Introduction

In our previous reports, we analyzed the properties and participation of human enzymes in the metabolism of physiological and xenobiotic compounds, including natural products (Rendić and Guengerich 2012, 2015, 2021). The analysis showed an overwhelming participation of the cytochrome P450 (P450, CYP) enzymes (~ 95%) in the metabolism of the compounds. P450 enzymes catalyze a great number of metabolic reactions and have important effects on the biological activities (physiologic, therapeutic, and/or toxic) of xenobiotics such as drugs, natural products, “general chemicals” (e.g., pesticides, pro-carcinogens, various environmental chemicals), and physiological compounds. In addition to P450s, other enzymes such as microsomal flavin-containing monooxygenase (FMO), monoamine oxidase (MAO), and aldehyde oxidase (AOX) enzymes participate in the metabolism of these compounds, although to a lower extent (~ 2%, 1%, and 2%, respectively). Other oxidoreductase enzymes participate to an extent of < 1% (Rendić and Guengerich 2012, 2015). The mechanism, kinetics, and metabolic properties of P450 enzymes (Guengerich 2022) and oxidative metabolism, and the gene regulation of non-cytochrome P450 enzymes have been discussed recently (Pang et al. 2022). In the present paper, we discuss mechanisms and metabolic properties of human FMO, MAO, NAD(P)H quinone oxidoreductase (NQO), molybdenum-containing hydroxylases (AOX and xanthine oxidoreductase (XOR) enzymes) in the oxidation of drugs, physiological and natural products, and other (general) chemicals as substrates and inhibitors of these enzymes, in the context of their participation in the metabolism of the compounds and also possible toxic effects that might result from oxidation and reduction reactions.

The review is divided into four parts, addressing these four sets of enzymes. Only the human enzymes are discussed and data are only presented for these. The experimental data are presented in tables, and the published kinetic values were categorized according to the values and effects presented in Table 1.

Table 1 Values and limits used for evaluation of kinetic data*

Results and discussion

Flavin-containing monooxygenase (FMO)

As reported previously, FMOs participate in ~ 2% of reactions involved in the metabolism of xenobiotics, natural products, and physiological compounds (Rendić and Guengerich 2015). There is a higher fraction of FMO enzymes involved in the metabolism of general chemicals when compared to the metabolism of drugs, natural products, or physiological compounds. The reactions catalyzed by FMO enzymes are predominately detoxication reactions and include N-, S-, P-, and Se-atom oxygenations, depending on the substrate structure (Table 2).

Table 2 Examples of substrates, products, and reactions catalyzed by FMO enzymes

In some cases, FMOs are involved in the activation of substrates to toxic products. When calculating the participation of FMO enzymes in activation reactions, we found that FMOs participate in ~ 1% in the reactions, catalyzed predominately by FMO1 and FMO3 and related to the formation of N- or S-oxides. These results show equal participation of FMO enzymes in detoxication reactions and the formation of potentially toxic products. For comparison, P450 enzymes participate in ~ 66% of reactions involving the formation of toxic products and in 95% of the overall oxidations and reductions of xenobiotics and natural products (Rendić and Guengerich 2012, 2015).

Enzymes

In the literature, different terminology has been used for these enzymes: FMO(s), FAD-containing amine oxidases, microsomal oxygenases containing flavin, and mixed-function microsomal amine oxidases. The enzyme was discovered by the late Prof. Daniel Ziegler, who worked with the enzyme from swine, and frequently scientists simply referred to this as “Ziegler’s Enzyme” for many years (Pettit et al. 1964; Ziegler 1988, 2002; Ziegler and Pettit 1966). The enzymes are found in the endoplasmic reticulum of most organs and tissues, predominately in the liver and in the lungs, kidneys, digestive tract, brain, and others (Dannan and Guengerich 1982).

The human FMO enzymes are characterized by the following features: FMO enzymes contain 1 mol FAD/mol enzyme, Mr ~ 65 kDa, and about 535 amino acids. In humans, 11 FMO genes have been identified, encoding five active FMOs (FMO1–5) and six pseudogenes. FMOs are differentially distributed in organs, and the amino acid sequences of the orthologous forms of the enzymes in different animal species are 80–90% similar (Cashman 2004; Henderson et al. 2014; Hines 2006; Huang et al. 2021; Koukouritaki et al. 2002; Krueger et al. 2009; Nagashima et al. 2009; Phillips and Shephard 2017, 2020; Shimizu et al. 2011, 2015; Ziegler 1988).

The most frequently represented reactions catalyzed by FMO enzymes are N- and S-oxygenations (Elfarra 1995; Furnes and Schlenk 2004; Krause et al. 2003), although some oxygenations are known for phosphorus and selenium atoms (Hodgson and Levi 1992; Jones et al. 2017; Rooseboom et al. 2001) (Table 2).

FMO1 is the major form expressed in the neonatal liver and kidneys and small intestine of adults. FMO2 is the most abundant in human lungs and is expressed in the liver and kidneys at a minor level. The non-functional variant FMO2*2 is predominant in humans, but in some ethnic groups that have been studied (Afro- and Hispanic-Americans) the variant FMO2*1 is present (Krueger et al. 2005). The developmental expression pattern for human hepatic FMO1 and FMO3 shows that relatively high levels of FMO1 expression are observed throughout prenatal development, in particular during the embryonic period, but FMO3 is essentially absent in the fetal liver. In the human liver, FMO3 is the most abundant enzyme and predominantly oxidizes tertiary amines, including a large number of clinically important drugs and amines ingested in food. FMO3 is a highly polymorphic enzyme, and polymorphism is related to a rare hereditary disorder of the inability to metabolize trimethylamine (a disorder called trimethylaminuria) (Phillips et al. 1995). FMO3 enzymes have been associated with some clinically relevant drug–drug or drug–chemical interactions because a large number of clinically important drugs (as well as natural products, e.g., indoles, tyramine, trimethylamine) possess amine structures. FMO4 is present at a low level in multiple tissues (e.g., liver, kidneys, brain). FMO5 is highly expressed in the adult human liver.

There are significant differences between individuals and ethnic groups in both expression and functional activity. Genetic polymorphism in the human FMO genes (in major part associated with the FMO3 gene) may lead to changes in N- and/or S-oxygenations of drugs, xenobiotics, and endogenous substances.

Following the P450s (Rendić 2002; Rendić and Di Carlo 1997), FMOs are the most important enzymes involved in the monooxygenation of amine-containing xenobiotics or amines that are formed during the biotransformation of drugs, general chemicals, natural products, and physiological compounds (Rendić and Guengerich 2015). Reactions catalyzed by FMO enzymes have been generally considered as detoxications but there are exceptions to this rule. All FMO enzymes possess the structural features by which FAD and NADPH are bound. Important endogenous roles for the FMO family have been suggested, including the regulation of cellular stress resistance and major cellular metabolic activities that involve central carbon metabolism (Huang et al. 2021; Krueger and Williams 2005).

Typical substrates include aliphatic, basic amines and some aromatic primary amines, secondary amines, tertiary amines, N-arylamides, heteroaromatic amines, hydroxylamines, and hydramines (e.g., metamizole, N,N-dimethylaniline). Substrates of FMOs (e.g., N-alkyl arylamines including N-methylaniline and N,N-dimethylaniline) can be substrates for both FMO and P450 enzymes, depending on the structural and electronic properties of substituents and basicity of the amines. FMO enzymes predominantly catalyze N-oxidation of most of the cyclic and acyclic secondary amines (Hanson et al. 2010) (Tables 3, 4, 5), while P450s tend to catalyze N-dealkylation reactions because of the chemical mechanisms involved (Seto and Guengerich 1993).

Fig. 1
figure 1

Human FMO enzymes participating in the metabolism of drugs (data calculated for major and minor enzymes from Table 3; a total of 114 drugs used in calculations)

`

Table 3 Examples of drugs as substrates in oxygenation reactions catalyzed by human FMO enzymes
Table 4 Examples of general chemicals as substrates in reactions catalyzed by human FMO enzymes
Table 5 Examples of natural products and physiological compounds as substrates for human FMO enzymes

FAD, NADPH, and O2 are required for the FMO catalyzed reactions, but the FAD is tightly bound to the enzyme and does not need to be added (i.e., acts as a prosthetic group instead of a cofactor (Dixon and Webb 1964)). Of the human FMO enzymes, FMO3 is the prominent enzyme that converts nucleophilic heteroatom-containing chemicals, drugs, and xenobiotics to more polar materials, which are generally more efficiently excreted in the urine. The substrate specificity for FMO3 is distinct from that of FMO1. Of the five FMO families, FMO1 and FMO3 are the most prevalent in drug metabolism in humans (Fig. 1, Table 3). A similar participation pattern of the enzymes was found for general chemicals (Table 4). For natural products and physiological compounds, the most prominent enzymes were FMO3 and FMO1, followed by FMO2 and FMO4, with low participation of FMO5 (Table 5).

In general, FMO enzymes have not been reported to be very inducible. However, induction of FMO4 and FMO5 cDNA has been reported in human hepatocytes by the drug rifampicin (Rae et al. 2001), and the tricyclic antidepressants imipramine and chlorpromazine were reported to upregulate recombinant FMO3 catalyzed methimazole S-oxidation in a concentration-dependent manner (Adali et al. 1998, 1999; Cherrington et al. 1998) (Table 3). In addition, FMO5 mRNA was upregulated in HepG2 cells by the natural product (herbal medicine) St. John’s wort and its active component hyperforin, as well as by the synthetic progestin R5020 in a breast cancer cell line that stably expresses B-receptors (YB cells) (Miller et al. 1997).

Inhibition of FMO3 was reported by dietary indoles such as indole-3-carbinol (contained in Brussels sprouts (Cashman et al. 1999a)) and decreased expression and activity of FMO3 was observed for endogenously formed nitric oxide (Ryu et al. 2004) (Table 6).

Table 6 Examples of natural products and physiological compounds as inhibitors of human FMO enzymes

The potential for adverse reactions due to drug–drug interactions is less likely for drugs predominately metabolized by FMO than for P450 enzymes. However, physiological factors can influence FMO function, and this may have clinical implications (Cashman and Zhang 2006; Ryu et al. 2004). For instance, in the case of mammalian FMO3, which does not appear to be very inducible (vide supra), inter-individual variations in FMO3-dependent metabolism of drugs, other chemicals, and endogenous compounds are more likely to be caused by genetic and ethnic polymorphisms (Cashman 2002b; Cashman et al. 2000; Cashman and Zhang 2002; Hisamuddin and Yang 2007). However, human FMO enzymes can activate drugs (e.g., antibiotics, antibacterial, antitubercular, CNS stimulants), natural products, and general chemicals to toxic products, resulting in adverse reactions (Table 7).

Table 7 Examples of compounds activated to toxic products by human FMO enzymes**

Reactions

Human FMO3 N-oxygenates primary, secondary, and tertiary amines but only human FMO1 is highly efficient at N-oxygenating tertiary amines. Both human FMO1 and FMO3 S-oxygenate many nucleophilic sulfur-containing substrates, and in some cases, reactions proceed with high stereoselectivity (Cashman 2000).

N-oxygenations

The N-oxygenation reactions of primary amines catalyzed by FMO enzymes, which occur without splitting the C–N bond, can result in the formation of toxic nitroso compounds. The reaction usually creates potentially toxic hydroxylamines in the first step, which can be further oxidized into oxime and nitroso compounds (Fig. 2) (e.g., sulfamethoxazole and amphetamine N-oxidation). N-Oxygenations of secondary amines, e.g., cyclic and acyclic secondary amines, are catalyzed by FMO enzymes, and those of N-alkyl- and N-aryl amines are generally catalyzed by both FMO and P450 enzymes (e.g., N-methylamphetamine, Tables 3, 7).

Fig. 2
figure 2

General reaction of N-oxygenation of primary amines by FMO

S-oxygenations

Compounds containing a sulfur atom as a part of the structure are present in physiological compounds such as amino acids and derivatives (e.g., cysteine, methionine, glutathione), lipids, and enzyme cofactors (e.g., biotin, thioredoxin, lipoic acid, coenzyme A) and in natural products (e.g., the toxin amanitin and various compounds isolated from onions, radishes, and watercress). The characteristic odor and healing properties of plants of the genus Allium are attributed to sulfur-containing compounds. A number of drugs and general chemicals (e.g., solvents, insecticides) are substrates for S-oxygenation.

S-Oxygenation reactions (Fig. 3) occur by mechanisms similar to N-oxygenation (vide infra), catalyzed by FMO enzymes (also called sulfoxidases). In addition, P450s may be involved (Rendić 2002). Substrates in these reactions include thiocarbamides, thiones, thioamides, sulfides (aromatic and aliphatic), thiols, and mercaptopurines (Table 2). Some intermediates formed in S-oxidations (e.g., sulfenes, sulfines) are reactive and potentially toxic because they can react with proteins and lipids in cells (Table 7). The final products (S-oxides) of the S-oxygenation reactions may also exert toxic effects (Furnes and Schlenk 2004; Shimizu et al. 2007; Siddens et al. 2014).

Fig. 3
figure 3

Typical oxygenation reaction catalyzed by FMO enzymes, where R denotes part of the molecule and X is a heteroatom, usually N or S

Mechanism of oxygenation of heteroatoms (N- and S-oxygenation)

Compounds possessing a soft nucleophilic heteroatom are substrates of FMO enzymes. Structure–activity studies suggest that in addition to nucleophilicity, the size and charge of potential substrates are important parameters limiting access to the enzyme-bound hydroxylating intermediate form of the enzyme (4a-hydroperoxide) (Ziegler 2002).

The mechanism of oxygenation of nucleophilic groups catalyzed by FMO enzymes is presented in the context of the following three steps (Phillips and Shephard 2019; Siddens et al. 2014; Ziegler 1988) (Fig. 4): (1) NADPH binds to the enzyme and reduces FAD to FADH2 (a rapid reaction). The result is the formation of a ternary complex (Enzyme-FADH2-NADP+). (2) FADH2 binds molecular oxygen, as a co-substrate, and produces a relatively stable C4a-hydroperoxyflavin (also a rapid reaction). The cofactor NADP+ remains attached to the enzyme during the reaction, stabilizing the complex. (3) The C4a-hydroperoxyflavin is a strong electrophile and can oxygenate a nucleophilic group, with an attack of activated oxygen (electrophile) atom from the C4a-hydroperoxyflavin molecule on the nucleophilic atom (nitrogen, sulfur, phosphorus) in the substrate molecule, without prior binding of the substrate to the enzyme. The transfer of the oxygen atom to a substrate (reaction of monooxygenation of the substrate) results in the formation of 4a-hydroxyflavin. (3a) If there is no substrate that can be oxygenated near the enzyme, the C4a-hydroperoxyflavin releases H2O2, the oxidized form of the enzyme, and NADP+. (4) Removal of the water molecule (dehydration) (and release of NADP+ from the complex) regenerates the oxidized form of the enzyme (slow reaction).

Fig. 4
figure 4

Oxygenation of substrates with FMO enzymes

Access to the active form of oxygen on the prosthetic group (flavin) is observed for non-ionizable lipophilic amines and amines that are found in the form of mono-cations at physiological pH (step 3). Amines that possess two cationic groups at physiological pH (and amines with one or more anionic groups) cannot approach the active site and are not preferred substrates for FMO enzymes. These structural requirements prevent many endogenous substances from being substrates of the enzymes.

The catalytic cycle and mechanism of monooxygenation catalyzed by FMO enzymes differ significantly from the mechanism that P450s generally use in catalysis. The latter mechanism takes place via an intermediate reactive form of oxygen (FeO3+) that involves radical species (Ziegler 2002). An interesting kinetic feature of the FMO mechanism is that (in general, with a given FMO) the kcat does not vary much and the Km varies among substrates, and the Km is not a measure of inherent affinity for the enzyme (Kd).

Oxidations of ketones by FMOs in Baeyer–Villiger oxidations

FMOs, like other flavin-based monooxygenases in general, utilize flavin 4a-hydroperoxides in their mechanisms (Walsh 1979), with the hydroperoxide acting as an electrophile to oxygenate nitrogen or sulfur (Fig. 4). Flavin 4a-hydroperoxides can also act as nucleophiles, when deprotonated, catalyzing Baeyer–Villiger reactions with carbonyls (Fig. 5) (Walsh and Chen 1988). This is an important reaction in some bacteria, allowing the breaking of a (ketone) ring structure to generate acidic products that can be degraded (e.g., by fatty acid oxidation enzymes) for use as a carbon source. An example of a mammalian enzyme that does this is human FMO5 (Fiorentini et al. 2016; Walsh 1979).

Fig. 5
figure 5

Baeyer–Villiger oxidation of cyclohexanone by the flavoprotein monooxygenase FMO5 (Guengerich and Yoshimoto 2018). Some uncoupling also occurs to generate H2O2 (Fiorentini et al. 2016; Walsh 1979)

FMO5 appears to be adapted for the nucleophilic Baeyer–Villiger chemistry. Examples of reactions attributed to FMO5 are presented in Fig. 5, including four drugs (Fiorentini et al. 2016, 2017; Lai et al. 2011; Meng et al. 2015). This is an interesting reaction, in that the lactones can be readily cleaved to open-chain products by the action of esterases or by non-enzymatic base-catalyzed hydrolysis (Fig. 6).

Fig. 6
figure 6

Some Baeyer–Villiger C–C oxidations of drugs catalyzed by FMO (Guengerich and Yoshimoto 2018)

Thus, a C–C oxygen insertion reaction can be utilized to cleave a C–C bond (Guengerich and Yoshimoto 2018). Recently an alternate flavin mechanism involved in some oxygenations has been shown to involve a flavin N5-oxide (Teufel et al. 2015), but it is unknown whether this intermediate could also be involved in Baeyer–Villiger oxidations.

Substrates and reactions catalyzed by human FMO enzymes

Substrates contain nucleophilic heteroatoms nitrogen, sulfur, phosphorus, or selenium. As already pointed out, the best substrates are cyclic and acyclic amines that are not ionized at physiological pH (Kim and Ziegler 2000; Rettie et al. 1994; Rooseboom et al. 2001; Ziegler 1988). Many drugs possessing nucleophilic heteroatoms in their structure are substrates of these enzymes (Phillips and Shephard 2017; Sawada and Yokosawa 1991; Yamazaki et al. 2014; Cashman, 2000) (Table 3), as well as general chemicals (Table 4) and natural products and physiological compounds (Table 5). Additional substrates are iodides and boron-containing compounds (Jones and Ballou 1986). Drug oxidations are the most studied group of reactions with human FMOs (Tables 3, 4, 5), followed by general chemicals and physiological compounds. In addition, FMO-catalyzed reactions are predominately detoxication reactions, with some examples of contributions of the reactions to bioactivation and formation of toxic products or intermediates (Table 7) (Cashman 2002a).

In many of the cases, the results presented were obtained using purified and recombinant human enzymes expressed in different systems. Although information obtained by studies in such systems is of great value for further research, the results obtained may not be representative of the most important processes occurring in cells or tissues. In addition, some FMO-catalyzed reactions can also be catalyzed by other enzymes in cells, e.g., P450 (Tables 3, 4, 5) and AOX enzymes (Table 3). The participation of P450 enzymes in the metabolism of the FMO substrates by N-oxidation may be a minor contribution to overall metabolic reactions of the compound in some cases (e.g., N-oxygenations of cediranib, C-1311, benzydamine, selegiline, dapsone (Table 3)) or might predominate in the overall metabolic pathway of a compound, e.g., disulfoton, methiocarb, phorate, sulprofos (Table 4), M-04579 (Table 3). Dapsone N-oxygenation is, for instance, catalyzed by several P450 enzymes (P450s 1A2, 2C, 2D6, 2E1, 3A4) with high or intermediate Km values, contributing to its activation to toxic N-hydroxylamine formation (Li et al. 2003; Winter et al. 2000). Dapsone was, in addition, reported to be a substrate-dependent activator of P450 2C9 enzyme activity and thus activating its own oxidation (Hummel et al. 2004). However, P450-catalyzed N-oxidation of dapsone appears to be of minor importance to its overall metabolism (Rendić and Guengerich 2021). In addition to being substrates of P450 enzymes, FMO substrates can also be either strong P450 inhibitors with the potential for drug–drug interactions (e.g., cimetidine Rendić et al. 1983, 1979) (Fig. 9), or weak inhibitors of P450 enzymes with minor potential for inducing drug–drug interactions (e.g., ranitidine) (Fig. 10) (Rendić et al. 1982, 1983).

An additional characteristic of the reactions catalyzed by FMO enzymes is stereoselectivity which, depending on the substrate, can occur with high or low selectivity for a substrate or product formed. Stereoselectivity can occur regarding both N- and S-oxygenations (Tables 3, 4, 5). For instance, no selectivity is observed for product formation by N-oxygenation of two geometric isomers of clomiphene, but high regioselectivity in the conversion of only one of the two isomers of GSK5182 has been reported (the Z-isomer) (Table 3). For sulindac sulfide (a sulindac metabolite), a high degree of stereoselectivity towards the R-isomer was observed (Table 3), and stereoselectivity for N-oxidation is reported for deprenyl (Table 3) and trans-(S)-(-)-N-1′-nicotine oxide (Table 5). Stereoselectivity was also reported for S-oxidation of the L-isomer (FMO4) and the D-isomer of methionine (FMO3) (Table 5), N-oxygenation of (S)-N-methylamphetamine, and S-oxygenation of (R)-sulindac sulfide (Table 3).

In addition to their interaction with FMO and/or P450 enzymes, the drugs/chemicals that interact with FMOs can also induce or inhibit the activity of drug transporters. Clozapine, for instance, is a substrate for FMO3-catalyzed N-oxygenation (Table 3) and also a substrate and/or inhibitor of P450 enzymes (Rendić 2002). The drug is a substrate in P450 1A2, 2D6, and 3A4 catalyzed N-demethylations, and P450 1A2 and 3A4 catalyzed N-oxygenation (Fig. 7) (Buur-Rasmussen and Brøsen 1999; Murray et al. 2018; Tugnait et al. 1999). Furthermore, clozapine N-oxide is reported to be an inhibitor of P450 2B6 and 2C19 enzymes (Giri et al. 2017). In addition, clozapine was reported to be an inhibitor of the drug transporter P-glycoprotein, with the potential to affect the pharmacokinetic properties of co-administered drugs (Liu et al. 2021b; Wang et al. 2006). This example illustrates the complexity of predicting possible drug–drug interactions when a drug is a substrate and/or inhibitor of multiple drug-metabolizing enzymes and/or drug transporters, the properties which are also affected by the properties of the co-administered drug(s).

Fig. 7
figure 7

Clozapine metabolism by human FMO and P450 enzymes (Fang et al. 1998; Tugnait et al. 1999, 1997)

In the reactions of drug substrates of FMO enzymes, the oxygenated products produced are usually more polar (Table 3) and may be more rapidly eliminated from the body or maybe substrates in conjugation reactions. As shown in Table 3, drugs belonging to several important therapeutical categories are substrates of FMO enzymes, e.g., anticancer (cediranib), antiulcer (cimetidine, ranitidine), antidepressants, CNS stimulants (amphetamine and derivatives), and antibacterial drugs (sulfamethoxazole). In some cases, substrates of the FMO enzymes are metabolites produced by the catalytic activity of other enzymes, e.g., S-methyl esonarimod, sulindac sulfide, 3-hydroxynabumetone, tazarotenic acid, and S-methyl-N,N-diethyldithiocarbamate (a disulfiram metabolite). The data also show that in humans FMO3 and FMO1 are the most frequently represented among the FMO enzymes catalyzing the metabolism of drugs (Fig. 1), as well as with the general chemicals (possessing a tertiary amine group, thiols, thiolates, sulfides, thiourea derivatives, and organothiophosphate insecticides) (Table 4), and natural products (e.g., (S)-nicotine, phenethylamine, cysteamine, and methionine-containing compounds) (Table 5). In the case of natural compounds as substrates the enzymes often exert stereoselectivity for a particular isomer (e.g., l-methionine as substrate) or for the formation of a particular isomer (e.g., formation of trans-(S)-(-)-N-1′-nicotine oxide). Also, the products of the reactions are, in some cases, more toxic than the parent compounds (Table 7). Prominent among the reactions producing reactive metabolites are those involving thiourea and derivatives (e.g., thiourea, thioacetazone, ethionamide) as substrates. The metabolite(s) of the compounds are potentially carcinogenic compounds formed by the oxygenation of a sulfur atom. Exposure to thiourea, for instance, can damage bone marrow, causing reductions in the number of red blood cells, white blood cells, and/or blood platelets. Thiourea and derivatives are oxidized by FMO1, FMO2, and FMO3 enzymes with the formation of sulfinic and sulfenic acids (Tables 4, 7); however, the toxicity of thiourea and its derivatives was assigned to the activity of the FMO3 enzyme (Smith and Crespi 2002). In some cases, the same activation reaction (i.e., S-oxidation) might also be catalyzed by P450 enzymes (e.g., activation of the insecticides methiocarb and aldicarb) (Costa et al. 2003; Fujino et al. 2016) (Tables 4, 7).

Examples of substrates and reactions resulting in the formation of non-toxic metabolites

Nicotine (stimulant, agonist at nicotinic acetylcholine receptors)

Nicotine N′-oxygenation is one of the direct detoxication pathways for nicotine, accounting for 4–7% of total urinary nicotine metabolites (Fig. 8). Several FMO enzymes catalyze the reaction, and the role of this reaction increases in subjects with deficient P450 2A6 activity. While all of the recombinant FMO enzymes can mediate nicotine N-oxide formation, FMO1, FMO2, and FMO3 exhibit the highest activity. It was reported that oxidation of nicotine in humans occurs with a certain degree of stereoselectivity, and the formation of trans-nicotine N-1′-oxide catalyzed by FMO3 has been reported as a highly stereoselective probe of human FMO3 (Cashman et al. 1995) (Table 5). In other animal species (rat, swine, rabbit) the oxidation is catalyzed by FMO1, and approximately the same amounts of nicotine isomers are formed (Cashman 2000; Cashman et al. 1992; Park et al. 1993; Perez-Paramo et al. 2019) (Fig. 8).

Fig. 8
figure 8

Oxygenation of nicotine by FMO enzymes

Cimetidine (histamine H2 receptor antagonist)

Cimetidine S-oxygenation has been suggested as a stereoselective functional probe of human FMO3 activity (Cashman 2000; Cashman et al. 1995; Lu et al. 1998). FMO1 produces more of the S-oxide-(−)-enantiomer and FMO3 generates mainly the S-oxide-( +)-enantiomer (with no activity for FMO5) (Hai et al. 2009) (Table 3) (Fig. 9).

Fig. 9
figure 9

Oxygenation of cimetidine by FMO enzymes

Ranitidine (histamine H2 receptor antagonist)

The FMO enzymes in human liver microsomes formed the S- (13–18%) and N-oxides (66–76%) as products. Recombinant human FMO1, FMO2, FMO3, and FMO5 all formed the N-oxide, with FMO3 as the major enzyme. S-Oxide formation catalyzed by FMO3 was reported to be very low, as well as N-oxide formation by FMO5. Based on these results, it has been suggested that ranitidine N-oxide formation can be used as an in vivo probe to determine hepatic FMO3 activity (Cashman 2000; Chung et al. 2000a, b; Overby et al. 1997) (Table 3, Fig. 10).

Fig. 10
figure 10

Oxygenation of ranitidine by FMO enzymes

Chlorpromazine (antipsychotic, phenothiazine)

The N-oxide derivative of chlorpromazine is a stable and pharmacologically active chlorpromazine metabolite. Chlorpromazine is a substrate for both FMO and P450 enzymes (Table 3, Fig. 11). In humans, it is metabolized to 7-hydroxy-N-desmethylchlorpromazine in reactions catalyzed by multiple P450 enzymes (Rendić 2002). Chlorpromazine N-oxide, formed by FMO1 as a major enzyme, is oxidized to a sulfoxide by P450 enzymes (chlorpromazine N,S-dioxide formation) and generates additional metabolites (7-hydroxy, N-desmethyl, 7-hydroxy-N-desmethyl, and N-desmethyl sulfoxide derivatives). The in vivo metabolites are formed in the order: chlorpromazine N-oxide > chlorpromazine sulfoxide > 7-hydroxychlorpromazine > norchlorpromazine sulfoxide > norchlorpromazine. Chlorazepine N-oxide was also reduced back to chlorpromazine (Beckett et al. 1988; Cashman et al. 1993b; Chetty et al. 1994; Jaworski et al. 1990; Ohmiya and Mehendale 1984). This example illustrates the complexity of drug metabolism and activity when metabolic reactions are components of multiple metabolic pathways and effects (Adali et al. 1998, 1999).

Fig. 11
figure 11

Chlorpromazine oxygenations by FMO and P450 enzymes

Dimethylamphetamine (CNS stimulant and anorectic)

N,N-Dimethylamphetamine is an N-methylamphetamine analog with weaker central nervous system stimulant activity. One of the metabolites of dimethylamphetamine in humans is the stable N-oxide (Fig. 12), possessing much lower neurotoxic potential compared to amphetamine and N-methylamphetamine (Lee et al. 2009a, b; Ricaurte et al. 1989). The reaction is catalyzed by FMO1 (as the major enzyme) and FMO3. The reaction catalyzed by FMO1 was reported to be enantioselective for L-N-oxide formation (Table 3).

Fig. 12
figure 12

N-Oxygenation of dimethylamphetamine

Sulfides

Sulfide drugs and general chemicals, or their metabolites, are oxidized to S-oxides by human FMO enzymes (Tables 3, 4). The reaction of sulfide oxidation showed differential structurally dependent stereoselectivity. For instance, sulfoxidation of methyl and ethyl p-tolyl sulfides by recombinant human FMO3 proceeds with little stereochemical preference, whereas sulfoxidation of the n-propyl and n-butyl homologs demonstrated increasing selectivity for formation of the (R)-sulfoxide. In addition, S-oxidation of methyl-p-tolyl sulfide by FMO1 was stereoselective for (R)-sulfoxide formation (Table 4).

Examples of reactions resulting in the formation of toxic metabolites

N-oxygenations

Trimethylamine (an agonist of human TAAR5 (trace amine associated receptor 5))

In humans, FMO3 is polymorphic and can be associated with clinically relevant drug–drug or drug–chemical interactions. FMO3 enzyme polymorphism in humans is related to a rare hereditary disorder of the inability to metabolize trimethylamine. This leads to the accumulation of trimethylamine and to a disorder called trimethylaminuria, which results in a so-called “fish odor” syndrome (Al-Waiz et al. 1987; Dolphin et al. 1997; Phillips et al. 1995).

In humans, trimethylamine is formed mainly from the metabolism of phosphatidylcholine/choline, carnitine, betaine, dimethylglycine, and ergothioneine from food by intestinal microflora in the colon. It is absorbed into the bloodstream and transformed into trimethylamine N-oxide (TMAO) (Fig. 13) by hepatic FMO1 and FMO3 but can be also converted to (mono)methylamine, dimethylamine, and ammonia within the colon. Although the oxidation of trimethylamine to its N-oxide had been known for years, the detrimental effects of TMAO were discovered only recently. Elevated TMAO plasma levels have been correlated with an elevated risk for cardiovascular disease (atherosclerosis and thrombosis) and were implicated in reverse cholesterol transport and glucose and lipid homeostasis. High plasma TMAO levels were also positively associated with the incidence of gallstone disease in humans (Gatarek and Kaluzna-Czaplinska 2021; Papandreou et al. 2020; Schneider et al. 2018; Steel et al. 1988; Zhu et al. 2018). The major enzyme involved in trimethylamine N-oxygenation is FMO3 (Table 5). In some individuals, due to the genetic polymorphism of FMO3, decreased trimethylamine oxidation occurs (Fig. 13) with an accumulation of trimethylamine resulting in “fish odor.” Trimethylamine N-oxide accounts for almost 98% of the administered dose of the parent compound trimethylamine. However, in individuals deficient in the FMO3 the formation of toxic trimethylamine N-oxide is reduced to 80%, with the remainder (i.e., 20%) being present as trimethylamine. This polymorphism in amine metabolism, due to attenuated catalytic activity of FMO3, is heritable (Cashman et al. 2003; Phillips and Shephard 2020; Shimizu et al. 2014).

Fig. 13
figure 13

N-Oxygenation of trimethylamine by FMO3

The ratio of trimethylamine to TMAO in urine is used as an index of FMO3 activity, FMO3 polymorphism, and the occurrence of trimethylaminuria.

Amphetamine (CNS stimulant, anorexic)

Multiple mechanisms are involved and interact to promote neurotoxicity from amphetamine and derivatives, which are widely abused psychostimulant drugs (Carvalho et al. 2012; Yamamoto et al. 2010). Oxygenation of the amino group of amphetamine occurs less in humans because deamination and aromatic hydroxylation predominate, catalyzed by P450 enzymes (Bach et al. 1999; Miranda et al. 2007). N-Oxygenation of amphetamine is catalyzed by FMO3, and reactive and toxic metabolites are formed that can contribute to the toxic effects of amphetamine by participating in the autooxidation of dopamine, norepinephrine, and serotonin (Tables 3, 7).

Potential toxic effects are ascribed to amphetamine hydroxylamine. A proposed mechanism of amphetamine activation is N-oxygenation to a hydroxylamine in the first step, which is then re-oxygenated with FMO3 to form an unstable intermediate that, after spontaneous dehydration, is transformed into a trans-oxime (Cashman et al. 1999b; Szöko et al. 2004) (Fig. 14).

Fig. 14
figure 14

N-Oxygenation of amphetamine

N-Methylamphetamine (CNS psychostimulant)

N-Methylamphetamine (methamphetamine) is an illicit, highly addictive psychostimulant amphetamine derivative that is widely abused. Large doses of the drug are associated with serious neuropsychiatric consequences including agitation, anxiety, hallucinations, paranoia, and psychosis (Jayanthi et al. 2021). N-Methylamphetamine can severely damage the central nervous system and is toxic to the cardiovascular system (Halpin et al. 2014; Tan et al. 2021; Zhao et al. 2021). Metabolism of N-methylamphetamine proceeds with the initial formation of N-methylamphetamine hydroxylamine, and the final product is phenyl propanone (Tables 3, 7) (Fig. 15). The formation of phenyl propanone oxime and the nitrone are proposed as part of an overall detoxication process, with the potentially toxic effects ascribed to N-methylamphetamine hydroxylamine (Cashman et al. 1999b; Szöko et al. 2004).

Fig. 15
figure 15

N-Oxygenation of N-methylamphetamine

Arecoline (tetrahydropyridine alkaloid)

The alkaloid arecoline, a major constituent of areca nuts, has been classified as a Class I carcinogen by the International Agency for Research on Cancer (IARC) (IARC Working Group on the Evaluation of Carcinogenic Risks to Humans, 2004). Arecoline is converted to the metabolite arecoline N-oxide by human FMO1 and FMO3, with FMO1 as the major enzyme (Tables 5, 7) (Fig. 16). Arecoline N-oxide was shown to be mutagenic in bacterial tester strains and to induce DNA damage in mammalian test systems, including cultured fibroblasts. The metabolite arecoline N-oxide is further converted to mercapturic acid derivatives in vivo (Das and Giri 2020; Giri et al. 2007; Lin et al. 2011; Oliveira et al. 2021).

Fig. 16
figure 16

N-Oxygenation of arecoline by FMO enzymes

S-Oxygenations

Substances with a sulfur atom can be oxygenated with FMO enzymes to form electrophilic intermediates (e.g., thiols, thioamide, 2-mercaptoimidazole, thiocarbamate, thiocarbamide metabolites). Such electrophilic metabolites can bind to cellular proteins and inactivate enzymes in the endoplasmic reticulum, e.g., P450s (Başaran and Can Eke 2017; Jones and Ballou 1986).

Thiourea and derivatives (organosulfur compounds)

Thiourea (also called thiocarbamide) is a pro-carcinogenic, moderate to a highly toxic substance that is oxidized to carcinogenic products by FMO enzymes. The thiourea moiety is part of chemicals with different applications, including rodenticides, bactericides, components used in the manufacture of rubber, and therapeutic agents. Some derivatives of thiourea are known toxins (e.g., phenylthiourea) (Henderson et al. 2014; Smith and Crespi 2002).

Thiourea is oxygenated via a sulfenic acid to a sulfinic acid by human FMO1, FMO2, and FMO3, with FMO2 as a major enzyme (Tables 4, 7) (Fig. 17). The sulfinic acid formed can be detoxicated in the cells by reaction with glutathione. Similarly, N-substituted derivatives of thiourea (e.g., N′-(4-imidazole-ethyl)thiourea derivatives) exerted cytotoxicity and are activated by oxygenation of the sulfur atom to sulfenic acids (Furnes and Schlenk 2004; Kim and Ziegler 2000; Onderwater et al. 2006; Smith and Crespi 2002).

Fig. 17
figure 17

Oxygenation of thiourea by FMO enzymes

Fenthion (organophosphate, insecticide)

Fenthion, an inhibitor of human acetylcholinesterase, is a substrate in the reaction of S-oxygenation catalyzed by FMO1, FMO3, and FMO5, with FMO1 being the major enzyme. The reaction is characterized by high Km values and by the stereoselective formation of (R)-( +)- sulfoxide (Table 4) (Fig. 18). At lower concentrations, fenthion is predominately metabolized by multiple P450 enzymes, with P450 1A2 as the major one (Furnes and Schlenk 2004, 2005; Gadepalli et al. 2007; Leoni et al. 2008) (Fig. 18).

Fig. 18
figure 18

Oxygenation of fenthion by FMO enzymes

Monoamine oxidase (MAO)

We previously reported that human MAOs participate in ~ 1% of the metabolism of xenobiotic and physiological compounds, including natural products. In the metabolism of general chemicals, MAO enzymes participate in ~ 2%, drugs ~ 1%, and natural and physiological chemicals ~ 1% (Rendić and Guengerich 2015). The previous analysis indicated more extensive participation of MAO enzymes in the metabolism of general chemicals when compared to the metabolism of drugs and natural products and physiological compounds, but this pattern may reflect more basic studies and efforts at drug discovery (Rendić and Guengerich 2012, 2015).

Enzymes

Two MAO enzymes are known (MAO A and MAO B), which are encoded by the MAOA and MAOB genes. The enzymes are primarily involved in the catalytic oxidative deamination of endogenous monoamines (Bach et al. 1988; Benedetti 2001; Bortolato et al. 2008; Bortolato and Shih 2011; Edmondson and Binda 2018; Grimsby et al. 1990; Ramsay 2012; Shih et al. 1990; Strolin Benedetti et al. 2007) The MAOs are mitochondrial, membrane-bound enzymes, and are located in many tissues, of which the most significant may be the brain. The enzymes are present also in the liver, where they catalyze the oxidative deamination of some xenobiotics.

MAO A is present in the brain, small intestine, heart, placenta, liver, portal system, and peripheral adrenergic neurons, and it is selective for the metabolism of norepinephrine and serotonin. MAO B is found in blood platelets, cerebral glial cells, and hepatic cells and is relatively selective for the metabolism of benzylamine and phenylethylamine. Physiological substrates are amines that are oxidized to aldehydes, which may be reduced by aldehyde reductase to alcohols. In vivo inhibition of MAO with either irreversible or nonselective compounds permits the up-take of high concentrations of tyramine and other sympathomimetic molecules into the blood circulation, where they gain access to peripheral adrenergic neurons, trigger catecholamine release, and cause a marked and rapid increase in blood pressure (Lavian et al. 1993).

Substrates

The substrates are nitrogen-containing compounds, including primary, secondary, and tertiary amines (Kalgutkar et al. 2001; Strolin Benedetti et al. 2007) (Tables 8, 9, 10). Substrates have also been synthesized as prodrugs (e.g., dopamine prodrugs synthetized as esters, amides, dimeric amides, carrier-mediated, peptide transport-mediated, cyclic, chemical delivery systems) to enhance their bioavailability in the treatment of Parkinson’s disease (Haddad et al. 2017; Sozio et al. 2012). Endogenous substrates include biogenic and dietary amines, monoamine hormones, and neurotransmitters such as serotonin, dopamine, norepinephrine, and epinephrine, as well as tyramine, tryptamine, 2-phenylethylamine, 5-hydroxytryptamine, monoacetyl putrescine (a precursor to γ-aminobutyric acid (GABA), adrenaline, and metanephrine) (Bortolato and Shih 2011). Similar to FMO enzymes, the substrates of MAO enzymes are often substrates for other drug-metabolizing enzymes as well (e.g., P450 and/or FMO enzymes).

Table 8 Examples of drugs as substrates in oxidation reactions catalyzed by human MAO enzymes
Table 9 Examples of natural products and physiological compounds as substrates in oxidation reactions catalyzed by human MAO enzymes
Table 10 Examples of general chemicals as substrates in oxidation reactions catalyzed by human MAO enzymes

Inhibitors

The inhibitors of MAO enzymes are developed and tested either as selective or nonselective reversible or irreversible inhibitors. Many compounds (drugs, natural products, as well as general chemicals) have been shown to inhibit MAO enzymes. In the clinic, drugs are used either as selective or nonselective MAO inhibitors in the therapy of several neuropsychiatric disorders (mood disorders, Parkinson's disease, Alzheimer’s disease) (Table 11). Tested natural products have shown a variety of activities and some of them were selective and strong as either MAO B (e.g., (−)-maackiain and (−)-medicarpin) or MAO A (e.g., apigenin) inhibitors (Table 12). In addition, extensive work has been done to synthesize derivatives of natural products as MAO inhibitors to be used as CNS drugs (Gulcan and Orhan 2020; Lu et al. 2013; Mathew and Kim 2020) (Table 13).

Table 11 Drugs as inhibitors of human MAO enzymes
Table 12 Examples of natural products as inhibitors of human MAO enzymes (Kong et al. 2004; Lee et al. 2008)
Table 13 Examples of general chemicals as inhibitors of human MAO enzymes

Activators and inducers

Valproic acid, which has been widely used in clinics for the treatment of multiple neuropsychiatric disorders such as epilepsy and bipolar disorder, exerts its activity by regulating the brain levels of serotonin. The compound was reported to increase MAO A catalytic activity, mRNA levels, and promoter activity (Wu and Shih 2011).

Bavachin, a Psoralea corylifolia L. seed compound, has been also reported to be an activator of the activity (Zarmouh et al. 2015), along with clomipramine (Reid et al. 1988).

Reactions

The general reaction catalyzed by MAO enzymes (Ramsay and Albreht 2018) is shown in Fig. 19.

Fig. 19
figure 19

Typical reaction catalyzed by MAO enzymes, where R denotes part of the molecule

MAO enzymes catalyze oxidative deamination reactions, including cleavage of C–N bonds with the formation of several chemical species with neurotoxic potential, e.g., hydrogen peroxide, ammonia, and aldehydes. As a consequence, prolonged excessive activity of these enzymes can lead to mitochondrial damage and neurodegenerative disorders.

Oxidative deamination reactions are also catalyzed by P450 enzymes. However, the mechanism catalyzed by MAO enzymes differs from the reaction catalyzed by P450s in that one of the products of the overall reaction is hydrogen peroxide (Fig. 19), while in the reactions catalyzed by cytochromes P450 the product is a water molecule, i.e. fully reduced oxygen (Guengerich 2022).

Substrate oxidation by MAO enzymes

The MAO enzymes share similar overall structures, with nearly identical FAD-binding domains, but contain varied substrate binding sites. It should be noted that, in contrast to the FMOs, AOX, XOR, and NADPH-P450 reductase, the MAO enzymes have the flavin covalently attached to the protein via a histidine residue. As flavoprotein oxidases, they catalyze substrate oxidation via two half-reactions. In the reductive half-reaction two hydrogen atoms are transferred to the MAO FAD complex when it accepts a hydride equivalent from the substrate, while in the oxidative step the MAO FADH2 complex is oxidized to form MAO FAD by molecular oxygen (generating H2O2) (Figs. 20, 21). Due to the ability of the flavin prosthetic group to accept either one or two electrons (i.e., as a biological “transformer” (Walsh 1979)), several mechanisms have been proposed for the transfer of electrons from the substrate to the prosthetic group (Behl et al. 2021; Edmondson et al. 2007; Fitzpatrick 2010; Gaweska and Fitzpatrick 2011; Ramsay and Albreht 2018; Scrutton 2004).

Fig. 20
figure 20

Substrate oxidation by MAO enzymes (Edmondson et al. 2007)

Fig. 21
figure 21

Deamination of benzylamines by MAO enzymes

Drugs as substrates of MAO enzymes

Numerous drugs possessing a nucleophilic heteroatom are substrates of MAO enzymes (Table 8). Knowledge of the involvement of either MAO A, MAO B, or both enzymes in the metabolism of a drug allows for the prediction of drug–drug interactions with selective or non-selective MAO inhibitors. It should be emphasized that these are mitochondrial enzymes and that in vitro studies with microsomes will not include these enzymes or evaluate their metabolic potential. The metabolism of a drug that is deaminated by both forms of MAO is not necessarily inhibited in vivo by selective MAO A or MAO B inhibitors. If a drug is metabolized by MAOs, competitive interactions can occur with other drugs that are MAO substrates, e.g., with β-adrenoceptor agonists and antagonists, prodrugs of dopamine, and serotonin 5-HT1-receptor agonists, as well as with primaquine, flurazepam, and citalopram (Benedetti 2001; Masuo et al. 2017).

Drugs or drug metabolites that are substrates for human MAOs include β-blockers (i.e., amines formed by dealkylations of β-blockers), primaquine, β-phenylethylamine, phenelzine (also an irreversible inhibitor), almotriptan, bicifadine, citalopram, and its active metabolite desmethylcitalopram, rizatriptan, and zolmitriptan (Table 8).

The drug ozanimod is oxidatively deaminated to a pharmacologically active metabolite by MAO B, yielding the major circulating active compound. The reaction follows a prior N-dealkylation reaction catalyzed by P450 3A4 (Fig. 22) (Table 8). Also involved in the overall metabolism of ozanimod are P450s 1A1 and 2C8, aldehyde dehydrogenase, and alcohol dehydrogenase, plus reductive metabolism by gut microflora (Surapaneni et al. 2021; Tran et al. 2020).

Fig. 22
figure 22

Ozanimod metabolism by P450 and MAO enzymes

Drugs as MAO inhibitors

In addition to being substrates of MAO enzymes, many nitrogen-containing drugs are also MAO inhibitors (Table 11) and were among the first agents shown to be efficacious in the treatment of clinical depression (Fernandez and Chen 2007; Kalgutkar et al. 2001; Suchting et al. 2021). For instance, the therapeutic effects of some antidepressants, hydrazine derivatives (e.g., iproniazid), and tranylcypromine are based on irreversible inhibition of the MAO enzyme and result in the accumulation of sympathetic amines in adrenergic neurons.

As already mentioned, the drugs used in clinical practice are either nonselective and irreversible MAO enzymes inhibitors or selective inhibitors for either MAO A or MAO B enzymes. Some irreversible inhibitors include rasagiline (MAO A and B, selective MAO B inhibitor), tranylcypromine (MAO A and B, nonselective), iproniazid (MAO A and B, nonselective), phenelzine (MAO A and B, nonselective inhibitor), selegiline (MAO A and B, selective MAO B inhibitor at lower concentrations/doses), pargyline (MAO A and MAO B, partially MAO B selective inhibitor), iproniazid (MAO A and B, nonselective inhibitor), clorgyline (MAO A, MAO B, selective MAO A inhibitor), ladostigil (MAO A and B, non-selective inhibitor), and isocarboxazid (MAO A and B, non-selective inhibitor). Some selective reversible MAO enzyme inhibitors are lazabemide (MAO B selective inhibitor), befloxatone (MAO A and B, selective MAO A inhibitor), toloxatone (MAO A and B, selective MAO A inhibitor), brofaromine (MAO A and B, selective MAO A inhibitor), and moclobemide (MAO A selective inhibitor) (Table 11).

Due to the observed toxic effects, irreversible inhibitors of MAO enzymes have been largely replaced in therapy with selective reversible inhibitors.

Antidepressant drugs, besides inhibiting the active uptake of amines into presynaptic cells (Stahl 1998), also exert inhibitory activity on MAO enzymes with potencies dependent on the model and experimental conditions used. For instance, when testing mitochondrial MAO activity in mouse, rat, dog, and monkey brains with antidepressant drugs (zimeldine, imipramine, maprotiline, and nomifensine) (which inhibit MAO A and MAO B at high concentrations), inhibition was dependent on the species used and experimental conditions applied. Imipramine, for instance, inhibited MAO B more strongly than MAO A activity in mouse and rat brains. When dog and monkey brains were investigated, MAO A activity was inhibited with greater potency than MAO B activity at high concentrations of imipramine; at low concentrations, however, MAO B activity was more strongly inhibited. Also, maprotiline and nomifensine inhibited mouse and rat brain MAO B activity more strongly than MAO A activity, while the inverse was found for dog and monkey brain models (Egashira et al. 1999).

As an example, the non-selective MAO A and MAO B irreversible inhibitor phenelzine (Table 11, Fig. 23) elevates brain levels of the monoamine neurotransmitters 5-hydroxytryptamine (serotonin), noradrenaline, and dopamine. Phenelzine is also a substrate for MAO enzymes, and different metabolites are formed including β-phenylethylamine, phenylacetic acid, p-hydroxyphenyl acetic acid, β-phenylethylidenehydrazine, and phenylethyldiazenephenylethylidenehydrazine. Of these metabolites, neuroprotective/neuro-rescue activity has been suggested for the metabolite β-phenylethylidenehydrazine and irreversible inactivation of MAO enzymes has been ascribed to the formation of phenethyl free radicals (Ortiz de Montellano et al. 1983; Rumyantseva et al. 1991). Thus, besides its MAO inhibiting activity, phenelzine also elevates brain levels of γ-aminobutyric acid (GABA) which may also contribute to its anxiolytic effects, and the effects ascribed to the phenelzine intermediate metabolite β-phenylethylidenehydrazine, a weak MAO inhibitor (Baker et al. 2019; Parent et al. 2002). Phenelzine may also ameliorate the effects of oxidative stress by reducing the formation of reactive metabolites (aldehydes, hydrogen peroxide, ammonia/ammonia derivatives) produced by the interaction of MAO with biogenic amines, as well as by inhibiting primary amine oxidase (Baker et al. 2019; Matveychuk et al. 2021). This example illustrates the complex interactions of the parent drug and its metabolite(s) on the final effects.

Fig. 23
figure 23

Phenelzine oxidation to β-phenylethylidenehydrazine

The first generation of non-selective (iproniazid, tranylcypromine, phenelzine) and irreversible MAO A inhibitors was shown to produce associations with the “cheese reaction,” whereas MAO B inhibitors (used at the recommended selective dosage) did not produce the effect. The cheese reaction occurs because of potentiation of the sympathomimetic activity of ingested tyramine present in cheese and other fermented food. This cheese reaction provoked by inhibition of MAO A may consequently produce a hypertensive crisis due to the increased release of norepinephrine. In contrast to irreversible MAO A inhibitors, reversible MAO A inhibitors (e.g., the antidepressants moclobemide and brofaromine) exert limited tyramine potentiation activity (McCabe 1986; Youdim and Weinstock 2004). In addition, some of the early MAO inhibitors have been withdrawn from the market due to hepatotoxic reactions (e.g., nialamide, pargyline). Because of the observed toxic effects, nonselective irreversible inhibitors of MAO enzymes have been replaced with selective reversible inhibitors in clinical therapy. At present, drugs that are inhibitors of MAO A are used and investigated for the treatment of depression, while selective MAO B inhibitors (e.g., rasagiline, selegiline), are used in the treatment of Parkinson's disease, avoiding severe side effects. It has been suggested that MAO B inhibitor drugs might be effective in the treatment of Alzheimer’s disease (Finberg 2014; Özdemir et al. 2021; Shulman et al. 2013; Szökő et al. 2018; Yamada and Yasuhara 2004; Youdim 1975; Youdim and Bakhle 2006).

Natural products and physiological compounds, derivatives, preparations, and MAO enzymes

The physiological substrates of MAO enzymes are brain neurotransmitters (e.g., serotonin, dopamine, norepinephrine, and epinephrine), as well as trace amines (e.g., tyramine, tryptamine, 2-phenylethylamine, octopamine, 3-iodothyronamine (Table 9). The products of oxidative deamination are aldehydes and H2O2, both of which have some potential toxicity in cells (Tables 9, 14). The formation of potentially toxic metabolites has been associated with neurodegenerative disorders of the central nervous system such as Parkinson’s disease and dementia. Thus, the reactions of physiological compounds catalyzed by MAO enzymes are examples of the bioactivation of non-toxic amines to potentially toxic metabolites. However, in the cells, the aldehydes that are formed are either oxidized to polar carboxylic acids by the activity of aldehyde dehydrogenases (ALDH) or reduced to alcohols or glycols by aldehyde reductases (AKR enzymes). These polar products can often be excreted through kidneys and/or participate in conjugation reactions. Dopamine and norepinephrine can alternatively participate as substrates in methylation reactions catalyzed by catechol O-methyltransferase (COMT) to form 3-methoxytyramine and epinephrine, respectively, or participate in conjugation reactions such as sulfoconjugation (Behl et al. 2021; Buu 1985; Danielczyk et al. 1988; Ji et al. 2005; Rivett et al. 1982).

Table 14 Examples of compounds activated to toxic products by human MAO enzymes**

Some natural products are substrates of MAO enzymes, but there is also a growing interest in testing natural products and compounds as inhibitors of MAO enzymes (Table 12) for possible use in the treatment of Parkinson’s disease (Zarmouh et al. 2016) or to explain possible side effects or their toxicity when ingested.

When using natural products, care should be taken because the same preparation may contain a compound that inhibits the enzymes, as well as compounds that act as enzyme activators. For example, ethanolic extracts of the seeds of Psoralea corylifolia L. contain flavanone bavachinin, which showed competitive MAO A and MAO B inhibition. P. corylifolia L. extracts also contain its analog bavachin, which has stimulatory properties (Zarmouh et al. 2015).

General chemicals and synthetic derivatives of natural products as MAO enzyme inhibitors

In the group of general chemicals as substrates of MAO enzymes, special attention has been focused on the tetrahydropyridine compound MPTP, which is oxidized to the neurotoxic products MPDP+ and MPP+ by MAO B as the major enzyme. During the reaction the enzyme is inactivated (Tables 10, 13, 14, Fig. 26).

Some of the synthetic MAO enzyme inhibitors are compounds with structures based on scaffolds of natural compounds known to be MAO inhibitors, e.g., caffeine, coumarin, piperazine, and chalcone (a structural isomer of coumarin) (Tables 12, 13). Caffeine, an adenosine receptor antagonist (A2A), is a weak and reversible MAO A and MAO B inhibitor, both in vivo and in vitro (Grzelczyk et al. 2021; Haj Ahmed et al. 2020). Its structure has been used to design compounds having both A2A receptor antagonist and MAO A and MAO B inhibition activity. The compounds have been developed with the potential for treating Parkinson's disease. Although structural modifications of caffeine led to strong MAO inhibitors, the MAO inhibitory activity of caffeine itself is not likely to be of pharmacological relevance in typical coffee consumption (Petzer et al. 2013; Petzer and Petzer 2015).

The β-carboline alkaloids, which are also components of coffee (and also present in cigarette smoke), were reported to be reversible, competitive, and strong inhibitors of MAO enzymes and linked to a reported lower incidence of Parkinson’s disease in coffee drinkers and cigarette smokers (Herraiz and Chaparro 2005, 2006). However, the MAO inhibitory activity of natural products may be dependent on and affected by the type of product used (e.g., type of coffee), as well as by the method of preparation of the sample for testing (e.g., light or dark roasted coffee beans) (Grzelczyk et al. 2021) (Table 12).

Coumarin, for instance, exhibited nonselective intermediary MAO A and MAO B inhibitory activity, but some of its natural derivatives exhibited selective strong MAO A (osthenol) or MAO B inhibitory activity (rutamarin) (Table 12). Furthermore, some synthetic coumarin derivatives exerted strong, selective inhibition of either MAO A or MAO B activity. Many other derivatives of various natural compound-based structures (e.g., indoles, chromones, chalcones, carboxamides, benzylamine, sulfonamide, benzofuran, pyrazole, pyrrole, quinazolinone, and others) were synthesized and reported to exhibit strong, selective inhibition of either MAO A or MAO B activity (Table 13) (Patil et al. 2013).

Stereoselective inhibition of MAO enzymes was reported for enantiomers of the 8-aminoquinoline derivative NPC1161. Racemic NPC1161 exerted both MAO A and MAO B inhibitory activity with 3.7-fold selectivity of MAO A compared to MAO B, while the (S)-( +)- enantiomer was shown to be an intermediate (MAO A) and strong (MAO B) mixed-type irreversible inhibitor with about tenfold selectivity for inhibition of MAO B over MAO A. The (R)-(−)-enantiomer was shown to be a mixed-type nonselective intermediate reversible inhibitor (Table 13). Stereoselective MAO inhibition was also observed in the interaction of enantiomers of antimalarial drug primaquine. Racemic primaquine and (R)-( −)-primaquine were weak and very weak inhibitors, respectively, both being nonselective inhibitors. (S)-( +)-Primaquine was also a weak inhibitor but showed 1.5-fold selectivity for inhibition of MAO A over MAO B (Chaurasiya et al. 2021) (Table 11).

Examples of compounds bioactivated to toxic products by MAO enzymes

Examples of compounds that are bioactivated to toxic products by MAO enzymes include physiological compounds (e.g., neurotransmitters dopamine, serotonin, noradrenaline, the biogenic amine kynuramine) that are metabolized to toxic aldehydes, the antidepressant drug nomifensine (by forming dihydroisoquinolinium ions exerting risks of anemia and hepatotoxicity), and the general chemical MPTP, which was shown to be mechanism-based inhibitor inactivating the enzyme and forming 1,4-dihydropyridine adducts. Benzylamine, widely used as a model substrate for MAO B, is converted to toxic benzaldehyde, which is consequently reduced and deactivated to an alcohol by aldehyde dehydrogenase (Table 14).

Examples of metabolic reactions of MAO substrates

Tyramine

The aromatic amine tyramine, which is both a natural product and physiological compound, is oxidatively deaminated preferentially by MAO A. The product of its metabolism is the toxic 4-hydroxyacetaldehyde, which is converted to nontoxic 4-hydroxyphenyl acetic acid by aldehyde dehydrogenase (ALDH) (Tables 9, 14) (Fig. 24). Alternatively, tyramine can be hydroxylated to dopamine by P450 2D6 in a reaction considered as the main elimination/detoxication pathway for tyramine (Niwa et al. 2004). In a minor reaction, tyramine is converted to tyrosol by alcohol dehydrogenase (ADH) and, in human liver microsomes, to a trans-oxime by FMO3 through a hydroxylamine intermediate (Lin and Cashman 1997a; Niwa et al. 2011; Phillips and Shephard 2019). The oxidative deamination reaction can potentially be inhibited by MAO A inhibitors, resulting in an enhanced concentration of other sympathomimetics in peripheral adrenergic neurons and causing a rapid increase in blood pressure and the onset of the cheese reaction (McCabe 1986).

Fig. 24
figure 24

Deamination of tyramine by MAO enzymes

Dopamine and other neurotransmitters

MAO plays a central role in the metabolism of the neurotransmitter dopamine, as well as norepinephrine and serotonin (Table 9). Dopamine metabolism is complex (Meiser et al. 2013) and, in addition to MAO enzymes, dopamine is also a substrate for catechol O-methyl transferase (COMT). 3,4-Dihydroxyphenylacetaldehyde (DOPAL), a product of the MAO-catalyzed deamination reaction, is toxic and is converted to 3,4-dihydroxyphenylacetic acid (DOPAC) by aldehyde dehydrogenase (ALDH), which rapidly exits the neurons and is also a substrate for COMT, producing homovanillic acid (Fig. 25). In addition to DOPAL, the oxidative deamination produces H2O2, which (in the presence of divalent metal atoms) may form hydroxyl radicals (OH⋅). The formation of toxic species from dopamine (and also from other neurotransmitter substrates of MAO enzymes) has been suggested to contribute to catecholaminergic denervation in Parkinson’s disease. Cytoplasmic dopamine levels are maintained at low, non-toxic levels by the combined activity of the vesicular monoamine transporter (VMAT) and MAO and ALDH enzymes (Goldstein 2020; Goldstein et al. 2012).

Fig. 25
figure 25

Activation and deactivation of dopamine

MPTP

MPTP, a selective nigrostriatal neurotoxin, is bioactivated by MAO B (and less effectively by MAO A) to 2,3-MPDP+, and this intermediate undergoes further oxidation to MPP+ by MAOs (Fig. 26). MPTP and its two primary metabolites are competitive and mechanism-based inactivators of MAO A and MAO B enzymes (Trevor et al. 1988, 1987b). To express the selective nigrostriatal neurotoxicity of MPTP, bioactivation by MAO B is required, leading to the formation of the potentially reactive products MPDP+ and (the 4-electron oxidation product) MPP+. The latter product accumulates in brain striatal tissue, is a substrate for dopaminergic active uptake systems, and is an inhibitor of mitochondrial NADH dehydrogenase, a respiratory chain enzyme located in the inner mitochondrial membrane (Peterson et al. 1985; Singer et al. 1988; Trevor et al. 1987a). Both reactions, MPTP activation to MPP+ and its deactivation by N-demethylation, are catalyzed by MAO B and P450s (Fig. 26) (Bajpai et al. 2013; Hanna et al. 2001; Herraiz et al. 2006; Nakamura et al. 2020; Trevor et al. 1987a; Uehara et al. 2015).

Fig. 26
figure 26

Bioactivation and detoxication of MPTP by MAO and P450 enzymes

NAD(P)H-quinone oxidoreductase (NQO) enzymes

NAD(P)H quinone oxidoreductase 1 (NQO1) and NAD(P)H quinone oxidoreductase 2 (NQO2), are homodimeric flavoproteins containing one molecule of non-covalently bound FAD per monomer. These enzymes are members of a larger mammalian quinone oxidoreductase family and catalyze the reduction of quinones and similar molecules possessing quinone-like structures, e.g. quinone imines, benzotriazine oxides, tocopherols (Fig. 27). Nitro groups are also reduced by NQO enzymes. These enzymes use both NADH and NADPH and were termed “DT diaphoras” in the early literature because they use both DPNH (NADH) and TPNH (NADPH) (former names used for these pyridine nucleotides) (Ernster et al. 1962).

Fig. 27
figure 27

Typical reaction catalyzed by NQO enzymes

These enzymes are generally considered to be detoxicating enzymes that protect cells by catalyzing the 2-electron reduction of quinones and thus participate in the protection of cells against toxicity. NQO enzymes are constitutively expressed in a variety of tissues and also in many solid tumors. The latter property has been considered in the context of potential targets for the activation of certain bioreductive anticancer agents (e.g., activation of the anticancer drug mitomycin C in tumor cells) (Siegel et al. 2004; Workman 1994). In our previous reports, human quinone reductase enzymes were classified in the group of “other oxidoreductases” (Rendić and Guengerich 2012, 2015). These enzymes participate in < 1% of the metabolism of xenobiotics and natural products, including drugs. The enzymes were also classified in the group of enzymes participating to the extent of < 4% of the activation of chemical carcinogens.

Changes in the activity of NQO1 are associated with different pathologies (including cancer and cardiovascular and neurodegenerative diseases), and these properties have been considered in the context of potential targets for the treatment of the diseases. Induction or depletion (knockout) of NQO1 was shown to be associated with decreased or increased susceptibilities to oxidative stress, respectively. Human NQO1 is often over-expressed in cancer cells, and the enzyme has been considered as a possible drug target. Two common polymorphic forms of human NQO1, pR139W and pP187S, were found to be associated with an increased risk of several forms of cancer. Dicumarol and some structurally related compounds act as competitive inhibitors of both variants. In addition, NQO1 was reported to be inhibited by nicotinamide, and resveratrol inhibited both NQO1 and NQO2 (Megarity and Timson 2019; Nolan et al. 2012; Pey et al. 2019). On the other hand, quercetin was shown to increase NQO1 transcription in human MCF-7 human breast cells (Valerio et al. 2001), and resveratrol increased NQO1 protein levels in K562 cells (Hsieh et al. 2006).

A review of the literature and examples of drugs, physiological, and environmental compounds that interact with NQO1 and NQO2 enzymes is provided in a recently published article (Rashid et al. 2021). The compounds are presented as being either activated (e.g., mitomycin C, doxorubicin, porfiromycin) or inactivated (e.g., acetaminophen, menadione, amrubicin) by NQO enzymes. One of the physiological substrates of the NQO1 enzyme is the highly unstable DOPA quinone, formed by auto-oxidation of dopamine catechol. DOPA-quinone may induce neuronal damage resulting from the formation of reactive oxygen species, e.g. superoxide radicals and hydrogen peroxide. In a mouse model, that quinone formed up-regulates astroglial NQO, which might reduce the potentially toxic dopamine quinone to more stable hydroquinone, a detoxication reaction catalyzed by NQO enzymes (Drukarch et al. 2001; van Muiswinkel et al. 2000).

Molybdenum-containing hydroxylases

Molybdenum hydroxylases are cytosolic molybdoflavoproteins with a molecular mass of approximately 300 kDa (Hille 2005). Human molybdenum-containing hydroxylase enzymes were classified in the group of “other oxidoreductases” in our previous work (Rendić and Guengerich 2012); (Rendić and Guengerich 2015). According to this classification, as mentioned before, the enzymes from this group participate in < 1% of the metabolism of xenobiotics and natural products, including drugs. The enzymes were also classified in the group of enzymes participating to the extent of < 4% of the activation of chemical carcinogens (Rendić and Guengerich 2012); (Rendić and Guengerich 2015).

Enzymes

The molybdoflavoenzyme family in humans is composed of aldehyde oxidase (AOX), xanthine oxidoreductase (XOR), sulfite oxidase, and an enzyme known as mitochondrial amidoxime-reducing component (Terao et al. 2020) (Tables 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29). This review focuses on AOX and XOR, which are known to play roles in the metabolism of drugs and other xenobiotics (Tables 15, 19, 23). AOX and XOR enzymes also catalyze the metabolism of physiological compounds (Tables 17, 25) and are involved in both detoxications and activation of substrates to toxic/pharmacologically active intermediates or products (Tables 21, 22, 29).

Table 15 Examples of drug substrates in reactions catalyzed by human AOX1
Table 16 Drug inhibitors of human AOX1
Table 17 Natural products and physiological compounds as substrates in reactions catalyzed by human AOX1
Table 18 Natural products and physiological compounds as inhibitors of human AOX1
Table 19 General chemicals as substrates in reactions catalyzed by human AOX1
Table 20 General chemicals as inhibitors of human AOX1
Table 21 Compounds activated to pharmacologically active metabolites by human AOX1
Table 22 Compounds activated to toxic metabolites by human AOX1
Table 23 Examples of drugs as substrates in reactions catalyzed by human XOR
Table 24 Drugs that inhibit human XOR
Table 25 Natural products and physiological compounds as substrates in oxidation reactions catalyzed by human XOR
Table 26 Examples of natural and physiological compounds as inhibitors of human XOR enzymes
Table 27 Examples of general chemical as substrates in reactions catalyzed by human XOR enzymes
Table 28 Examples of general chemicals as inhibitors of human XOR enzymes
Table 29 Compounds activated to toxic/pharmacologically active metabolites by human XOR

The functional XOR and AOX enzymes are homodimers composed of two identical subunits of approximately 150 kDa, each possessing three cofactor-binding domains connected by flexible linker regions (Terao et al. 2020). The N-terminal domain contains two distinct iron–sulfur (2Fe–2S) redox centers, the central domain binds FAD, and the C-terminal domain houses a molybdenum cofactor (Moco) within the active site. The molybdenum atom of the Moco is coordinated with a sulfido ligand that is essential for catalytic activity. Whereas AOX exists only in a single form, mammalian XOR can interconvert between a dehydrogenase (XDH) and an oxidase (XO) (Battelli et al. 1973; Corte and Stirpe 1972; Della Corte and Stirpe 1968; Stirpe and Della Corte 1969). Accordingly, AOX and the XO utilize molecular oxygen as a final electron acceptor, whereas only XDH can transfer electrons to NAD+. With amino acid sequence identities of approximately 50%, AOX and XOR enzymes possess similarities in substrate specificity (e.g., aromatic azaheterocycles); however, the larger, more anionic active site of AOX is able to accommodate a wider range of substrates relative to XOR (Mahro et al. 2013).

The tissue distribution of both AOX and XOR is species-dependent. AOX expression in humans is distributed across many different tissues, including liver (major), kidneys, lungs, gastrointestinal tract, skin, male reproductive tissues, and endocrine tissues, most notably the adrenal glands (Moriwaki et al. 2001; Terao et al. 2016). Constitutive expression of XOR in human tissues is low, and consequently, XOR activity is primarily present in the liver and gastrointestinal tract, as well as in lactating breast and kidney (Battelli et al. 2016a; Bortolotti et al. 2021). Notably, XOR is located in the vascular endothelium and can also be released into the systemic circulation, e.g., as a consequence of hepatic or intestinal damage (Kumar et al. 2018; Pritsos 2000).

The exact physiological roles of AOX and XOR are not well-defined, particularly with regard to AOX. Of the four human molybdenum hydroxylases, only sulfite oxidase is an essential enzyme (Duran et al. 1978; Shih et al. 1977; Terao et al. 2020; Veldman et al. 2010). XOR is responsible for the conversion of hypoxanthine to xanthine and of xanthine to uric acid (Balis 1976; Krenitsky et al. 1986). Consequently, XOR deficiency leads to the accumulation of xanthine, a condition referred to as xanthinurea (Kumar et al. 2018). Xanthinurea is an autosomal recessive disorder and is categorized as either Type I (Nakamura et al. 2012), which is associated with a deficiency in XOR alone, or Type II, which is associated with a deficiency of both XOR and AOX (Reiter et al. 1990). In addition, XOR is capable of reducing nitrates to nitrites and both AOX and XOR have been shown to reduce nitrites to nitric oxide (Maia and Moura 2018; Maia et al. 2015). Because AOX and XO utilize molecular oxygen as an electron acceptor, both enzymes produce reactive oxygen species (hydrogen peroxide and/or superoxide) as by products in catalyzing the oxidation of substrates. Oxidative damage has been linked to the development of cancer (Oberley 2002), and both AOX and XOR have been implicated in tumor growth and development (Kusano et al. 2019; Qiao et al. 2020; Takeuchi et al. 2018).

Reactions

Molybdenum hydroxylases catalyze the transfer of an oxygen atom, ultimately derived from water, to a substrate in a two-electron redox reaction (Fig. 28) (Kisker et al. 1997). The enzymes oxidize carbon atoms of a number of different aldehyde and heteroaromatic rings. In general, aromatic azaheterocyclic compounds are better substrates of molybdenum-hydroxylases than aldehydes.

Fig. 28
figure 28

General reaction catalyzed by molybdenum hydroxylases

The catalytic mechanism of molybdenum hydroxylases used to oxidize aromatic azaheterocycles and aldehydes involves the oxidation of an electrophilic carbon, typically located adjacent to a nitrogen in heterocyclic substrates (Alfaro and Jones 2008). The process begins with deprotonation of a hydroxyl group on the Moco by a conserved glutamate residue, followed by a nucleophilic attack on the electron deficient carbon atom of the heteroaromatic substrate. Hydride transfer from the electrophilic carbon of the substrate to the sulfur of the Moco then follows, resulting in a reduction of the molybdenum from Mo(VI) to Mo(IV). While the reaction could proceed via a tetrahedral intermediate in a step-wise mechanism, it is believed to proceed via a concerted mechanism (Fig. 29). The reaction intermediate is hydrolyzed, releasing the oxidized product, and a water molecule replaces the lost hydroxyl ligand on the molybdenum. The reducing equivalents are shuttled from the Moco to FAD via the iron-sulfur clusters. FADH2 is then reoxidized by molecular oxygen via a one or two-electron transfer, generating superoxide ion or H2O2, respectively. The oxidized products are structurally similar to those generated by P450 enzymes. However, the oxygen molecule used to oxidize substrates of molybdenum hydroxylases is derived from water (Garattini and Terao 2012), unlike P450s which use molecular oxygen as the source of the oxygen in the product (Guengerich 2001). Accordingly, the inclusion of H218O in incubations with molybdenum hydroxylases is utilized as a reaction phenotyping strategy for these enzymes.

Fig. 29
figure 29

Catalytic cycle and proposed mechanism for the oxidation of aromatic heterocycles and aldehydes by the molybdenum hydroxylases using quinazoline as an example (glutamate numbering represents human AOX) (Alfaro and Jones 2008)

In addition to oxidation reactions, AOX and XOR are capable of catalyzing reduction reactions. Both AOX and XOR have been demonstrated to reduce nitrite to nitric oxide (Fig. 30), an important signaling molecule involved in numerous physiological functions, including vasodilation, platelet aggregation, and immune response (Godber et al. 2000; Li et al. 2009; Maia et al. 2015). AOX is also known to reduce a variety of other functional groups, including N- and S-oxides, heterocycles, and nitro groups (Amano et al. 2018; Cashman et al. 2020; Dalvie and Di 2019; Ogiso et al. 2018; Pryde et al. 2010; Sung et al. 2020).

Fig. 30
figure 30

General reaction of nitrite reduction catalyzed by molybdenum hydroxylases

While reductive reactions and mechanisms have received less attention relative to oxidation reactions catalyzed by molybdenum hydroxylases, Maia and Moura have described the mechanism of nitrite reduction to nitric oxide (Fig. 31), which takes place at the Moco center (Maia and Moura 2018). A reducing substrate, such as an aldehyde or aromatic heterocycle, is required to reduce the Moco from Mo(VI) to Mo(IV) as previously described in Fig. 29. The nitrite reduction then proceeds via sequential one electron transfer to two molecules of nitrite, reoxidizing the Moco from Mo(IV) to Mo(V) and then back to Mo(VI). Maia and Moura also demonstrated that the reaction is independent of the FAD center with experiments using an FAD inhibitor or enzyme lacking FAD.

Fig. 31
figure 31

Proposed mechanism for reduction of nitrite to nitric oxide in the presence of a reducing substrate by the molybdenum hydroxylases (Maia and Moura 2018)

Aldehyde oxidase 1 (AOX1)

Mammalian aldehyde oxidases (AOXs) are cytosolic molybdoflavoenzymes involved in the metabolism of drugs, natural and physiological compounds, and general chemicals (Tables 15, 17, 19). The enzymes participate not only in the detoxication of toxic metabolites endogenously formed by other enzymes such as P450s (e.g., aldehyde intermediates) but also in the production of toxic and therapeutically active metabolites (Tables 21, 22), and the generation of reactive oxygen species (ROS) as a byproduct of their enzymatic activity.

Enzymes

Different animal species are characterized by a different complement of aldehyde oxidase genes clustering at a short distance on the same chromosome (chromosome 2 in humans) (Terao et al. 2016). Humans contain a single active gene, AOX1, and two pseudogenes, while rodents are characterized by four active genes. Both AOX1 and AOX3 are major enzymes present in rodent liver (with the exception of guinea pigs, which only express AOX1 in the liver). The mouse Aox1 enzyme bears 85% sequence identity with human AOX1, whereas mouse Aox3 is only 65% identical to human AOX1 (Garattini et al. 2008). Primates, like humans, have only a single functional AOX enzyme (AOX1) in the liver, which bears 96% sequence identity with the human enzyme (Hoshino et al. 2007). Accordingly, marked species differences in AOX-mediated metabolism are common, and these differences present in a substrate-dependent manner (Beedham et al. 1987; Choughule et al. 2015; Crouch et al. 2018; Dalvie et al. 2013; Diamond et al. 2010; Hutzler et al. 2014; Sahi et al. 2008). However, AOX catalytic activity generally tends to be highest in monkeys and humans and lowest in mice and rats, whereas rabbits and guinea pigs tend to fall somewhere in between. Dog liver is completely devoid of an active AOX enzyme (Terao et al. 2016).

The human AOX1 protein has been reported in many tissues, including liver, pancreas, kidney, adrenal gland, thyroid gland, prostate, bladder, gastrointestinal tract, testis, bronchi, uterus, and skin (Moriwaki et al. 2001). The liver contains the highest concentration of AOX1 protein, though substantial quantities are also present in the adrenal glands. AOX1 mRNA expression has been found in many human tissues as well (Terao et al. 2016).

Humans have functionally inactive AOX1 allelic variants as well as variants encoding enzymes with different catalytic activities (i.e., slow and rapid metabolizers) (Foti et al. 2016; Hartmann et al. 2016; Mota et al. 2019). In addition, single nucleotide polymorphisms affecting the FAD binding site have been demonstrated to increase the rate of superoxide production (Foti et al. 2017). The clinical relevance of these variants has yet to be established. Garrido and Leimkühler demonstrated that the L438V variant, which produces a higher ratio of superoxide/H2O2 relative to the wild-type enzyme, is more extensively inactivated over time (i.e. inactivated by ROS production) relative to the wild type enzyme (Garrido and Leimkühler 2021). The L438V variant, which bears a single nucleotide polymorphism affecting the FAD binding site, produces superoxide at a rate of 75% compared to the amount of H2O2 produced, whereas the wild-type enzyme produces only 10% superoxide in comparison to H2O2.

As the name suggests, AOX catalyzes not only the aldehydes but can also catalyze the oxidation of aromatic azaheterocycles, as well as reductive reactions. AOX is best understood for its role in xenobiotic metabolism, but the enzyme may have additional physiological functions. Species differences in the expression and activity of AOX present a challenge in defining the physiological role(s) of human AOX. The list of endogenous substrates of AOX includes N-methylnicotinamide, pyridoxal, and all-trans-retinaldehyde (Table 17) (Johns 1967; Zhong et al. 2021). AOX has also been shown to be capable of reducing nitrites to nitric oxide and has been proposed to be involved in adipogenesis (Heid et al. 2020; Maia et al. 2015; Weigert et al. 2008). Due to its function in producing ROS, AOX may also contribute to pathological conditions resulting from oxidative stress (e.g., cancer). However, the exact physiological/pathological roles of AOX remain poorly understood.

Substrates

AOX is characterized by broad substrate specificity, in contrast to XOR, which has a specificity more limited to purine-like compounds. AOX has been most frequently reported to oxidize aromatic azaheterocycles, e.g. substituted pyrroles, pyridines, pyrimidines, purines, pteridines, and quinolines, among others (Dalvie and Di 2019; Garattini and Terao 2012; Kitamura et al. 2006; Manevski et al. 2019; Pryde et al. 2010). In addition, compounds containing iminium ions (often intermediate metabolites generated by P450s or MAOs) are relatively common AOX substrates. Aliphatic and aromatic aldehydes (which also often arise as intermediate metabolites) are oxidized by AOX as well. However, compounds containing aldehydes tend to be more efficiently oxidized by ALDH. AOX substrates for reduction include nitro-containing compounds, sulfoxides, N-oxides, and nitrites, as well as heterocycles such as isoxazoles and isothiazoles. More recently, AOX has also been demonstrated to be capable of hydrolyzing amides (Sodhi et al. 2015). While AOX and P450s have somewhat opposing substrate preferences due to different catalytic mechanisms (e.g. P450s prefer to oxidize electron-rich carbon atoms, whereas AOXs prefer electron deficient carbon atoms), they can sometimes share heterocyclic substrates and produce the same hydroxylated products (e.g., idelalisib) (Jin et al. 2015).

Inhibitors

A number of drugs, natural compounds, and general chemicals have been reported to inhibit human AOX (Tables 16, 18, 20). The strongest known inhibitor of AOX is the selective estrogen receptor modulator, raloxifene (IC50 2.9 nM), while several other drugs have also been also reported to strongly inhibit the enzyme, particularly phenothiazines (e.g. perphenazine, IC50 33 nM) (Obach et al. 2004). 17 β-Estradiol and 17α-ethinylestradiol are also strong inhibitors. In addition to a number of general chemical compounds (e.g., phenothiazines, dibenzazepines, flavonoids, purines, and pyrimidines) (Table 20), a number of diet-derived natural products have been found to inhibit AOX activity, including various catechins, flavones, flavonoids, flavonols, and flavanonols (Table 18). The compounds most commonly used to phenotype AOX reactions are raloxifene, menadione, and the time-dependent inhibitor hydralazine. Each of these three compounds is selective for AOX over XOR, but they exhibit varying degrees of inhibition toward various microsomal enzymes (Zientek and Youdim 2015). As the inhibition kinetics of AOX is complex (and substrate-dependent), the use of multiple probe substrates is recommended when results are used to assess the potential for drug–drug and/or drug–chemical interactions (Barr and Jones 2013). Clinically significant drug–drug interactions involving AOX inhibition have not been reported, with the exception of zaleplon and cimetidine. Zaleplon inhibits not only the AOX-mediated metabolism (of zaleplon) but also the CYP3A-mediated pathway, which represent approximately 70% and 30% of the fractional metabolism of zaleplon, respectively (Renwick et al. 2002).

Inducers

Limited reports are available on the regulation of AOX expression, particularly regarding the human enzyme. Maeda et al. demonstrated regulation of the human AOX gene involving the Nrf2 pathway (Maeda et al. 2012). In addition, Zhou et al. reported increased AOX-mediated metabolism of methotrexate, increased AOX protein levels, and a minimal increase in AOX mRNA following treatment of human hepatocytes with the nonsteroidal anti-inflammatory drug nimesulide (Zhou et al. 2020). Others have demonstrated induction of AOX activity, mRNA, and/or protein in animals by various compounds (e.g. phthalazine, dioxin) (Johnson et al. 1984; Rivera et al. 2005). Notably, androgens have been reported to increase AOX expression in rodents, while estrogens reduced expression, which is consistent with sex-dependent AOX activity observed in rodents (Al-Salmy 2001; Beedham 1985; Garattini and Terao 2012). Humans, however, appear to exhibit no sex-dependent differences in AOX activity, although estrogens are known to inhibit human AOX in vitro (Obach 2004).

Reactions

AOX catalyzes the oxidation of aromatic and aliphatic aldehydes into the corresponding carboxylic acids, hydroxylation of electrophilic carbon atoms on heteroaromatic rings, oxidation of iminium ion intermediates to the corresponding lactams, and hydrolysis of amides. AOX is also reported to catalyze the reductive metabolism of nitro groups, N-oxides, sulfoxides, isoxazoles, isothiazoles, benzisoxazoles, nitrite, and hydroxamic acids. Certain reductive transformations (e.g., nitro-reduction) have the potential to cause toxicity due to the formation of reactive metabolites. Reductive reactions may be accelerated in the presence of a reducing substrate, as was demonstrated for the reduction of dantrolene in the presence of N-methylnicotinamide (Amano et al. 2018).

AOX catalyzes the oxidation of heteroaromatic rings, iminium ions, and aldehydes. During the oxidation of the substrate, the enzyme is reduced and reoxidized with molecular oxygen, therefore behaving as an oxidase (Fig. 32). For both AOX and the XO form of XOR, the reduction of molecular oxygen produces the reactive oxygen species (ROS), hydrogen peroxide and superoxide anion, with AOX favoring the production of hydrogen peroxide (Foti et al. 2017). However, it has been suggested that AOX may produce more than 20-fold higher amounts of superoxide versus XOR, based on the enzymatic activities of the two enzymes in the human liver (Krenitsky et al. 1972; Kundu et al. 2007). Although NAD+ is the preferred electron acceptor for the XDH form of the XOR enzyme, it is capable of transferring electrons to O2 and thus producing ROS by acting as an NADH oxidase (Sanders et al. 1997).

Fig. 32
figure 32

Typical reaction catalyzed by AOX and XO enzymes

Garrido and Leimkühler recently reported that AOX is inactivated in a substrate-dependent manner by ROS production, with a high turnover substrate inactivating the enzyme more rapidly than a low turnover substrate (Garrido and Leimkühler 2021). Alternatively, because the enzyme inactivation was not prevented by ROS scavengers (catalase and superoxide dismutase) in their studies, Abbasi et al. reported that the enzyme inactivation results, not from ROS production, but rather from the rate-limiting transfer of electrons to O2, which is required to reoxidize the enzyme (Abbasi et al. 2019). Similar to the results obtained by Garrido and Leimkühler with AOX, Lynch and Fridovich previously reported that XO is autoinactivated by ROS production (Lynch and Fridovich 1979). Differences in incubation conditions were cited as a possible explanation for the discrepancy between the findings in the two AOX studies.

Other compounds can also serve as electron acceptors for AOX as well. While some accept electrons at the FAD site (e.g. 5-nitroquinoline) (Abbasi et al. 2019), where O2 accepts electrons, others may directly receive electrons from the Moco site (e.g. 2,6-dichlorophenolindophenol) (Foti et al. 2017; Garrido and Leimkühler 2021).

Examples of substrates and reactions resulting in the formation of nontoxic metabolites

Oxidation of aromatic heterocycles

Phthalazine (general chemical)

Phthalazine is a bicyclic heteroaromatic compound that is rapidly oxidized by AOX to phthalazone (Beedham et al. 1987, 1995) (Fig. 33). Phthalazine is commonly used as an AOX probe substrate (Table 19). The time-dependent AOX inhibitor hydralazine is a derivative of phthalazine. Because phthalazine and phthalazone are metabolites of hydralazine, it is not a suitable inhibitor to evaluate the AOX-mediated metabolism of phthalazine.

Fig. 33
figure 33

Oxidation of phthalazine by AOX

DACA (antineoplastic, DNA intercalating dual topoisomerase I/II poison)

While most heterocyclic substrates of AOX are oxidized on an electrophilic carbon atom adjacent to a nitrogen atom, in some cases the oxidation may occur para to the nitrogen, as is the case for the antineoplastic agent DACA (Fig. 34). DACA is oxidized to an acridone product by AOX on the carbon atom opposite the nitrogen in the acridine moiety (Schofield et al. 2000) (Table 15).

Fig. 34
figure 34

Oxidation of DACA by AOX

Lapatinib (antineoplastic, EGFR inhibitor)

Lapatinib, an anticancer agent known to be associated with hepatotoxicity, is debenzylated by P450 3A enzymes to a metabolite that subsequently undergoes metabolic activation via P450 3A enzymes (Castellino et al. 2012) (Fig. 35). Lapatinib and its debenzylated metabolite are also oxidized by AOX to AO-M1 and M3, respectively (Dick 2018), which may serve as a detoxication pathway in opposition to the bioactivation pathway (Table 15).

Fig. 35
figure 35

Bioactivation of lapatinib by P450 3A enzymes and oxidation by AOX

Zaleplon (nonbenzodiazepine sedative hypnotic)

Zaleplon is a dual AOX and P450 substrate, undergoing oxidation by AOX to 5-oxo-zaleplon (Table 15, Fig. 36) and N-dealkylation by P450 3A4 to desethyl-zaleplon. Species differences are noted in the fractional metabolism by AOX and P450, with 5-oxo-zaleplon representing the major metabolite in humans and monkeys (approximate fm,AO (fraction of metabolism due to AOX)) = 0.7 in humans) and desethyl-zaleplon representing the major metabolite in rodents (Crouch et al. 2018; Kawashima et al. 1999; Strelevitz et al. 2012). In addition, an interaction between cimetidine and zaleplon is one of the only known clinically relevant drug–drug interactions associated with AOX metabolism (Renwick et al. 2002). However, cimetidine inhibits not only the AOX metabolism pathway but also the P450 3A4 pathway as well.

Fig. 36
figure 36

Oxidation and N-deethylation of zaleplon by AOX and P450 3A enzymes

Idelalisib (antineoplastic, phosphatidylinositol 3-kinase inhibitor)

Despite differences in catalytic mechanism and general substrate preference, some metabolites may be produced by both AOX and P450 enzymes, as is the case for idelalisib. Both AOX and P450 3A4 catalyze the oxidation of idelalisib to the inactive metabolite GS-563117, which is a mechanism-based inactivator of P450 3A (Jin et al. 2015; Ramanathan et al. 2016) (Table 15) (Fig. 37).

Fig. 37
figure 37

Oxidation of idelalisib by AOX and P450 3A enzymes

Oxidation of aldehydes

All-trans-retinaldehyde (vitamin A derivative)

Retinal is the most well-studied endogenous substrate of AOX (Ambroziak et al. 1999; Zhong et al. 2021). The aldehyde undergoes oxidation to produce the active metabolite retinoic acid, a reaction that is catalyzed by both AOX and ALDH (Table 17) (Fig. 38). Zhong and coworkers recently reported that ALDH1A1 serves as the low Km, low kcat enzyme contributing to the biosynthesis of retinoic acid in the human liver, while AOX serves as the high Km, high kcat enzyme (Zhong et al. 2021).

Fig. 38
figure 38

Oxidation of retinaldehyde to retinoic acid by AOX

Citalopram aldehyde (metabolite of SSRI citalopram)

Aldehydes are relatively uncommon in parent drug molecules. However, they are often generated by enzymes such as P450s. Citalopram is demethylated and converted to an aldehyde metabolite by a combination of P450 and MAO enzymes (Rochat et al. 1998) (Table 8) (Fig. 39). The aldehyde metabolite undergoes subsequent oxidation by AOX to the carboxylic acid (Table 15).

Fig. 39
figure 39

Oxidation of citalopram by P450, MAO, and AOX enzymes to citalopram propionic acid

Oxidation of iminium ions

Like aldehyde intermediates, AOX has been found to oxidize iminium ion intermediates generated by P450s or MAOs. As iminium ions have the potential to produce toxic effects, AOX-mediated oxidation of iminium ion intermediates may serve as a detoxication pathway.

KW-2449 (multikinase inhibitor)

KW-2449 is a multikinase inhibitor that was previously under investigation for the treatment of leukemia. The drug, which has been discontinued from further development, displayed unexpected rapid metabolism to an oxo-piperazine metabolite in clinical trials (Hosogi et al. 2018). The pharmacologically active metabolite was determined to be generated via sequential metabolism by MAO B to an iminium ion intermediate, followed by AOX-mediated oxidation to the oxo-piperazine metabolite (Table 15, Fig. 40). In addition, the iminium ion intermediate was found to be a time-dependent inhibitor of AO, reducing exposure of the active metabolite following repeat dosing (Table 16). Interspecies differences in both MAO B and AOX likely contributed to the failure to recognize this metabolic pathway prior to clinical trials.

Fig. 40
figure 40

Oxidation of KW-2449 to an oxo-piperazine metabolite by MAO B and AOX

Reduction of N-oxides

Oxycodone N-oxide (metabolite of the opioid analgesic oxycodone)

Oxycodone is extensively metabolized by P450 enzymes, but it is also converted to oxycodone N-oxide by FMO3 (Cashman et al. 2020). The N-oxide metabolite was found to be retro-reduced back to oxycodone by AOX, quinone reductase, and hemoglobin (Table 15) (Fig. 41). Consequently, interindividual variability in AOX activity could potentially influence the duration of action and toxicity of oxycodone across patients.

Fig. 41
figure 41

N-Oxidation of oxycodone by FMO and retro-reduction by AOX and other enzymes

Reduction of S-oxides

Sulindac (NSAID)

Fewer examples of AOX-mediated S-oxidation are available in the literature relative to N-oxidation. The reduction of the nonsteroidal anti-inflammatory drug sulindac to its pharmacologically active sulfide metabolite represents one example of this reaction (Sung et al. 2020) (Table 15). Alternatively, the sulfide can be oxidized back to sulindac by FMO enzymes, which also oxidize sulindac to an inactive sulfone metabolite (Table 3, Fig. 42).

Fig. 42
figure 42

S-Oxidation of sulindac by FMO and sulfoxide reduction by AOX

Reduction of heterocycles

Ziprasidone (second generation antipsychotic)

Ziprasidone represents an example of a drug containing a heterocycle that is reduced by AOX (Miao et al. 2005; Prakash et al. 1997). The drug is extensively metabolized to multiple metabolites, including several oxidative P450 metabolites. However, the major circulating metabolite results from AOX-mediated reductive cleavage of the benzisothiazole ring, followed by methylation of the thiol (Table 15, Fig. 43).

Fig. 43
figure 43

Benzisothiazole reduction and thiol methylation of ziprasidone by AOX and thiol methyltransferase

Toxic effects of drugs as substrates of AO catalyzed reactions

Several examples of therapeutically successful candidate drugs tested in animal models have been removed from further testing due to differences in the formation of a toxic metabolite in preclinical species relative to humans (e.g., SGX523) (Hutzler et al. 2013; Manevski et al. 2019). Some examples of drugs converted into toxic metabolites by AOX are listed in Table 22. Alternatively, examples of drugs that are converted to active metabolites by AOX are listed in Table 21, which includes metabolites possessing desirable cytotoxic properties (e.g., anti-cancer agents).

Reduction of nitro-groups

Drugs containing nitro groups have been associated with mutagenicity and genotoxicity. AOX is capable of reducing nitro-groups to their corresponding amines, producing a hydroxylamine intermediate in the process. As hydroxylamines are reactive species that have the potential to produce toxic effects, the toxicities (e.g., hepatotoxicity) associated with nitro-aromatic containing drugs such as nimesulide, nitrazepam, dantrolene, and others (Table 22) may result, at least in part, from AOX-mediated nitro-reduction of these drugs.

Nitrazepam (benzodiazepine)

Koneshi et al. demonstrated that AOX participates in the reduction of nitrazepam to hydroxylaminonitrazepam and aminonitrazepam (Konishi et al. 2017). The aminonitrazepam metabolite is further metabolized by N-acetyltransferases (NATs) to N-acetylaminonitrazepam (Table 15, Fig. 44).

Fig. 44
figure 44

Nitro reduction and N-acetylation of nitrazepam by AOX and NAT

Dantrolene (skeletal muscle relaxant)

Dantrolene, like nitrazepam, contains a nitro-group that is reduced to a hydroxylamine, followed by a second reduction to aminodantrolene (Amano et al. 2018; Ogiso et al. 2018) (Table 22). Aminodantrolene is also N-acetylated by NAT2 (Fig. 45). Dantrolene carries a black box warning for severe hepatotoxicity, which is attributed to the formation of the hydroxylamine intermediate.

Fig. 45
figure 45

Nitro reduction and N-acetylation of dantrolene by AOX and NAT

Oxidation of heterocycles to poorly soluble lactam metabolites

SGX523 (antineoplastic, c-MET inhibitor)

SGX-523 is an anticancer agent that was discontinued in clinical trials due to renal toxicity that went undetected in toxicity studies conducted in rats and dogs (Infante et al. 2013). Diamond et al. determined that the toxicity likely resulted from precipitation of an AOX metabolite in the renal tubules, which was not observed in animals studies due to species differences in AO activity (Diamond et al. 2010) (Tables 15, 22). The 2-quinolinone (Fig. 46) metabolite was undetected in dog liver post-mitochondrial supernatant fraction (S9) and only trace amounts were produced in rat S9, unlike human and monkey S9, in which the 2-quinolinone was a major metabolite. Following administration of SGX-523 to cynomologus monkeys, the 2-quinolinone metabolite was present in urine at concentrations 70-fold higher than the parent drug and its solubility was only 3% of the parent solubility. A structural analog of SGX-523, JNJ-38877605, was also discontinued during clinical trials due to renal toxicity, which was presumed to occur via the same AOX-mediated mechanism (Lolkema et al. 2015). JNJ-38877605 has a difluoro-substituted ether linkage rather than a thioether linkage and is otherwise structurally identical to SGX-523.

Fig. 46
figure 46

Oxidation of SGX-523 to a poorly soluble lactam metabolite by AOX

Methotrexate (antineoplastic, antirheumatic, antifolate)

Methotrexate is also known to cause renal toxicity, particularly when administered in high doses (Jordan et al. 1999). As with SGX-523 and JNJ-38877605, renal toxicity is believed to be associated with AOX-mediated oxidation to 7-hydroxymethotrexate, a poorly soluble metabolite (Table 22, Fig. 47).

Fig. 47
figure 47

Oxidation of methotrexate to a poorly soluble lactam metabolite by AOX

Hydrolysis of anilides

GDC-0834 (antirheumatic, Bruton’s tyrosine kinase inhibitor)

GDC-0834 was previously under investigation for the treatment of rheumatoid arthritis but was discontinued from further development due to rapid hydrolysis of the anilide moiety, producing an aniline metabolite (Liu et al. 2011; Sodhi et al. 2015) (Fig. 48). While neither preclinical nor clinical toxicity was reported as a concern or reason for discontinuation of GDC-0834, anilines are known to have the potential to produce toxic effects, if for no reason than guilt by association. Lepri et al. evaluated a series of anilide-containing compounds for their susceptibility to AOX-mediated amide hydrolysis and found several of the compounds to be AOX substrates, thus highlighting this reaction as a potential source of toxicity for anilide-containing drugs (Lepri et al. 2017) (Table 15).

Fig. 48
figure 48

Hydrolysis of GDC-0834 to an aniline metabolite by AOX

Sodhi et al. have proposed a mechanism for the hydrolysis of GDC-0834, on the basis of in silico modeling, that involves a nucleophilic reaction between a hydroxyl group of the Moco and the carbonyl group of the anilide (Sodhi et al. 2015). The authors speculated that this reaction would be more likely to take place with the Moco in the reduced state (Fig. 49) due to the higher electron density, which would require the presence of a reducing substrate. Once the enzyme has initially been reduced, the proposed hydrolysis mechanism does not require any transfer of electrons in order to complete the catalytic cycle, meaning that the entire process would take place at the Moco center without the involvement of the FAD and 2Fe-2S centers.

Fig. 49
figure 49

Proposed mechanism of hydrolysis of GDC-0834 by AOX

Xanthine oxidoreductase (XOR)

Xanthine oxidoreductase (XOR), perhaps the better-known molybdenum hydroxylase, plays an important role in the catabolism of endogenous purines and pyrimidines in humans, as well as drugs such as thiopurines and methyl-xanthine compounds (Tables 23, 25, 27). XOR is less promiscuous than AOX, preferring substrates that are more purine-like, although some compounds are substrates for both enzymes.

Enzymes

Unlike AOX, XOR can exist in two interconvertible forms, as a dehydrogenase (XDH), which prefers NAD+ as an electron acceptor, or an oxidase (XO), which can only transfer electrons to O2. Based upon studies with rat liver and bovine milk XOR, the XDH form can be postranslationally modified to the XO form either irreversibly through limited proteolysis, or reversibly through the formation of two disulfide bonds involving four cysteine residues (Battelli et al. 1973; Corte and Stirpe 1972; Della Corte and Stirpe 1968; Stirpe and Della Corte 1969). In both cases, either reversible or irreversible conversion from XDH to XO, the modification takes place within a peptide that links the Moco- and FAD-containing domains.

Tissue distribution of XOR is species-dependent, with lower constitutive expression in humans relative to other mammals, presumably due to promotor suppression (Xu et al. 2000). Human XOR enzymes have been found in the lactating mammary gland, intestine, liver, lungs, kidneys, and vascular endothelium, with the highest specific activity in the liver and intestine (Linder et al. 1999; Moriwaki et al. 1996). In addition, relative levels of an inactive enzyme (for example, de-molybdo and/or de-sulfo forms of the enzyme) may contribute to species differences in the tissue distribution of XOR activity (Battelli et al. 2014). For example, in human milk, active XOR was found to account for < 2% of the total enzyme content, and the xanthine oxidizing activity in human milk was found to be 2–3 orders of magnitude lower than bovine milk, despite similar total XOR enzyme content (Abadeh et al. 1992).

XOR has been detected in the vascular endothelial cells of various human tissues (Kooij et al. 1992; Linder et al. 1999; Moriwaki et al. 1993). While XOR is a cytosolic enzyme like AOX, it has also been detected on the outer surface of bovine and porcine endothelial cells (Vickers et al. 1998). XOR is also present (in the XO form) in circulating plasma, although the constitutive presence in plasma is species-dependent (Al-Khalidi and Chaglassian 1965). For example, studies evaluating the plasma stability of a quinoxaline-containing compound within a series of cyanopyridine derivatives revealed a species-specific oxidation of the quinaoxaline mediated by plasma XOR (Sharma et al. 2011). The quinoxaline-containing compound was rapidly degraded in rat, mouse, and guinea pig plasma but not in dog, monkey, or human plasma. The lactam metabolite was also detected in human liver cytosol; however, a further evaluation to distinguish whether the metabolite was produced by human hepatic XOR vs AOX was not reported. Though plasma levels of XOR in healthy humans is low, increased levels of circulating plasma XOR have been associated with various pathological conditions, primarily related to hepatic injury, including viral hepatitis, toxic agents, transplantation, hypoxia, and ischemia/reperfusion (Battelli et al. 2014). Other conditions that have been associated with elevated circulating XOR include pneumonia, type 2 diabetes, post-surgical procedures, and sickle cell disease, among others (Battelli et al. 2014).

Products of XOR catalyzed reactions have been associated with both beneficial and toxic effects and elevated XOR activity has been connected to different pathological conditions causing tissue damage and cell necrosis (Battelli et al. 2016b; Bortolotti et al. 2021; Harrison 2002). Uric acid, the product of xanthine oxidation, and NO, the product of nitrite reduction, play a role in blood pressure regulation and vascular tone. In addition, uric acid has anti-oxidant activity, contributes to the inflammatory response, and promotes gluconeogenesis and fat accumulation. Consequently, XOR may play a role in the pathogenesis of metabolic syndrome and insulin resistance. While ROS produced by XOR can contribute to pathological conditions associated with oxidative stress such as cancer, uric acid plays a role in preventing these pathological conditions. Beneficial effects are also derived from ROS production, for example, via their bactericidal action.

Substrates

Similar to AOX, XOR utilizes a variety of heterocycles and some aldehydes as substrates. However, XOR has a narrower specificity than AOX, generally preferring substrates that are more purine-like (Tables 23, 25, 27). Oxidation of purines occurs at C-atoms in positions C2-, C6-, and C8- (Okamoto et al. 2013) (Fig. 50). The affinity of C-atoms for oxidation increases with the number of adjacent N-atoms.

Fig. 50
figure 50

Oxygenation of purine compounds by XOR enzymes

The substrate specificity for AOX and XOR does sometimes overlap. For example, 6-mercaptopurine is oxidized by both AOX and XOR to 6-thioxanthine, whereas oxidation of 6-thioxanthine to 6-thiouric acid is catalyzed by XOR alone (Choughule et al. 2014). In some cases, a compound may be a dual substrate for XOR and AOX, but the site of metabolism may differ. Examples of this observation in the literature, however, have been determined using non-human sources of XOR and/or AOX. For example, 6-deoxyacyclovir was reported to be oxidized to the active metabolite acyclovir by bovine milk XOR and to the inactive metabolite 8-hydroxy-6-deoxyacyclovir by rabbit liver AOX (Krenitsky et al. 1984). In fact, many compounds not reported in this review are substrates for mammalian XOR or AOX but have not been confirmed to be substrates of the human enzymes. XOR studies in particular have been frequently carried out using a non-human enzyme source, most often bovine milk XOR owing to its wide availability and low cost. Due to species differences in AOX/XOR substrate specificities, it cannot be assumed that a substrate for a non-human enzyme is also a substrate for the human enzyme. For example, using enzyme-selective inhibitors in multiple species of S9 preparations, the 6-oxopyrimidine metabolite of VU424238, for example, was found to be oxidized to a 2,6-oxopyrimidine metabolite by either AOX or XOR in a species-dependent manner (Crouch et al. 2017).

Inhibitors

As previously mentioned, XOR is responsible for the conversion of hypoxanthine to xanthine and xanthine to uric acid, the accumulation of which is associated with the pathophysiology of gout. Consequently, inhibition of XOR is a therapeutic strategy in the treatment of gout, and multiple drugs that inhibit XOR are on the market (allopurinol, febuxostat, and topiroxostat). Allopurinol has been in use since the 1960s, whereas febuxostat only received FDA approval in 2009. Topiroxostat is not available in the United States but was approved for use in Japan in 2013. Allopurinol is a substrate of both XOR and AOX and is converted to the active metabolite oxypurinol, which forms a covalent bond with Mo(IV) and strongly inhibits XOR (Okamoto et al. 2008). Febuxostat is characterized as a structure-based inhibitor and can bind to the enzyme regardless of the Moco oxidation state (Okamoto et al. 2003). Topiroxostat initially displays competitive inhibition, followed by a covalent type of inhibition, based on studies with bovine milk XOR (Matsumoto et al. 2011). Several additional purine and nonpurine-like compounds have been found to inhibit XOR but are not utilized as clinical XOR inhibitors (Tables 24, 26, 28).

Inducers

Constitutive activity of XOR in human tissues is relatively low in comparison with other mammals (Harrison 2002). Increased activity and/or expression of XOR in various tissues has been associated with several compounds, including interferon, 3-methylcholanthrene, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), prolactin, and cortisol in mice (Ghezzi et al. 1984; McManaman et al. 2000; Sugihara et al. 2001), and sodium metabisulfite and phenytoin in rats (Ercan et al. 2015; Ekaidem et al. 2014). TCDD and 3-methylcholanthrene also increase AOX activity in mouse liver, and these effects were diminished in aryl hydrocarbon receptor-null mice (Sugihara et al. 2001). Cytokines have also been shown to increase XOR activity on a transcriptional and/or post-translational level in rodent, bovine and human cultured cells of various tissues (Dupont et al. 1992; Page et al. 1998; Pfeffer et al. 1994). Menadione, an inhibitor of AOX, stimulates human XOR activity, presumably by functioning as an electron acceptor to reoxidize the enzyme following substrate oxidation (Ferreira Antunes et al. 2016).

Examples of substrates and reactions resulting in the formation of nontoxic metabolites

Hypoxanthine and xanthine (endogenous purines)

XOR catalyzes the oxidation of hypoxanthine into xanthine and further oxidizes xanthine into uric acid (Fig. 51) (Table 22) (Okamoto et al. 2013). In the purine metabolism, hydroxylation of hypoxanthine (6-hydroxypurine) initially takes place at the 2-position, yielding xanthine (2,6-dihydroxypurine). The next hydroxylation occurs at the 8-position, affording uric acid (2,6,8-trihydroxypurine). Studies using bovine milk XOR indicated that xanthine accumulates prior to conversion into uric acid, suggesting that oxidation of the 2-position influences oxidation of the 8-position (Okamoto et al. 2013). In addition, 6,8-dihydroxypurine was not detected. Dimethylated (theophylline, theobromine) and trimethylated xanthine derivatives (caffeine) are better substrates of cytochrome P450 enzymes than XOR.

Fig. 51
figure 51

Oxidation of hypoxanthine to xanthine and uric acid by XOR

Allopurinol and oxypurinol (antigout, XOR inhibitor)

Allopurinol is a substrate and specific XOR inhibitor. An analog of hypoxanthine, allopurinol has a nitrogen atom in 8-position with a carbon atom in the 7-position. Both AOX and XOR can metabolize allopurinol to the active metabolite oxypurinol (Fig. 52) (Tables 15, 23), although this reaction is probably mediated primarily by AOX. Allopurinol is both a competitive (at lower concentrations), and uncompetitive inhibitor (at higher concentrations) of hypoxanthine and xanthine oxidations, catalyzed by XOR (Table 24). Oxypurinol is an uncompetitive inhibitor and covalently binds to the reduced form of XOR (Okamoto et al. 2008; Spector 1988; Spector et al. 1986) (Table 24). Allopurinol/oxypurinol inhibit the conversion of hypoxanthine and xanthine to uric acid, thus regulating blood urate levels and is used to treat gout and hyperuricemia. In addition, it was suggested that allopurinol, by suppression of XOR activity, ameliorates myocardial inefficiency and poor vascular flow, and accordingly, may present an innovative contribution to the future treatment of ischemia and reperfusion (I/R) injury in heart failure patients (Harzand et al. 2012; Lee et al. 2009a, b). Allopurinol is also associated with potentially life-threatening severe cutaneous adverse reactions (Table 29) for which the HLA-B*5801 allele has been identified as a genetic risk factor (Hung et al. 2005).

Fig. 52
figure 52

Oxidation of allopurinol to oxypurinol by AOX and XOR enzymes

When human liver tissue is harvested, it is commonly perfused with a solution containing allopurinol to prevent XOR-related oxidative damage. Barr et al. reported the presence of both allopurinol and oxypurinol at micromolar concentrations in cytosolic human liver fractions obtained from livers perfused with an allopurinol-containing solution, with a corresponding lack of XOR activity in these samples (Barr et al. 2014). Importantly, the authors noted that commercial liver fractions are likely to contain residual allopurinol and/or oxypurinol and should be screened prior to use in metabolism studies.

Acyclovir (prodrug, antiviral, antiherpetic)

6-Deoxyacyclovir is an example of a prodrug that is activated by the catalytic activity of XOR (Fig. 53). 6-Deoxyacyclovir is converted into the active drug acyclovir via 6-oxidation (Krenitsky et al. 1986; Rees et al. 1986) (Table 29). Rabbit liver AOX was found to oxidize both 6-deoxyacyclovir and acyclovir at the 8-position to inactive metabolites (Krenitsky et al. 1984), but whether or not this deactivating reaction is catalyzed by human AOX has not been reported.

Fig. 53
figure 53

Oxidation of 6-deoxyacyclovir to the active metabolite acyclovir by XOR

6-Mercaptopurine (antineoplastic)

6-Mercaptopurine, a thiopurine drug, can be oxidized to 6-thioxanthine (6TX) and 6-thiouric acid (6TUA) through 6TX as an intermediate (Fig. 54). Both AOX and XOR are found to be involved in the formation of the 6TX intermediate, whereas only XOR was responsible for the conversion of 6TX to 6TUA (Choughule et al. 2014) (Tables 15, 23). In addition, both the xanthine dehydrogenase (XDH) and xanthine oxidase (XO) forms of XOR were evaluated and found to contribute to the formation of 6TX and 6TUA in studies with human liver cytosol in the presence and absence of NAD+, the preferential electron acceptor for XDH.

Fig. 54
figure 54

Oxidation of 6-mercaptopurine to 6-thiouric acid by AOX, XO, and XDH

Toxic effects of drugs as substrates of XOR catalyzed reactions

Pyrazinamide (antituburculosis prodrug)

Pyrazinamide is a prodrug used to treat tuberculosis, but it is associated with dose-related hepatotoxicity. Pyrazinamide is converted into the active metabolite pyrazioic acid by amidases, and it can also be oxidized by XOR to 5-hydroxypyrazinamide (5-OH-PZN) (Fig. 55) (Lacroix et al. 1989; Yamamoto et al. 1987) (Table 23). Both metabolites can undergo further conversion to 5-hydroxypyrazinoic acid (5-OH-PA) via the action of XOR on pyrazinoic acid or amidases on 5-hydroxypyrazinamide. The 5-OH-PA metabolite is proposed to be primarily responsible for the hepatotoxicity associated with pyrazinamide, as inhibition of amidase activity decreased pyrazinamide-induced hepatotoxicity, but did not prevent pyrazinoic acid-induced hepatotoxicity in rats (Shih et al. 2013). These data were also supported by in vitro studies demonstrating increased toxicity of pyrazinoic acid and 5-OH-PA relative to pyrazinamide in HepG2 cells. In addition, greater hepatotoxicity was observed in tuberculosis patients receiving pyrazinamide who had higher urine ratios of pyrazinoic acid/pyrazinamide and 5-OH-PA/pyrazinamide.

Fig. 55
figure 55

Metabolism of pyrazinamide to active and toxic metabolites via amidase and XOR catalyzed reactions

Concluding remarks

We have presented an overview of the metabolic reactions of drugs, natural products, physiological compounds, and other (general) chemicals catalyzed by the major non-P450 human oxidoreductase enzymes, i.e., FMOs, MAOs, NQOs, and molybdenum hydroxylases (AOX and XOR). All of these enzymes, in addition to their roles of facilitating excretion of exogenous and endogenous compounds, also catalyze reactions producing toxic products from both physiological compounds (e.g., bioactivation of neurotransmitters by MAO enzymes activity), as well as from xenobiotic compounds under specific conditions (e.g., supra-physiological substrate concentrations, anaerobic vs aerobic conditions, presence of specific inhibitors, presence/absence of cofactors, enzyme polymorphism). The participation of non-P450 oxidoreductases in the activation reactions forming toxic products is relatively low, compared to P450 enzymes (Rendić and Guengerich 2012, 2015). However, important therapeutic agents (including antibiotics, antibacterial, antitubercular, and CNS stimulants) (Table 7) are substrates in some bioactivation reactions catalyzed by FMO enzymes. An important role of MAO inhibitors is as drugs that are used in the clinic to treat depression (Table 11), and potential roles exist for natural products and their derivatives (Tables 12, 13). In addition, potential roles for toxic/reactive metabolites in the MAO-catalyzed metabolic reactions have to be considered with neurotransmitters as substrates (i.e., formation of aldehydes and H2O2) (Table 14). The toxic products can be eliminated by detoxication reactions catalyzed by aldehyde dehydrogenases and aldehyde reductases. Another example of detoxication is illustrated by the deactivation of highly reactive DOPA quinone, which might be formed by oxidation of dopamine or 3,4-l-DOPA by the catalytic activity of tyrosine oxidase (Asanuma et al. 2003; Ito et al. 2020). DOPA quinone might be deactivated by NQO enzymes or by conjugation with glutathione. These examples illustrate the multiple factors that can affect bioactivation/detoxication reactions and the outcome of the metabolic reactions with a particular compound as substrate. The literature also indicates that a number of compounds that are substrates of non-P450 oxidoreductases are also substrates of one or more P450 or other enzymes, and they might also interact with drug transporters in addition. Thus, multiple metabolic properties of a compound/drug have to be considered when drug–drug metabolic interactions or toxicity caused by a compound is evaluated.