CrossRef] [PubMed]
  119. Go, Y.-M.; Chandler, J.D.; Jones, D.P. The cysteine proteome. Free Radic. Biol. Med. 2015, 84, 227–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Weyler, W.; Hsu, Y.P.P.; Breakefield, X.O. Biochemistry and genetics of monoamine-oxidase. Pharmacol. Ther. 1990, 47, 391–417. [Google Scholar] [CrossRef]
  121. Hubalek, F.; Edmondson, D.E. Thiol reactivities as probes of MAO A and MAO B structure and function. In Flavins and Flavoproteins 2002; Chapman, S.K., Perham, R.N., Scrutton, N.S., Eds.; Rudolf Weber: Berlin, Germany, 2002; pp. 217–222. [Google Scholar]
  122. Vintem, A.; Price, N.; Silverman, R.; Ramsay, R. Mutation of surface cysteine 374 to alanine in monoamine oxidase A alters substrate turnover and inactivation by cyclopropylamines. Bioorganic Med. Chem. 2005, 13, 3487–3495. [Google Scholar] [CrossRef] [PubMed]
  123. Wu, H.F.; Chen, K.; Shih, J.C. Site-directed mutagenesis of Monoamine Oxidase-A and Oxidase-B—Role of cysteines. Mol. Pharmacol. 1993, 43, 888–893. [Google Scholar]
  124. Zhong, B.Y.; Silverman, R.B. Identification of the active site cysteine in bovine liver monoamine oxidase B. J. Am. Chem. Soc. 1997, 119, 6690–6691. [Google Scholar] [CrossRef]
  125. Murphy, M.P. Mitochondrial Thiols in Antioxidant Protection and Redox Signaling: Distinct Roles for Glutathionylation and Other Thiol Modifications. Antioxid. Redox Signal. 2012, 16, 476–495. [Google Scholar] [CrossRef] [PubMed]
  126. Alves-Figueiredo, H.; Silva-Platas, C.; Lozano, O.; Vazquez-Garza, E.; Guerrero-Beltran, C.E.; Zarain-Herzberg, A.; Garcia-Rivas, G. A systematic review of post-translational modifications in the mitochondrial permeability transition pore complex associated with cardiac diseases. Biochim. Biophys. Acta-Mol. Basis Dis. 2021, 1867, 165992. [Google Scholar] [CrossRef] [PubMed]
  127. Cao, X.; Li, X.-M.; Mousseau, D.D. Calcium alters monoamine oxidase-A parameters in human cerebellar and rat glial C6 cell extracts: Possible influence by distinct signalling pathways. Life Sci. 2009, 85, 262–268. [Google Scholar] [CrossRef] [PubMed]
  128. Egashira, T.; Sakai, K.; Sakurai, M.; Takayama, F. Calcium disodium edetate enhances type A monoamine oxidase activity in monkey brain. Biol. Trace Elem. Res. 2003, 94, 203–211. [Google Scholar] [CrossRef]
  129. Kosenko, E.A.; Venediktova, N.I.; Kaminsky, Y.G. Calcium and ammonia stimulate Monoamine Oxidase A activity in brain mitochondria. Biol. Bull. 2003, 30, 449–452. [Google Scholar] [CrossRef]
  130. Samantaray, S.; Chandra, G.; Mohanakumar, K.P. Calcium channel agonist, (+/−)-Bay K8644, causes a transient increase in striatal monoamine oxidase activity in Balb/c mice. Neurosci. Lett. 2003, 342, 73–76. [Google Scholar] [CrossRef]
  131. Ramsay, R.R.; Armstrong, S. Calcium ions alter monoamine oxidase A activity. In Flavins and Flavoproteins 2011; Miller, S.H.R., Ed.; Lulu Press: Raleigh, NC, USA, 2011. [Google Scholar]
  132. Cao, X.; Wei, Z.; Gabriel, G.G.; Li, X.; Mousseau, D.D. Calcium-sensitive regulation of monoamine oxidase-A contributes to the production of peroxyradicals in hippocampal cultures: Implications for Alzheimer disease-related pathology. BMC Neurosci. 2007, 8, 73. [Google Scholar] [CrossRef] [Green Version]
  133. Pennington, P.R.; Wei, Z.; Rui, L.; Doig, J.A.; Graham, B.; Kuski, K.; Gabriel, G.G.; Mousseau, D.D. Alzheimer disease-related presenilin-1 variants exert distinct effects on monoamine oxidase-A activity in vitro. J. Neural. Transm. 2011, 118, 987–995. [Google Scholar] [CrossRef]
  134. Wei, Z.; Gabriel, G.G.; Rui, L.; Cao, X.; Pennington, P.R.; Chlan-Fourney, J.; Nazarali, A.J.; Baker, G.B.; Mousseau, D.D. Monoamine Oxidase-A Physically Interacts with Presenilin-1(M146V) in the Mouse Cortex. J. Alzheimers Dis. 2012, 28, 403–422. [Google Scholar] [CrossRef] [Green Version]
  135. Shih, J.C.; Chen, K. MAO-A and -B gene knock-out mice exhibit distinctly different behavior. Neurobiology 1999, 7, 235–246. [Google Scholar] [PubMed]
  136. Godar, S.C.; Fite, P.J.; McFarlin, K.M.; Bortolato, M. The role of monoamine oxidase A in aggression: Current translational developments and future challenges. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2016, 69, 90–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Naoi, M.; Maruyama, W.; Shamoto-Nagai, M. Type A and B monoamine oxidases distinctly modulate signal transduction pathway and gene expression to regulate brain function and survival of neurons. J. Neural Transm. 2018, 125, 1635–1650. [Google Scholar] [CrossRef]
  138. Cheng, Y.; Buchan, M.; Vitanova, K.; Aitken, L.; Gunn-Moore, F.J.; Ramsay, R.R.; Doherty, G. Neuroprotective actions of leptin facilitated through balancing mitochondrial morphology and improving mitochondrial function. J. Neurochem. 2020, 155, 191–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Quartey, M.O.; Nyarko, J.N.K.; Pennington, P.R.; Heistad, R.M.; Klassen, P.C.; Baker, G.B.; Mousseau, D.D. Alzheimer Disease and Selected Risk Factors Disrupt a Co-regulation of Monoamine Oxidase-A/B in the Hippocampus, but Not in the Cortex. Front. Neurosci. 2018, 12, 419. [Google Scholar] [CrossRef]
  140. Fitzgerald, J.C.; Ufer, C.; Billett, E.E. A link between monoamine oxidase-A and apoptosis in serum deprived human SH-SY5Y neuroblastoma cells. J. Neural Transm. 2007, 114, 807–810. [Google Scholar] [CrossRef] [PubMed]
  141. Rybaczyk, L.A.; Bashaw, M.J.; Pathak, D.R.; Huang, K. An indicator of cancer: Downregulation of Monoamine Oxidase-A in multiple organs and species. BMC Genom. 2008, 9, 134. [Google Scholar] [CrossRef] [Green Version]
  142. Li, P.C.; Siddiqi, I.N.; Mottok, A.; Loo, E.Y.; Wu, C.H.; Cozen, W.; Steidl, C.; Shih, J.C. Monoamine oxidase A is highly expressed in classical Hodgkin lymphoma. J. Pathol. 2017, 243, 220–229. [Google Scholar] [CrossRef]
  143. Wu, J.B.; Shao, C.; Li, X.; Li, Q.; Hu, P.; Shi, C.; Li, Y.; Chen, Y.T.; Yin, F.; Liao, C.P.; et al. Monoamine oxidase A mediates prostate tumorigenesis and cancer metastasis. J. Clin. Investig. 2014, 124, 2891–2908. [Google Scholar] [CrossRef] [Green Version]
  144. Gross, M.E.; Agus, D.B.; Dorff, T.B.; Pinski, J.K.; Quinn, D.I.; Castellanos, O.; Gilmore, P.; Shih, J.C. Phase 2 trial of monoamine oxidase inhibitor phenelzine in biochemical recurrent prostate cancer. Prostate Cancer Prostatic Dis. 2021, 24, 61–68. [Google Scholar] [CrossRef]
  145. Krueger, M.J.; Mazouz, F.; Ramsay, R.R.; Milcent, R.; Singer, T.P. Dramatic species differences in the susceptibility of Monoamine-Oxidase-B to a group of powerful inhibitors. Biochem. Biophys. Res. Commun. 1995, 206, 556–562. [Google Scholar] [CrossRef]
  146. Egashira, T.; Takayama, F.; Yamanaka, Y. The inhibition of monoamine oxidase activity by various antidepressants: Differences found in various mammalian species. Jpn. J. Pharmacol. 1999, 81, 115–121. [Google Scholar] [CrossRef]
  147. Inoue, H.; Castagnoli, K.; van der Schyf, C.; Mabic, S.; Igarashi, K.; Castagnoli, N. Species-dependent differences in monoamine oxidase A and B-catalyzed oxidation of various C4 substituted 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridinyl derivatives. J. Pharmacol. Exp. Ther. 1999, 291, 856–864. [Google Scholar] [PubMed]
  148. Ben Ramadan, Z.; Wrang, M.L.; Tipton, K.F. Species differences in the selective inhibition of monoamine oxidase (1-methyl-2-phenylethyl)hydrazine and its potentiation by cyanide. Neurochem. Res. 2007, 32, 1783–1790. [Google Scholar] [CrossRef] [PubMed]
  149. Yagodina, O.V.; Basova, I.N. Comparative enzymological study of catalytic properties of liver monoamine oxidases in frogs. J. Evol. Biochem. Physiol. 2010, 46, 350–356. [Google Scholar] [CrossRef]
  150. Senatori, O.; Pierucci, F.; Parvez, S.H.; Scopelliti, R.; Nicotra, A. Monoamine oxidase in Teleosts. Biog. Amines 2003, 17, 199–213. [Google Scholar] [CrossRef]
  151. Wang, J.; Edmondson, D.E. H-2 Kinetic Isotope Effects and pH Dependence of Catalysis as Mechanistic Probes of Rat Monoamine Oxidase A: Comparisons with the Human Enzyme. Biochemistry 2011, 50, 7710–7717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Santillo, M.F.; Liu, Y.T.; Ferguson, M.; Vohra, S.N.; Wiesenfeld, P.L. Inhibition of monoamine oxidase (MAO) by beta-carbolines and their interactions in live neuronal (PC12) and liver (HuH-7 and MH1C1) cells. Toxicol. In Vitro 2014, 28, 403–410. [Google Scholar] [CrossRef]
  153. Brummett, B.H.; Krystal, A.D.; Siegler, I.C.; Kuhn, C.; Surwit, R.S.; Zuechner, S.; Ashley-Koch, A.; Barefoot, J.C.; Williams, R.B. Associations of a regulatory polymorphism of monoamine oxidase-A gene promoter (MAOA-uVNTR) with symptoms of depression and sleep quality. Psychosom. Med. 2007, 69, 396–401. [Google Scholar] [CrossRef] [Green Version]
  154. Kunugi, H.; Ishida, S.; Kato, T.; Tatsumi, M.; Sakai, T.; Hattori, M.; Hirose, T.; Nanko, S. A functional polymorphism in the promoter region of monoamine oxidase-A gene and mood disorders. Mol. Psychiatry 1999, 4, 393–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Manca, M.; Pessoa, V.; Lopez, A.I.; Harrison, P.T.; Miyajima, F.; Sharp, H.; Pickles, A.; Hill, J.; Murgatroyd, C.; Bubb, V.J.; et al. The Regulation of Monoamine Oxidase A Gene Expression by Distinct Variable Number Tandem Repeats. J. Mol. Neurosci. 2018, 64, 459–470. [Google Scholar] [CrossRef] [Green Version]
  156. Roohi, J.; DeVincent, C.; Hatchwell, E.; Gadow, K. Association of a Monoamine Oxidase-A Gene Promoter Polymorphism with ADHD and Anxiety in Boys with Autism Spectrum Disorder. J. Autism Dev. Disord. 2009, 39, 67–74. [Google Scholar] [CrossRef] [Green Version]
  157. Bortolato, M.; Floris, G.; Shih, J.C. From aggression to autism: New perspectives on the behavioral sequelae of monoamine oxidase deficiency. J. Neural Transm. 2018, 125, 1589–1599. [Google Scholar] [CrossRef]
  158. Pregeljc, D.; Jug, U.; Mavri, J.; Stare, J. Why does the Y326I mutant of monoamine oxidase B decompose an endogenous amphetamine at a slower rate than the wild type enzyme? Reaction step elucidated by multiscale molecular simulations. Phys. Chem. Chem. Phys. 2018, 20, 4181–4188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Fowler, J.S.; Logan, J.; Shumay, E.; Alia-Klein, N.; Wang, G.-J.; Volkow, N.D. Monoamine oxidase: Radiotracer chemistry and human studies. J. Label. Compd. Radiopharm. 2015, 58, 51–64. [Google Scholar] [CrossRef] [PubMed]
  160. Fowler, J.S.; Logan, J.; Wang, G.-J.; Volkow, N.D.; Telang, F.; Ding, Y.-S.; Shea, C.; Garza, V.; Xu, Y.; Li, Z.; et al. Comparison of the binding of the irreversible monoamine oxidase tracers, [11C]clorgyline and [11C]l-deprenyl in brain and peripheral organs in humans. Nucl. Med. Biol. 2004, 31, 313–319. [Google Scholar] [CrossRef] [PubMed]
  161. Muller, T.; Mohr, J.D. Pharmacokinetics of monoamine oxidase B inhibitors in Parkinson’s disease: Current status. Expert Opin. Drug Metab. Toxicol. 2019, 15, 429–435. [Google Scholar] [CrossRef] [PubMed]
  162. Arnett, C.D.; Fowler, J.S.; MacGregor, R.R.; Schlyer, D.J.; Wolf, A.P.; Långström, B.; Halldin, C. Turnover of Brain Monoamine Oxidase Measured In Vivo by Positron Emission Tomography Using l-[11C]Deprenyl. J. Neurochem. 1987, 49, 527. [Google Scholar] [CrossRef] [PubMed]
  163. Naoi, M.; Maruyama, W.; Shamoto-Nagai, M. Rasagiline and selegiline modulate mitochondrial homeostasis, intervene apoptosis system and mitigate alpha-synuclein cytotoxicity in disease-modifying therapy for Parkinson’s disease. J. Neural Transm. 2020, 127, 131–147. [Google Scholar] [CrossRef]
  164. Müller, T. Pharmacokinetic drug evaluation of safinamide mesylate for the treatment of mid-to-late stage Parkinson’s disease. Expert Opin. Drug Metab. Toxicol. 2017, 13, 693–699. [Google Scholar] [CrossRef]
  165. Orayj, K.; Lane, E. Patterns and Determinants of Prescribing for Parkinson’s Disease: A Systematic Literature Review. Parkinsons Dis. 2019, 2019, 9237181. [Google Scholar] [CrossRef] [Green Version]
  166. Binde, C.D.; Tvete, I.F.; Gasemyr, J.; Natvig, B.; Klemp, M. A multiple treatment comparison meta-analysis of monoamine oxidase type B inhibitors for Parkinson’s disease. Br. J. Clin. Pharmacol. 2018, 84, 1917–1927. [Google Scholar] [CrossRef]
  167. Gillman, P.K. A reassessment of the safety profile of monoamine oxidase inhibitors: Elucidating tired old tyramine myths. J. Neural. Transm. 2018, 125, 1707–1717. [Google Scholar] [CrossRef]
  168. Gillman, P.K.; Feinberg, S.S.; Fochtmann, L.J. Revitalizing monoamine oxidase inhibitors: A call for action. CNS Spectr. 2020, 25, 452–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Mathew, B.; Carradori, S.; Guglielmi, P.; Uddin, M.S.; Kim, H. New Aspects of Monoamine Oxidase B Inhibitors: The Key Role of Halogens to Open the Golden Door. Curr. Med. Chem. 2021, 28, 266–283. [Google Scholar] [CrossRef] [PubMed]
  170. Mellado, M.; Mella, J.; Gonzalez, C.; Vina, D.; Uriarte, E.; Matos, M.J. 3-Arylcoumarins as highly potent and selective monoamine oxidase B inhibitors: Which chemical features matter? Bioorganic Chem. 2020, 101, 103964. [Google Scholar] [CrossRef] [PubMed]
  171. Stefanachi, A.; Leonetti, F.; Pisani, L.; Catto, M.; Carotti, A. Coumarin: A Natural, Privileged and Versatile Scaffold for Bioactive Compounds. Molecules 2018, 23, 250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Mladenović, M.; Patsilinakos, A.; Pirolli, A.; Sabatino, M.; Ragno, R. Understanding the Molecular Determinant of Reversible Human Monoamine Oxidase B Inhibitors Containing 2H-Chromen-2-One Core: Structure-Based and Ligand-Based Derived Three-Dimensional Quantitative Structure–Activity Relationships Predictive Models. J. Chem. Inf. Modeling 2017, 57, 787–814. [Google Scholar]
  173. Lutsenko, K.; Hagenow, S.; Affini, A.; Reiner, D.; Stark, H. Rasagiline derivatives combined with histamine H-3 receptor properties. Bioorganic Med. Chem. Lett. 2019, 29, 126612. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical mechanisms proposed for MAO-catalyzed oxidation of amines.
Figure 1. Chemical mechanisms proposed for MAO-catalyzed oxidation of amines.
Chemistry 03 00069 g001
Figure 2. Small molecule binding in the active site of MAO A and MAO B. Representative binding modes for the inhibitor 2-amino-5-(4-chlorophenyl)furan-3,4-dicarbonitrile bound to MAO A and MAO B obtained from molecular dynamics simulations [42]. The inhibitor is shown as orange sticks in MAO A and yellow sticks MAO B. Residues common to MAO A and MAO B are shown as blue sticks, with the key “aromatic cage” tyrosines labelled in black (TYR). Residues that differ are shown in green for MAO A and in magenta for MAO B and identified by the MAO A residue number followed by the MAO B residue number. These residues have been reported to be important in selectivity for small molecule binding. The flavin cofactor with the bent isoalloxazine ring is shown as blue sticks for both MAO A and B.
Figure 2. Small molecule binding in the active site of MAO A and MAO B. Representative binding modes for the inhibitor 2-amino-5-(4-chlorophenyl)furan-3,4-dicarbonitrile bound to MAO A and MAO B obtained from molecular dynamics simulations [42]. The inhibitor is shown as orange sticks in MAO A and yellow sticks MAO B. Residues common to MAO A and MAO B are shown as blue sticks, with the key “aromatic cage” tyrosines labelled in black (TYR). Residues that differ are shown in green for MAO A and in magenta for MAO B and identified by the MAO A residue number followed by the MAO B residue number. These residues have been reported to be important in selectivity for small molecule binding. The flavin cofactor with the bent isoalloxazine ring is shown as blue sticks for both MAO A and B.
Chemistry 03 00069 g002
Figure 3. Alternative pathways for oxidation complicate MAO kinetic studies.
Figure 3. Alternative pathways for oxidation complicate MAO kinetic studies.
Chemistry 03 00069 g003
Table 1. Some substrates, inhibitors and inactivators of MAO with their kinetic parameters.
Table 1. Some substrates, inhibitors and inactivators of MAO with their kinetic parameters.
MAO ACompoundMAO B
KM
(μM)
Vmax
(pmol.min−1.mg protein−1)
Substrates A KM
(μM)
Vmax
(pmol.min−1.mg protein−1)
137228Serotonin Chemistry 03 00069 i00110936.6
212680Dopamine Chemistry 03 00069 i002229702
140202-Phenylethylamine Chemistry 03 00069 i0034309
127182Tyramine Chemistry 03 00069 i004107343
Ki
(μM)
Reversible
Inhibitors B
Ki
(μM)
18 D-Amphetamine Chemistry 03 00069 i005250
3 MPP+ c Chemistry 03 00069 i006230
365 Safinamide Chemistry 03 00069 i0070.45
11.5 Moclobemide Chemistry 03 00069 i008>100
0.005 Harmine Chemistry 03 00069 i009121
0.028 Methylene Blue Chemistry 03 00069 i0105.5
KI
(μM)
kinact
(min−1)
Irreversible
Inhibitors C
KI
(μM)
kinact
(min−1)
3.10.12Phenelzine Chemistry 03 00069 i01150 a0.9 a
7.70.78Tranylcypromine Chemistry 03 00069 i0120.70.16
1930.25Selegiline d Chemistry 03 00069 i0130.10.53
9.70.007Rasagiline b Chemistry 03 00069 i0140.70.05
0.031.8Clorgyline Chemistry 03 00069 i0151.80.02
3.00.39Contilisant e Chemistry 03 00069 i0164.60.19
Data: A from [13]; B summarized in [14] from references therein; C from [15] except a [16] and b [17]. c MPP+ is 1-methyl-4-phenyl-pyridinium; d Selegiline is also known as l-deprenyl; e Contilisant is a multi-target compound which also inhibits cholinesterase and binds to the histamine H3 receptor [18].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramsay, R.R.; Albreht, A. Questions in the Chemical Enzymology of MAO. Chemistry 2021, 3, 959-978. https://doi.org/10.3390/chemistry3030069

AMA Style

Ramsay RR, Albreht A. Questions in the Chemical Enzymology of MAO. Chemistry. 2021; 3(3):959-978. https://doi.org/10.3390/chemistry3030069

Chicago/Turabian Style

Ramsay, Rona R., and Alen Albreht. 2021. "Questions in the Chemical Enzymology of MAO" Chemistry 3, no. 3: 959-978. https://doi.org/10.3390/chemistry3030069

Article Metrics

Back to TopTop