Next Article in Journal
Hormonal Gut–Brain Signaling for the Treatment of Obesity
Next Article in Special Issue
Pharmacokinetic Markers of Clinical Outcomes in Severe Mental Illness: A Systematic Review
Previous Article in Journal
Aggregation of Disordered Proteins Associated with Neurodegeneration
Previous Article in Special Issue
The MAOA rs979605 Genetic Polymorphism Is Differentially Associated with Clinical Improvement Following Antidepressant Treatment between Male and Female Depressed Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Human Cytochrome P450 1, 2, 3 Families as Pharmacogenes with Emphases on Their Antimalarial and Antituberculosis Drugs and Prevalent African Alleles

by
Chiratidzo R. Chamboko
1,†,
Wayde Veldman
1,†,
Rolland Bantar Tata
1,
Birgit Schoeberl
2 and
Özlem Tastan Bishop
1,*
1
Research Unit in Bioinformatics (RUBi), Department of Biochemistry and Microbiology, Rhodes University, Makhanda 6139, South Africa
2
Translational Medicine, Novartis Institutes for BioMedical Research, 220 Massachusetts Ave, Cambridge, MA 02139, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(4), 3383; https://doi.org/10.3390/ijms24043383
Submission received: 14 December 2022 / Revised: 30 January 2023 / Accepted: 6 February 2023 / Published: 8 February 2023
(This article belongs to the Special Issue Pharmacogenomics 2.0)

Abstract

:
Precision medicine gives individuals tailored medical treatment, with the genotype determining the therapeutic strategy, the appropriate dosage, and the likelihood of benefit or toxicity. Cytochrome P450 (CYP) enzyme families 1, 2, and 3 play a pivotal role in eliminating most drugs. Factors that affect CYP function and expression have a major impact on treatment outcomes. Therefore, polymorphisms of these enzymes result in alleles with diverse enzymatic activity and drug metabolism phenotypes. Africa has the highest CYP genetic diversity and also the highest burden of malaria and tuberculosis, and this review presents current general information on CYP enzymes together with variation data concerning antimalarial and antituberculosis drugs, while focusing on the first three CYP families. Afrocentric alleles such as CYP2A6*17, CYP2A6*23, CYP2A6*25, CYP2A6*28, CYP2B6*6, CYP2B6*18, CYP2C8*2, CYP2C9*5, CYP2C9*8, CYP2C9*9, CYP2C19*9, CYP2C19*13, CYP2C19*15, CYP2D6*2, CYP2D6*17, CYP2D6*29, and CYP3A4*15 are implicated in diverse metabolic phenotypes of different antimalarials such as artesunate, mefloquine, quinine, primaquine, and chloroquine. Moreover, CYP3A4, CYP1A1, CYP2C8, CYP2C18, CYP2C19, CYP2J2, and CYP1B1 are implicated in the metabolism of some second-line antituberculosis drugs such as bedaquiline and linezolid. Drug–drug interactions, induction/inhibition, and enzyme polymorphisms that influence the metabolism of antituberculosis, antimalarial, and other drugs, are explored. Moreover, a mapping of Afrocentric missense mutations to CYP structures and a documentation of their known effects provided structural insights, as understanding the mechanism of action of these enzymes and how the different alleles influence enzyme function is invaluable to the advancement of precision medicine.

1. Introduction

Precision medicine has been of interest to many scientists over recent years who envision a world where each person or groups of people can receive the right medications for their medical condition based on their genotype. Ideally, the genotype determines the therapeutic strategy, the appropriate dosage and the likelihood of benefit or toxicity. Additional factors affect patient response to medications such as age, sex, disease stage, and environment. However, genetic makeup, [1,2,3] is a critical factor, as it directly determines the likelihood to respond to a targeted therapy or how drugs are metabolized [2,4,5]. Pharmacogenomics, therefore, forms an integral part of precision medicine since it combines both pharmacology and genomics to determine how differences in the human genome, especially the ones located at pharmacogenes, affect drug metabolism and efficacy [5,6,7]. Such genomic differences are caused by genetic variations due to mutations such as single nucleotide polymorphisms (SNPs), insertions and deletions, or copy number variations [8]. These polymorphisms may be population specific or selected for across different populations, leading to slight differences in the expressed proteins or enzymes responsible for drug metabolism that are encoded by pharmacogenes [5,9,10]. The resultant effects of these differences on enzyme activity may be visible as minimal to no therapeutic response in some individuals while others may suffer from adverse drug reactions [7,11]. To minimize harmful side effects, while maximizing the therapeutic effectiveness of drugs in each patient, it is important to understand the effects of genetic variations on pharmacogenes and their associations with variable drug responses in different populations [12]. For instance, identifying the proto-oncoprotein KRAS genotype of a patient provides information about disease aggressiveness, which drugs to be prescribed (e.g., KRAS-G12C inhibitors sotorasib and adagrasib), and drug sensitivity [13].
Functional characterization of pharmacogenes segregates them into three main classes: drug transporters (e.g., P-glycoprotein 1 (P-gp), ATP binding cassette subfamily B member 1 (ABCB1)); drug metabolizers (cytochrome P450 enzymes (CYPs)); and drug targets (e.g., human epidermal growth factor receptor 2 (HER2)). There exists an intricate interplay between these classes of pharmacogenes and genetic variations involving any of them can potentially influence treatment outcomes or pharmacogenetic phenotypes [14,15,16].
The main drug metabolizing enzymes in humans belong to the CYP enzyme superfamily, and are actively involved in phase 1 metabolism of most therapeutic drugs and xenobiotics [17]. CYP enzymes are highly polymorphic, with newly discovered alleles being updated regularly in the Pharmacogene Variation (PharmVar) Consortium database [18]. Certain SNPs from CYP families have been shown to have a great impact on CYP function [19,20], while some have exerted no noticeable change in enzyme activity [1].
In this review, we first provide a general overview of CYP enzymes, and then focus on the three human CYP families (CYP1, 2, and 3) that are particularly important in the metabolism of drugs. We detail the enzymes of these three families, compare similarities and differences between them and combine the data with the 3D structural information where applicable. We further provide up to date variation data attached to these proteins. Although the diversity of CYP enzymes within the African Continent is well appreciated, with implications on drug resistance and adverse drug reactions [20,21], little is known about the underlying mechanisms responsible for such observations. Hence, as a next step, we outline the available data from the African perspective. Firstly, we convene on CYPs that are particularly involved in the metabolism of drugs used in the treatment of two important infectious diseases; malaria and tuberculosis, for which Africa shares the greatest burden [21,22]. Secondly, we review Afrocentric polymorphisms of CYP enzymes with specific focus on potential or known effects on antimalarial and antituberculosis drugs; and discuss the connections between the variations and their positions with possible explanations to their structural/functional effects. Collectively, in this review, by gathering the existing information and by pointing out the current gaps from different aspects, we aim to create a baseline understanding of CYP-drug-variation with specific emphases on malarial or tuberculosis drugs as well as the alleles that are prevalent in the African Continent. In the long term, we believe information shared here could contribute to the development of suitable therapeutic drugs and drug combinations for the treatment of prevailing infectious diseases within the Continent and elsewhere.

2. CYP Enzymes

CYP enzymes are a superfamily of heme-thiolate proteins that are present in all kingdoms of life [23]. They are essential for the production of steroids, cholesterol, prostacyclins, and thromboxane A2 [1,17], and partake in the detoxification of foreign chemicals and the metabolism of drugs [17,24,25]. In humans, they are mainly found within the mitochondria and endoplasmic reticulum of liver cells; however, they can be located in many other parts of the body including the kidneys, small intestine, placenta, and lungs [1,17]. There are 57 known human genes that encode for the different CYP enzymes, and these have been classified into 18 protein families [26]. The enzymes in the first three families (CYP1, CYP2, and CYP3) are responsible for the metabolism of 70–80% of clinical drugs [17,24] and most foreign chemicals; however, within these families, only enzymes CYP1A1, CYP1A2, CYP2A6, CYP2B6, CYP2C8, CYP2C9, CYP2C18, CYP2C19, CYP2D6, CYP2E1, CYP3A4, and CYP3A5 notably metabolize most of clinical drugs [1,27,28]. Most CYPs metabolize more than one drug, and also, a drug is generally metabolized by more than one CYP enzyme [19].

2.1. Human CYP Enzyme Nomenclature

The establishment of standardized CYP enzyme nomenclature dates back to 1987 [29]. At the time, enzymes which belonged to the P450 gene superfamily were classified based on amino acid sequence similarities such that enzymes from different gene families shared ≤36% similarity while those from the same subfamily had ≥70% similarity [29,30]. As more data became available, the classification and nomenclature of CYP enzymes also evolved [29,30,31]. While the original nomenclature was built on only 65 characterized P450 genes [29], the final nomenclature is based on 221 P450 genes [32]. The current classification of CYP enzymes relies more on phylogenetic tree clustering rather than the arbitrary sequence identity values which were subject to changes based on data availability [30,32,33].
The current nomenclature is as follows: CYPxyz, where CYP = Cytochrome P450 and x = family (Arabic numeral), y = subfamily (Capital letter, if more than one exists), z = gene name (Arabic numeral), and this is followed by the letter P if the gene in question is a pseudogene [30,32].
After the CYP nomenclature was agreed upon to gene level, allele representation remained ambiguous until later when the currently acceptable method was proposed for CYP2D6 [32,34]. Here, the name is maintained to the gene level followed by an asterisk, then the allele in Arabic numerals which may be followed by a capital letter denoting silent mutations in the gene as follows: CYPxyz*ij, where CYPxyz is as defined above, i = allele (Arabic numeral), j = silent mutation (Capital letter) [34,35,36]. Further updates have been applied to this method, particularly representing the silent alleles as three decimal point digits after the allele name as seen in the Pharmacogene Variation (PharmVar) Consortium database [18]. The consensus/reference allele, usually representing the major proportion of the population, confers normal activity of the CYP enzyme [35]. An allele is classified as a pharmacogenetic polymorphism if it occurs at a frequency of at least 1% in a population [19]. Sub-alleles are designated when an already-characterized sequence variant is discovered with additional non-causative variant/s [35]. These sub-alleles are given letters in addition to the number (e.g., CYP2B6*4A, CYP2B6*4B, CYP2B6*4C). However, when more than one effective variant exists on the same allele, the given allelic number is based on the variant causing the most serious effect, such as a splice defect (e.g., CYP2C19*2A). Unique allele numbers (e.g., CYP2B6*6) are given to combinations of variants that also occur on their own and are judged to be uniformly effective (e.g., different amino acid substitutions). Importantly, for practical purposes, the earliest alleles ever reported are not based on this nomenclature system as they have not been re-named.

2.2. Human CYP Enzyme Sequence, Structure, and Mechanism of Action

To understand ligand binding/selection and the general structure–function relationship of human CYP enzymes, one must have knowledge of the sequential, structural, and mechanistic features of CYPs.
Human CYP enzymes within families 1, 2, and 3 are known to have similar structure and function. Enzyme sequences within these families contain about 400–500 amino acids [28,37], and those that share ~40% sequence identity are grouped in the same family, while those sharing over 55% sequence homology are subgrouped together into the same sub-family [28,38]. According to [39] there are three conserved short regions/sequence motifs throughout the CYP superfamily. The first conserved region is the (A/G)XXXT which is located in helix I [37,39]. This is the oxygen binding and activation motif that contains the highly conserved threonine which plays an important role in the third step of the catalytic cycle of CYP enzymes [39]. The second most conserved region in CYP enzymes is the EXXR motif, which contains conserved glutamic acid and arginine that build a set of salt-bridge interactions which form the final tertiary structure of the enzymes [39]. The FXXGXXXCXG region is the third most conserved one in the CYP enzyme superfamily [40]. It is a heme-binding domain that contains phenylalanine, glycine, and cysteine, the three most conserved residues in the CYP superfamily, of which the conserved cysteine plays the role of the axial ligand to the heme [39,40].
In contrast to prokaryotic CYPs, eukaryotic CYPs are membrane proteins. Specifically, they are bitopic membrane proteins (spans the lipid bilayer only once) that are found on the cytoplasmic side of the endoplasmic reticulum or on the matrix side of the inner mitochondrial membrane [41]. They have an N-terminal transmembrane helical anchor that extends across the membrane bilayer and is joined by a flexible linker to a large globular domain that partly sits inside the membrane [42]. CYPs cannot function on their own and must bind with a redox partner for electron transfer [43]. These complexes are difficult to crystallize, resulting in limited protein–protein redox complexes in the Protein Data Bank (PDB). During CYP electron transport, reducing equivalents from nicotinamide adenine dinucleotide phosphate (NADPH) are transferred to molecular oxygen [44,45].
CYPs are typically divided into two major classes, class I and class II, based on their cellular location and their redox partners. Class l encompasses mitochondrial and bacterial CYPs which utilize two different redox partners: an iron–sulfur protein (ferredoxin/adrenodoxin) and a flavin-containing reductase (ferredoxin/adrenodoxin reductase). On the other hand, class II CYPs are microsomal monooxygenases that accept electrons from NADPH-cytochrome P450 oxidoreductase (POR). POR is a single polypeptide consisting of one molecule of flavin adenine dinucleotide (FAD) and one molecule of flavin mononucleotide (FMN) [43]. POR transfers electrons from the two-electron donor NADPH to the CYP heme group, with the FAD functioning as a dehydrogenase flavin and FMN as an electron carrier. The CYP enzymes reviewed in here belong to class II.
The fold structure of CYP enzymes is largely conserved throughout the superfamily [37,46]. All known 3D structures of these proteins have a general shape that includes 14 helices and loops denoted A–L (A, B, Β′, C, D, E, F, G, H, I, J, K, K′, L) [23] (Figure 1A). Other helices such as A′, B″, F′, G′, J′, K″, and L′ have been observed to occur in the CYP structure at times [46]. CYPs also contain 4 β-sheets—β1 (5 strands), β2 (2 strands), β3 (3 strands), β4 (2 strands) in their structures, with two additional sheets, β5 and β6, occurring sometimes [23,46]. Although the overall CYP structure is conserved, the size and shape of the active site can differ [43]. The B–C and F–G helix regions are the least conserved and contribute to substrate specificity, especially the B′ helix [43]. These regions change conformations for ligand entry, and together with the F–G loop, form a ‘‘roof” for the active site opposite to the heme ‘‘floor” [47,48]. The highest structural conservation in CYPs is centered around the heme–thiolate oxygen activation chemistry [43] (Figure 1B).
The CYP active site is situated in the protein hydrophobic core which is linked to the enzyme surface by channels [49,50] that substrates and O2 molecules enter, products exit, and water molecules move [51]. CYP enzymes contain a prosthetic heme group in their active site that is crucial for enzyme function during metabolism [43]. It is located between helices I and L and is covalently linked to the enzyme through the sulfur atom of a conserved cysteine residue that acts as the proximal axial thiolate ligand for the heme iron [43,52,53] (Figure 1B). The heme is regarded as a reactive center to activate oxygen and to oxidize the substrate [52]. Coordinated water gates controlled by the propionate side chains of heme, and their salt bridge partners facilitate water movement [43]. A commonly conserved acid/alcohol pair that is important for the P450 active cycle is situated on the I-helix [54], whilst the loop before the L-helix houses the cysteine axial thiolate ligand [52]. The oxidative prowess of CYP enzymes is linked to the formation of a coupled high-valent iron (IV)–oxo porphyrin π-cation radical species (Compound I) involving the iron ion of the enzyme’s heme group, dioxygen, two reducing, and two proton equivalents supplied within the CYP enzyme’s catalytic cycle [51,55,56]. The highly conserved β-bulge region (FXXGXXXCXG) containing the cysteine axial thiolate ligand is rigid so as to maintain a hydrogen-bonding distance from two peptide NH groups (seen in all CYPs) [43], even though the hydrogen-bonding geometry supports only one hydrogen-bond. These hydrogen-bonds play a role in controlling the heme iron redox potential [57,58], which would be too low for reduction by redox partners without the hydrogen-bonds. It is thought that in order to sustain a physiologically accessible range for the redox potential, the protein must have a satisfactory electrostatic environment around the cysteine ligand. While there is evidence pointing to Compound I as the ultimate oxidant in CYP enzymes, the oxidative versatility of the enzymes is likely linked to their plasticity, especially within the active site, and the ability of the catalytic cycle to be initiated by the entry of an oxidisable substrate into the active site [51]. On the other hand, unique features within the active sites/entry channels of different CYP enzymes—such as those due to mutations—modulate the stereo- and regioselectivity associated with their oxidation of different substrates and may result in different oxidized products of the same substrate [51,59,60].
All CYP enzymes in the 18 human families identified are involved in phase 1 metabolism [61,62], where they metabolize various endogenous and exogenous compounds such as xenobiotics and environmental chemicals [63]. These enzymes function as monooxygenases [64,65], carrying out the metabolism of drugs through oxidation reactions, including hydroxylation, epoxidation, sulfoxidation, C–C bond cleavage, and desaturation reactions [66,67,68,69,70,71,72,73,74]. The overall CYP canonical reaction comprises the reductive scission of the O–O bond of atmospheric dioxygen to release a single molecule of water with the transfer of a single oxygen atom to the substrate [75,76]. The convoluted catalytic cycle has multiple reactions with transient intermediates, where the product of one reaction is the substrate for the next reaction by the same enzyme. During the mechanism, the heme iron changes spin states [77] due to the binding of substrate, as the outer orbital has one or more unpaired electrons giving it a higher energy state. The following is a short description of the typical seven-step catalytic cycle seen in CYP enzymes (Figure 2) [51,78,79]:
Before the substrate enters the enzyme, the heme iron is coordinated to a water molecule and is in a low-spin ferric resting state (S = 1/2). Step 1: The cycle starts when the substrate first enters the active site and interacts with the resting state. The water molecules leave the pocket [71] which detaches the aqua ligand resulting in the high-spin FeIII–heme complex (S = 5/2). Step 2: The now more positive redox potential [80,81] allows electron transfer from the reducing partner, to the now ferrous FeII complex. Step 3: As the ferrous FeII complex is a good O2 binder, it takes up an O2 molecule and is transformed into the oxyferrous complex. Step 4: As the oxyferrous complex is a good electron acceptor, it is reduced and is transformed into the peroxo complex. Step 5: Around this time, the water molecules that departed the active site pocket return through a water gate and create a water channel that protonates the peroxo complex to give Compound 0 (Cpd 0) [51], which is debated to be a putative oxidant [82]. Step 6: The negatively charged Cpd 0 is a good Lewis base that accepts an extra proton, freeing a water molecule, and forms Compound I (Cpd I) which is a ferryl (FeIV)-oxo-π porphyrin cation radical and the ultimate oxidant. Step 7: Cpd I is thought to be responsible for the bond activation in the substrate via hydrogen abstraction, leading to substrate oxidation [73]. The product now leaves the pocket and a water molecule takes its place. The initial CYP enzyme resting/ferric state is now re-established and set for another cycle.
Figure 2. Mechanism of CYP enzymes in a seven-step catalytic cycle. Steps, explained in the text, are indicated with red numbers. Enzyme open, semi-open, and closed conformations at each stage denoted by a grey shape. The substrate, hydrogens, electrons, and oxygen are denoted by yellow, purple, green, and dark blue circles. Water is shown by a blue drop [51,78,79,83,84].
Figure 2. Mechanism of CYP enzymes in a seven-step catalytic cycle. Steps, explained in the text, are indicated with red numbers. Enzyme open, semi-open, and closed conformations at each stage denoted by a grey shape. The substrate, hydrogens, electrons, and oxygen are denoted by yellow, purple, green, and dark blue circles. Water is shown by a blue drop [51,78,79,83,84].
Ijms 24 03383 g002
Dubey and Shaik 2019 [51] researched the determining factors that allow the sequence of steps to occur, using molecular dynamics (MD) and quantum-mechanical/molecular-mechanical (QM/MM) calculations on mostly bacterial CYPs, namely bacterial CYPBM3 and CYPCAM, but also human CYP3A4. Substrate binding, O2 entrance, reductase attachment, and gating were researched using MD simulations [85,86,87,88]. Water channel formation and reorganization generally need about 100 ns of simulation time. Substrate binding and protein–protein interactions need between 350–400 ns and 1000–1500 ns, respectively. MD snapshots are used for their QM/MM calculations [73,74], which calculate the active species and their chemical events and provide geometries, electronic structures, mechanisms, and energies of the species in their native protein environment. It was discovered that the CYP catalytic cycle is initiated by substrate binding and from then on each step is automatically coordinated due to entropy rise [71] and switchable weak interactions [51].

2.3. Existing Structural Information in the Protein Data Bank (PDB)

As of November 2022, Protein Data Bank (PDB) [89] has a total of 177 structures for CYP 1, 2, and 3 family members. The enzymes CYP1A2, CYP2D6, and CYP3A4 were among the first few enzymes within the CYP family to have structures in the database, with the first structure being deposited in 2003. More than a decade later, many of the 23 human CYP enzymes from families 1, 2, and 3 have structures available. Currently, the enzymes with no structures within the database are CYP2A7, CYP2C18, CYP2F1, CYP2J2, CYP2S1, CYP2U1, CYP2W1, and CYP3A43. The enzyme CYP3A4 has the largest structural data available, with 87 structures in PDB. CYP2A6 and CYP2D6 have the second largest number of structures available, with 14 structures each. The PDB also contains variants of CYP enzymes within these three families such as CYP2C9 alleles *2, *3, *8, and *30. A summary of the best structures for each enzyme in the three families is shown in Table S1.
The structures of CYP enzymes in PDB are either in an open, partially open or closed conformation [83]. The open conformation is usually associated with ligand free proteins and the closed conformations are proteins with a ligand bound; however, closed ligand-free structures and partially open ligand-bound structures have been determined [83,84]. Variations in the pre-helix A, the B–C and F–G regions have been defined to cause the different conformations of human CYP enzymes [83,84,90,91,92]. CYP2D6 in an open and closed conformation as an example, is shown in Figure S1. When closed, the pre-helix A G42 residue is 14.3 Å from helix F’ residue E222, but in an open conformation, the distance between helix F’ E222 residue and the pre-helix A residue G42 increases by 0.6 Å to 14.9 Å, thus opening the substrate channel for ease of access to the active site [84]. Subsequent changes in the adjacent regions including the F and G helices, and the helix B-C loop that contains helix B’ follow, creating a CYP2D6 open conformation structure [84].

Antimalarial and Antituberculosis Drugs Complexed with CYP Enzymes

Currently, 3D structural data of CYPs with antimalarial drugs are extremely scarce and non-existent for the antituberculosis drugs. Furthermore, CYP structural information with its functional proteins and membrane partners are not available. This information deficiency hinders the attempts to understand the changes in the drug metabolism due to specific variations.
Of the 177 structures available for human CYP enzymes in families 1, 2, and 3, only two structures are complexed with antimalarial drugs: CYP2D6 with quinine (4WNV) and CYP2D6 with Quinidine (4WNU). X-ray diffraction method was used to obtain the complexes, and they were refined to resolutions of 2.26 Å and 2.35 Å for the quinidine and quinine complexes, respectively [84]. These antimalarial drugs are inhibitors of CYP2D6 and are not naturally metabolized by this enzyme [84,93,94]. They are bound to the active site of CYP2D6, and several differences in binding conformations were discovered [84].
Quinine is a the diastereomer of quinidine, and it is known that these two compounds differ in configurations of the hydroxylated chiral carbon which connects the aromatic quinolone moieties and the protonated and positively charged bicyclic quinuclidine rings of both compounds [84]. As seen in Figure 3, due to the configurational differences between quinine and quinidine, the quinolone rings of the two compounds bind to significantly different positions within the CYP2D6 active site, which leads to different poses overall for the two compounds. Additionally, in the quinuclidine ring of the bound quinidine, the protonated nitrogen is bound to a water molecule in the entrance channel and is positioned almost equidistant between the negatively charged Glu-216 (4.9 Å) and Asp-301 (4.8 Å) side chains [84]. The differences observed for the binding of quinidine (Figure 3A) and quinine (Figure 3B) within CYP2D6, are contributing to the differences in inhibition of the two compounds [84]. It has been reported that while quinine is a potent inhibitor, quinidine is more potent [95]. More information on the complexes can be found in the article by Wang et al. 2015 [84].

2.4. Structural Differences in Human CYP 1, 2, 3 Enzymes

As indicated in the previous sections, the 3D structures of mammalian CYPs are generally conserved; however, some sequence and structural variations in the B–C helix, F–G helix, and L helix regions are known [96]. The differing shapes and sizes of the active site cavities of human CYPs are a result of discrepancies of the F–G and B’ helices. The dissimilarity of these regions causes some variability of enzyme function.
CYP family 1 enzymes generally metabolize polynuclear aromatic hydrocarbons, and the structures of human CYP1A1 [92], 1A2 [96], and 1B1 [96] possess narrow active site pockets that suit the size and planarity of polynuclear aromatic hydrocarbons. For example, CYP1A2 has a preference for small planar aromatic or heterocyclic amine ligands, engendered by active site-facing polar residues Thr118, Ser122, and Thr124 that are located in the B’–C helices region, as well as residues Thr223 (F helix) and Asp320 (I helix) that are found on the roof of the active cavity [96]. On the other hand, because CYP2A6 has a smaller active site pocket, it prefers to bind small aromatic ligands [97] with the help of the substrate-orienting residue Asn297 [98]. CYP2C9 prefers weakly acidic substrates [99] due to its disordered arrangement of the F–G loop region and an additional turn at the N-terminal side of helix A [100]. CYP2C9 has a smaller active site cavity than CYP2C8 but bigger than CYP2A6 [101]. With a comparatively small active site pocket, CYP2E1 prefers neutral compounds with low molecular weight and fatty acids [102]. In the CYP2E1 active site cavity, a conserved Asp295 is critical for substrate recognition and ligand binding. The very flexible CYP3A4 enzyme has a large binding cavity, allowing many structurally diverse ligands to bind [103]. This could be aided by the fact that the heme surface of CYP3A4 enzymes is much more exposed to substrates [47,104]. CYP3A4 has comparatively short F and G helices, and a cluster of large hydrophobic phenylalanine residues situated on the roof of the active site in the ligand-free structure that are distorted upon ligand binding [104,105,106]. The shorter CYP3A4 F-helix does not cross above the active site, meaning the active site can expand and contract due to the positional fluctuations of helices F′ and G′ as well as the flexibility of the long connector between helix F and F′ [105]. A high sequence identity between CYP3A4 and CYP3A5 result in similar structure; however, the CYP3A5 active site cavity has extra space between the helix F–F’ and helix G’–G connectors due to the shorter Leu108 in CYP3A5 (Phe108 in CYP3A4), as well as the longer CYP3A5 Phe210 (Leu210 in CYP3A4) which increases the size of the upper region of the substrate binding cavity [107].

3. CYP 1, 2, 3 Enzyme Families and Antimalarial Drugs

Antimalarial drugs are used either for chemoprevention or for the treatment of uncomplicated and severe malaria. Current indications include, sulfadoxine pyrimethamine (SP) for intermittent preventive treatment in pregnancy (IPTp) and infants (IPTi), and seasonal malaria chemoprevention (SMC) in children, with the latter including the addition of amodiaquine, atovaquone + proguanil are also indicated as prophylaxis for travelers to endemic areas [108]. Treatment of uncomplicated malaria is by use of artemisinine-based combination therapies (ACT) including artemether + lumefantrine, artesunate + either amodiaquine, mefloquine, SP, or pyronaridine, and dihydroartemisinin + piperaquine while severe malaria requires prior intravenous or intramuscular artesunate administration followed by oral ACT [108]. Finally, primaquine is recommended alongside ACT to reduce transmission while quinine and clindamycin are recommended for the treatment of uncomplicated malaria in pregnancy [108]. Diverse toxicity profiles are associated with some antimalarial drugs, for example quinine: rare cardiovascular toxicity and hypoglycemia, mefloquine: dose-related serious neuropsychiatric toxicity, pyrimethamine + dapsone: agranulocytosis, SP + amodiaquine: potentially fatal reactions, chloroquine + proguanil: mouth ulcers and gastrointestinal upset, halofantrine: cardiotoxicity [109,110].
Antimalarial drug metabolism may result in bioactivation, altered toxicity, or preparation of the drug for excretion [111]. Like with other xenobiotics, phase I metabolism of most antimalarial drugs is performed by CYP enzymes—particularly those belonging to the CYP 1, 2 and 3 families [1,112,113,114,115].
CYP enzymes function by catalyzing the oxidation of organic substrates, rendering the substrates more hydrophilic [51,55]. Following oxidation, the substrate either becomes bio-activated—like with the prodrugs amodiaquine and proguanil [111,116], directly excreted, or conjugated in phase II metabolism prior to excretion like with dapsone and artemisinins [28,117,118]. CYP metabolism may also result in more toxic oxidized products as seen with dapsone metabolism and hemolytic anemia and methemoglobinemia in humans [119]. Some of the antimalarial drugs on the market with known CYP enzymes as metabolizers and their oxidized products are presented in Table 1.

4. CYP 1, 2, 3 Enzyme Families and Antituberculosis Drugs

Months-long tuberculosis (TB) treatment involves the combination of several antituberculosis drugs to prevent the development of a resistant strain [2,140]. For drug-susceptible TB, international guidelines have advised the combination of all first-line drugs isoniazid, rifampicin, ethambutol, and pyrazinamide for 2 months followed by 4 months of isoniazid and rifampicin [141]. However, resistant strains of TB have emerged leading to the development and usage of second line drugs [142,143]. For multi-drug resistant (MDR) TB, a 6-month “BPaLM” regimen of bedaquiline, pretomanid, linezolid, and moxifloxacin may be used in patients older than 14 years of age with MDR/rifampicin-resistant (RR) TB who have not had previous exposure to bedaquiline, pretomanid, and linezolid [144]. Moxifloxacin can be omitted here (“BPaL”) if the TB is resistant to fluoroquinolones. A new drug in the same class as linezolid that is currently in testing, namely sutezolid, could have better therapeutic value than linezolid [145]. Unfortunately, with TB, drug associated toxicity is common due to the amount of antibiotics used and the long treatment period. Typically, second-line TB drugs are less effective and more toxic than the first-line drugs.
Drug-susceptible TB treatment has a 95% cure rate (in optimal conditions); however, several unclear concerns exist such as the high variability of response, the likelihood of drug underexposure, the high prevalence of drug-related toxicity, and the selection of multidrug-resistant strains [141]. There is a scarcity of early biomarkers for predicting treatment efficacy, cure, and the determination of patients needing prolonged treatment [146]. Inter-individual variability in the pharmacokinetics of anti-tubercular drugs may be responsible for the variability of response as it’s been shown to have major influence over the sterilizing effect and selection of phenotypic resistance. Low maximum plasma concentrations are linked to treatment failure, relapse, and acquired drug resistance, whereas high plasma concentrations are linked to hepatoxicity [147,148]. Therefore individual/personalized treatment based on patient pharmacogenetics is extremely important.
The frequency of hepatotoxicity in patients receiving anti-tubercular treatment is between 2% and 28% [149]. Toxic metabolites play a role in hepatotoxicity, but specific mechanism is not known. The human liver metabolizes most drugs, requiring several reactions and drug metabolizing enzymes [150]. The rate-limiting step for the clearance of drugs from the body are phase I metabolic reactions, then during phase II metabolic reactions the drug and its metabolites associate with endogenous substances and are released from the body. This action does have a detoxification effect; however, several active metabolites that are simultaneously produced can cause liver damage.
First-line antituberculosis drugs undergo different pathways during metabolism, but none have been discovered to be directly metabolized by any enzymes from the CYP families. However, metabolism of the first-line antituberculosis drug isoniazid is indirectly influenced by CYP2E1, as is discussed in the next article section
The second-line antituberculosis drug bedaquilineis mainly metabolized by N-demethylation via CYP3A4, and less so via CYP1A1, CYP2C8, CYP2C18, and CYP2C19 [151,152]. Bedaquiline is metabolized to N-desmethyl-bedaquiline, N-didesmethyl-bedaquiline, and two hydroxyl metabolites [153,154]. Another study showed N-dealkylation of bedaquiline to produce the aldehyde metabolite M5, primarily mediated by CYP3A4 [152,155]. Bedaquiline has a novel mechanism of action towards mycobacterial ATP synthase [155]; however, despite being highly effective it has been linked to adverse cardiac and hepatic drug reactions [155,156,157]. Recently, another second-line antituberculosis drug linezolid, has been shown to be metabolized via CYP2J2 and CYP1B1, by 2-hydroxylation and de-ethyleneation of the morpholine moiety of linezolid [158] It has been suggested that the less studied CYP2J2 accounts for approximately 50% of linezolid hepatic metabolism [158]. The second-line antituberculosis drug delamanid is mostly metabolized by albumin and to a lesser extent by CYP3A4. The entire metabolic profile of delamanid has not yet been uncovered [159].

5. Inhibition and/or Activation of CYP Enzymes by Drugs, and Drug–Drug Interactions

As seen earlier, CYP enzymes constitute the most important contributors to oxidative metabolism of drugs, and their inhibited and/or induced activity is an important determinant of several drug–drug interactions [160]. Some of the presently available antimalarial and antitubercular drugs have been observed to inhibit or induce human CYP enzymes which may cause drug–drug interactions that could either cause adverse drug reactions or decrease the efficacy of the drugs metabolized by these enzymes [2,3,141,161,162].
Recently, OZ439, a potent synthetic ozonide that is currently used for the treatment of uncomplicated malaria was found to inhibit CYP3A4 through both direct and mechanism-based inhibition [163]. Since CYP3A4 is involved in the metabolism of several antimalarial drugs such as artemisinins and their derivatives, chloroquine, dapsone, halofantrine, lumefantrine, mefloquine, primaquine, and quinine (Table 1), a combination therapy involving OZ439 and any of these drugs may not be advisable. Moreover, chloroquine has been implicated in decreased activity of CYP2D6, suggesting autoinhibition of its metabolism by this enzyme, while its less toxic derivative, hydroxychloroquine, also inhibited the enzyme significantly [113]. More evidence over the years have suggested that artemisinin and its relative drugs dihydroartemisinin, artesunate and artemether inhibit CYP1A2, 2B6, 2C19, and 3A4 [113,160,164,165,166,167]. CYP1A2 and CYP2C19 were noticeably inhibited by artemisinin and dihydroartemisinin, and in healthy individuals, CYP2D6 activity was 66% inhibited in vivo by artemisinin [160,167,168]. The activity of CYP2B6 was also reported to be inhibited by artemisinin derivatives in vitro [164,169]; however, inhibition by artemisinin itself proved to be weak [164]. Thus, while some artemisinins may be autoinhibitive to one of their metabolising enzymes CYP2B6, a combination of artemisinins with either mefloquin, primaquin, chloroquine, or proguanil may prove counterproductive.
Induction of human CYP enzymes has also been described, with evidence of artemisinin as a potent inducer of CYP1A2, CYP2A6, and CYP3A4 transcription and activity [164]. Artemisinin induces CYP enzyme expression, by activating two closely related nuclear hormone receptors responsible for the transcriptional regulation of CYP enzyme expression [170,171]. These are the pregnane X receptor (PXR) and the constitutive androstane receptor (CAR) which undergo either separate or cooperate induction by the artemisinin [170,171].
Like with antimalarial drugs, antituberculosis drugs such as isoniazid, rifampicin and ciprofloxacin have also been implicated in the inhibition or induction of CYP enzymes. Isoniazid is an inhibitor of CYP enzymes CYP1A2, CYP2A6, CYP2C9, CYP2C19, CYP2E1, and CYP3A4 [161], while ciprofloxacin inhibits CYP1A2, CYP2D6, and CYP3A4, and because of this inhibition, metabolism by these CYP enzymes is slowed down and a buildup of their substrates within the body occurs [2,3,161,172]. This is particularly unfavorable, as enzyme inhibition can cause potentially serious adverse events [3,161]. Administering isoniazid and ciprofloxacin alongside antimalarials such as mefloquin, primaquin, artemisinin, and all the CYP3A4 metabolized drugs highlighted above may result in adverse reactions.
In contrast to isoniazid and ciprofloxacin, rifampicin is a well-known inducer of several CYP enzymes in the CYP2A, CYP2B, CYP2C, and CYP3A family subgroups, but evidence has highlighted the high induction of CYP3A4 specifically [2,161]. The induction of these CYP enzymes results in an increased elimination of administered drugs, and this often results in reduced pharmacological effects [2,172]. For instance coadministration of amodiaquine and rifampicin in healthy volunteers resulted in significant decreases in the critical pharmacokinetic parameters of the drug, as opposed to increases in those of the main metabolite desethylamodiaquine, leading to a significant increase in the metabolic ratio from 1.55 to 2.68 [173]. No other TB drugs have been documented to inhibit or induce the CYP enzyme system; however, there were some investigations carried out on pyrazinamide and ethionamide to identify any inhibition of CYP enzymes by these drugs [174]. The results from that study revealed that although the three drugs (isoniazid, pyrazinamide and ethionamide) are closely related chemically, they do not all inhibit CYP enzymes to the same degree and thus pyrazinamide and ethionamide were disregarded as inhibitors of CYP enzymes [174].
Isoniazid and its metabolic intermediates are regarded as the major source of hepatotoxicity in TB patients [175,176]. Isoniazid goes through hepatic metabolism by the N-acetyltransferase (NAT) enzyme system [177,178], being acetylated by NAT to acetyl-isoniazid and then biotransformed by hydrolysis via amidase to monoacetylhydrazine [2,179] (Figure 4). Monoacetylhydrazine is associated with hepatotoxicity and the most important detoxifying step to prevent this is a further acetylation step that produces the non-toxic product diacetylhydrazine [180,181]. However, in slow acetylators, extra monoacetylhydrazine is thought to be oxidized by CYP2E1 to toxic reactive metabolites instead of conversion to non-toxic diacetylhydrazine via NAT [182]. Moreover, when acetylation of isoniazid is slow it will be biotransformed to the toxic compound hydrazine and then most likely oxidized by CYP2E1 to produce reactive acetyl onium ions and acetyl radicals leading to hepatotoxicity [175,183,184]. Although probable, there is currently no explicit evidence of CYP2E1 oxidizing hydrazine and monoacetylhydrazine [185,186,187]. The enzyme glutathione s-transferase (GST) is an intracellular free radical scavenger that detoxifies the toxic reactive metabolites generated from antituberculosis drugs and other xenobiotics. Therefore, individuals with slower NAT2 activity, faster CYP2E1 activity, and slower GST activity are the most likely to experience hepatotoxicity from antituberculosis drugs [141,188,189,190,191]. Additionally, because rifampicin induces CYP2E1 [192] and isoniazid hydrolases [193], the incidence of hepatotoxicity has been found to be higher when isoniazid is combined with rifampicin.
The complexity increases when the HIV antiretroviral drug efavirenz is introduced. Efavirenz is transformed into inactive metabolites by CYP2B6, and less so by CYP2A6 [194]. Rifampicin induces CYP2B6 and should decrease efavirenz in the plasma [195]. Contrarily, a few patients prescribed with rifampicin and isoniazid have seen their efavirenz plasma levels increase (especially patients with slower activity CYP2B6 and NAT2 genotypes) [196,197,198]. This could be due to high isoniazid concentrations inhibiting CYP2A6 in NAT2 slow acetylators, as CYP2A6 activity is important for efavirenz clearance in CYP2B6 slow metabolizers [196,198,199].

6. Polymorphism in CYP Enzymes with Specific Focus on Antimalarial and Antituberculosis Drugs

The genes coding for CYP enzymes are highly polymorphic across populations, and this might lead to different enzyme activities impacting the efficacy of the drugs that they metabolize. The enzyme activity might increase, decrease or becomes nonexistent in some cases [7,11]. CYP gene variation engenders phenotypes classified as ultra-rapid, extensive, intermediate and poor metabolizers [35]. Extensive metabolizers have two “normal activity” alleles, generally called the *1 or consensus/reference allele. An ultra-rapid metabolizer usually has duplicated or multi-duplicated gene copies of the same allele (although this may not always be the case), intermediate and poor metabolizers have one and two defective alleles (e.g., gene inactivation or deletion), respectively [200,201,202].
CYP families 2 and 3 are implicated the most in terms of differences in metabolism due to variants. For example, a small study of adverse effects of CYP2A6 alleles and the antimalarial prodrug artesunate showed considerably higher adverse effects in patients with CYP2A6*1B variants [127]. CYP2A6*1B has a known ultra-rapid metabolism phenotype, which suggests that these patients experience adverse effects because of the accumulation of the active metabolite dihydroartemisinin. The polymorphism of CYP2C8 has resulted in missense mutations in alleles CYP2C8*2 and CYP2C8*3 that have been widely studied [133]. CYP2C8*2 occurs more frequently in African populations, while CYP2C8*3 is more frequent in Caucasians. Both of these alleles of CYP2C8 have been associated with slower metabolism, where there is a 50% reduction in enzyme activity in CYP2C8*2 and an 85% activity reduction in CYP2C8*3, as compared with the wild type/reference allele [203]. CYP2C8 along with CYP2B6, CYP3A4, and CYP3A5 metabolize drugs used in artemisinin-based combination therapy for malaria, and alleles such as CYP2C8*2, CYP2C8*3, CYP2B6*6, CYP3A4*1B, and CYP3A5*3 may affect patient response to this treatment [21]. The presence of alleles CYP2B6*6 and CYP3A4*1B vary significantly between African populations: 22% to 51% and 50% to 80%, respectively [204]. CYP2B6*6 is linked to a poor metabolizer phenotype for the antimalarial drug artemether [205]. Chloroquine is mostly metabolized by CYP2C8, but also by CYP2C9 and CYP3A5. The genetic variants of these enzymes may influence chloroquine pharmacokinetics [128]. For example, it has been seen that persons having “normal” CYP2C8 alleles (extensive metabolizers) have fewer gametocytes compared to persons with reduced-activity CYP2C8 alleles (*2, *3, and *4). CYP2C19 is known to metabolize proguanil and chlorproguanil, and its alleles CYP2C19*2 and CYP2C19*3 are largely associated with poor metabolism of these antimalarial drugs [111]. On the other hand, the CYP2C19*17 allele increases CYP2C19 expression and activity. Regarding treatment with the drug primaquine, there is evidence of patients with CYP2D6 poor-metabolizer alleles (*2, *4, *5, *10, *17, and *41) having therapeutic failure [206]. Lastly, CYP3A4 and CYP3A5 enzymes with poor-metabolizer phenotypes (CYP3A4*22, CYP3A5*3, CYP3A5*6, and CYP3A5*7) have been linked with quinine adverse reactions [207,208,209]. The impact of CYP genetic variants on antimalarial drugs and their clinical implications has also been reviewed in a recent article by Soyinka et al. 2022 [210].
With regards to TB, clearance of bedaquiline was 52% faster in Africans using a population pharmacokinetic model [211], which may be due to Africans expressing significantly more CYP3A5 compared to other populations [212]. Although bedaquiline is metabolized by CYP3A4, substrate specificity frequently overlaps between CYP3A4 and CYP3A5. A recent study found that in South Africans treated for MDR-TB, CYP3A5*3 was linked to slower bedaquiline clearance. The CYP3A5*3 allele results in nonfunctional CYP3A5 protein [213,214]. Several review articles mention that in East Asian populations, CYP2E1 RsaI/PstI polymorphisms are linked to an elevated risk of developing antituberculosis drug-induced liver injury [3,215,216,217,218]. In patients having the CYP2E1 RsaI/PstI c1/c1 genotype, isoniazid has a lower inhibitory effect on CYP2E1 activity compared to other genotypes, and therefore these patients have an increased chance of extra hepatotoxin production which then causes liver injury [216,218,219]. Moreover, the CYP2E1 rs2031920 variant genotype increases CYP2E1 activity, creating extra hepatotoxic metabolites of antituberculosis drugs (especially isoniazid) [220,221]. Conversely, some studies have found no significant connection between CYP2E1 genotype and antituberculosis drug-hepatoxicity [222,223]. In addition to CYP2E1, there may be a link between hepatotoxicity and polymorphisms in genes coding for CYP2C19: rs4244285 and rs4986893 polymorphisms may cause loss of gene functions and subsequently antituberculosis drug-induced liver injury [224]. Hypermethylation of CYP2E1 and CYP2D6 may increase the chance of liver injury from antituberculosis drugs as well [225,226].

Afrocentric Missense Mutations

In line with high genetic diversity in the Continent compared to the rest of the world, CYP enzymes have also been shown to have greater diversity in Africa, possessing novel unique alleles [20,227,228]. Furthermore, comparative studies have identified population structuring at CYP genes, possibly associated with intra-African differences in response to drug therapies as well as the high rates of adverse drug reactions registered in the Continent [20,204,229]. This highlights the need for optimization of drug therapy and drug development in Africa.
All CYP alleles related to the CYP enzymes in families 1, 2, and 3 that metabolize antimalarial and antituberculosis drugs were obtained from the PharmVar website (https://www.pharmvar.org/genes; accessed on 17 November 2022) in November 2022, and a full list of the alleles has been provided in Table S2. Of these, the alleles with the highest frequency in African populations and/or the alleles highly restricted to African populations, according to the recent comprehensive study by Zhou and Lauschke 2022 [230], were chosen from the list (Figure 2 and Table 2). The highest frequency being the allele in each sub-family that occurs most frequently in Africans (besides the reference allele), and highly restricted being the alleles that occur in Africans at least 4x the frequency of any other population type. These Afrocentric alleles were mapped to CYP enzyme structure (Figure 5), and a documentation of their known structural effects and functional consequences are provided (Table 2). A high proportion of Africans possess CYP2D6 alleles that are almost exclusive to the African population. A few notable alleles highly restricted to Africans are CYP2A6*17 (10.9%; second highest frequency of only 0.9% in Middle Eastern population), CYP2B6*18 (7%; second highest frequency of only 0.6% in Middle Eastern population), CYP2C9*9 (7.5%; second highest frequency of only 0.9% in Middle Eastern population), CYP2D6*17 (20.5%; second highest frequency of only 0.7% in admixed Americans), and CYP2D6*29 (8.9%; second highest frequency of only 0.4% in admixed Americans). The admixed American individuals with these CYP2D6 alleles are most likely African Americans. Moreover, worth mentioning, CYP2A6 alleles *23, *25, and *28 are almost exclusively found in Africans. All allele frequencies across all populations are available in the article by Zhou and Lauschke, 2022 [230].
A visual representation of where these 22 missense mutations are located on the general CYP structure is shown in Figure 5. Two of these Afrocentric missense mutations are situated near the catalytically important region that contains the conserved glutamic acid and arginine (EXXR) and decrease enzyme activity. As stated earlier, these two residues build salt bridge interactions which form the final tertiary structure of CYP enzymes [39]. The CYP2C9*5 missense mutation D360E is situated two residues away from the conserved EXXR region (Table 2) and was shown in molecular dynamics simulations to break the hydrogen bond between D360 and S478 (helix K and loop β4) leading to local structure destabilization [239]. Both aspartic acid and glutamic acid are negatively charged residues; however, glutamic acid is larger than aspartic acid. The other mutation, V365M from CYP2A6*17, forms part of the CYP2A6 active site region [231] and is four residues away from the conserved EXXR signature region (Figure 5A). Both valine and methionine are non-polar hydrophobic residues; however, methionine is larger and contains a sulfur atom [231]. The most frequent and restricted Afrocentric allele, CYP2D6*17 (20.5%; second highest frequency of only 0.7% in admixed Americans), contains mutation T107I that is located in substrate recognition site 1 (B′-helix) and causes changes in hydrogen bonds with surrounding residues within the active site cavity [247]. The hydrophilic threonine has an alcoholic side chain pointing down towards heme, whereas the hydrophobic isoleucine has a long aliphatic side chain points away from active site [246]. This CYP2D6*17 allele decreases enzyme activity, most likely due to the T107I mutation, as allele CYP2D6*2 contains the same mutations as CYP2D6*17 except for T107I but results in normal activity. Very concerning is allele CYP2B6*18, which is the only commonly found CYP2B6 missense mutation that causes enzyme inactivation [236,237], and it is highly restricted to Africans [230]. This allele contains mutation I328T (non-polar hydrophobic isoleucine to polar hydrophilic threonine) that is located in the substrate binding region. Another important Afrocentric allele, CYP2D6*29 (decreased enzyme function), contains substrate-binding region mutations V136I and V338M, with V136I also possibly contributing to cytochrome P450 reductase binding [245,248]. Expression studies reveal that both the V136I and the V338M mutations affect catalytic activity, and the effect is stronger when present together as seen in CYP2D6*29 [249].

7. Conclusions and Future Perspectives

Cytochrome P450 (CYP) enzyme families constitute the main pharmacogenes of humans, among which families 1, 2, and 3 are implicated in the phase I metabolism of most therapeutic or clinical drugs. An understanding of the mechanism of action of these enzymes is invaluable to the advancement of precision medicine. This review covers current knowledge on CYP enzymes in general, and more specifically on the first three CYP enzyme families with a focus on prevalent African alleles, particularly linking to two infectious diseases of poverty—malaria and tuberculosis—of which Africa has the highest burden.
Afrocentric alleles that may affect malaria treatment and the implicated antimalarials include: CYP2A6*17, CYP2A6*23, CYP2A6*25, CYP2A6*28 (artemisinin, artesunate), CYP2B6*6, CYP2B6*18 (artemisinin, artesunate, arteether), CYP2C8*2 (chloroquine), CYP2C9*5, CYP2C9*8, CYP2C9*9 (dapsone), CYP2C19*9, CYP2C19*13, CYP2C19*15 (dapsone, quinine, proguanil), CYP2D6*2, CYP2D6*17, CYP2D6*29 (chloroquine), and CYP3A4*15 (arteether, artemisinin, artemether, artelinic acid, chloroquine dapsone, halofantrine, lumefantrine, mefloquine, primaquine, quinine, quinidine).
Although there are reports on CYP enzymes that are involved in the metabolism of different antimalarial drugs, currently there is no literature regarding if first-line antituberculosis drugs are metabolized by any CYP enzymes. However, two second-line antituberculosis drugs bedaquiline and linezolid are known to be metabolized by CYP enzymes; the former mostly by CYP3A4 and the latter by CYP2J2 and CYP1B1. On another note, the first-line antituberculosis drug isoniazid is a known inhibitor while rifampicin is an inducer of several CYP enzymes, resulting in altered pharmacological effects of drugs that are metabolized by these CYPs. Further, variants of CYP2E1 and CYP2C19 have been implicated in altered isoniazid metabolism. Additionally, drug–drug interactions that influence the metabolism of antituberculosis, antimalarial, and other drugs, were explored.
As a first step to understand the mechanism by which these SNPs influence enzyme function, identified Afrocentric missense mutations were mapped to a CYP structure, and their known structural effects and functional consequences tabled. These afrocentric missence mutations indicate that they might play an important role in altering the metabolism of antimalarials amodiaquine, arteether, artelinic acid, artemether, artemisinin, artesunate, chloroquine, dapsone, halofantrine, lumefantrine, mefloquine, primaquine, proguanil, and quinine, as well as antituberculosis drugs bedaquiline and delamanid and thus need to be investigated further.
Several studies that have investigated the mechanism by which some CYP enzyme SNPs influence enzyme activity were covered in this review. These studies point to structural changes due to mutations as a major cause of the observed differences in the metabolism of these enzymes. Detailed structural studies will be required to fully elucidate the mechanisms by which these SNPs alter enzyme activity—with implications on precision medicine—especially as new mutants and phenotypic data are becoming available.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24043383/s1.

Author Contributions

Conceptualization, Ö.T.B.; supervision, Ö.T.B.; writing—original draft, C.R.C., W.V. and R.B.T.; writing—review and editing, C.R.C., W.V., R.B.T., B.S. and Ö.T.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Grants, Innovation and Product Development unit of the South African Medical Research Council with funds received from Novartis and GSK R&D (Grant # GSKNVS1/202101/002) under the Project Africa Genomic Research Approach for Diversity and Optimizing Therapeutics (GRADIENT) program. R.T.B is funded by GRADIENT as a postdoctoral fellow. C.R.C and W.V are funded as a PhD and postdoctoral fellow, respectively, by the National Research Foundation of South Africa (Grant Numbers: 130411 and 138566) and GRADIENT. The contents of this publication are solely the responsibility of the authors and do not necessarily represent the official views of the funders.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors appreciate Rhodes University, South Africa, from where this work was conducted.

Conflicts of Interest

The authors declare no conflict of interest exists regarding the publication of this paper.

References

  1. Zanger, U.M.; Schwab, M. Cytochrome P450 enzymes in drug metabolism: Regulation of gene expression, enzyme activities, and impact of genetic variation. Pharmacol. Ther. 2013, 138, 103–141. [Google Scholar] [CrossRef] [PubMed]
  2. Sarkar, S.; Ganguly, A. Current Overview of Anti-Tuberculosis Drugs: Metabolism and Toxicities. Mycobact. Dis. 2016, 6, 1000209. [Google Scholar] [CrossRef]
  3. Liu, X.; Ren, S.; Zhang, J.; Xu, D.; Jiang, F.; Jiang, P.; Feng, J.; Deng, F. The association between cytochrome P450 polymorphisms and anti-tuberculosis drug-induced liver injury: A systematic review and meta-analysis. Ann. Palliat. Med. 2021, 10, 6518–6534. [Google Scholar] [CrossRef] [PubMed]
  4. Ahmed, S.; Zhou, Z.; Zhou, J.; Chen, S.-Q. Pharmacogenomics of Drug Metabolizing Enzymes and Transporters: Relevance to Precision Medicine. Genom. Proteom. Bioinform. 2016, 14, 298–313. [Google Scholar] [CrossRef]
  5. Hassan, R.; Allali, I.; Agamah, F.E.; Elsheikh, S.S.M.; Thomford, N.E.; Dandara, C.; Chimusa, E.R. Drug response in association with pharmacogenomics and pharmacomicrobiomics: Towards a better personalized medicine. Brief. Bioinform. 2020, 22, bbaa292. [Google Scholar] [CrossRef]
  6. Johnson, J.A. Ethnic Differences in Cardiovascular Drug Response. Circulation 2008, 118, 1383–1393. [Google Scholar] [CrossRef]
  7. Wilke, R.A.; Lin, D.W.; Roden, D.M.; Watkins, P.B.; Flockhart, D.; Zineh, I.; Giacomini, K.M.; Krauss, R.M. Identifying genetic risk factors for serious adverse drug reactions: Current progress and challenges. Nat. Rev. Drug Discov. 2007, 6, 904–916. [Google Scholar] [CrossRef]
  8. Bergström, A.; McCarthy, S.A.; Hui, R.; Almarri, M.A.; Ayub, Q.; Danecek, P.; Chen, Y.; Felkel, S.; Hallast, P.; Kamm, J.; et al. Insights into human genetic variation and population history from 929 diverse genomes. Science 2020, 367, eaay5012. [Google Scholar] [CrossRef]
  9. Schärfe, C.P.I.; Tremmel, R.; Schwab, M.; Kohlbacher, O.; Marks, D.S. Genetic variation in human drug-related genes. Genome Med. 2017, 9, 117. [Google Scholar] [CrossRef]
  10. Hoehe, M.R.; Kroslak, T. Genetic variation and pharmacogenomics: Concepts, facts, and challenges. Dialog. Clin. Neurosci. 2004, 6, 5–26. [Google Scholar] [CrossRef]
  11. Weinshilboum, R. Inheritance and Drug Response. N. Engl. J. Med. 2003, 348, 529–537. [Google Scholar] [CrossRef]
  12. Katara, P.; Yadav, A. Pharmacogenes (PGx-genes): Current understanding and future directions. Gene 2019, 718, 144050. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, J.; Kang, R.; Tang, D. The KRAS-G12C inhibitor: Activity and resistance. Cancer Gene Ther. 2021, 29, 875–878. [Google Scholar] [CrossRef] [PubMed]
  14. Benet, L.Z. The Drug Transporter-Metabolism Alliance: Uncovering and Defining the Interplay. Mol. Pharm. 2009, 6, 1631–1643. [Google Scholar] [CrossRef] [PubMed]
  15. Henriques, B.C.; Yang, E.H.; Lapetina, D.; Carr, M.S.; Yavorskyy, V.; Hague, J.; Aitchison, K.J. How Can Drug Metabolism and Transporter Genetics Inform Psychotropic Prescribing? Front. Genet. 2020, 11. [Google Scholar] [CrossRef]
  16. Petzinger, E.; Geyer, J. Drug transporters in pharmacokinetics. Naunyn. Schmiedeberg. Arch. Pharmacol. 2006, 372, 465–475. [Google Scholar] [CrossRef]
  17. Lynch, T.; Price, A. The Effect of Cytochrome P450 Metabolism on Drug Response, Interactions, and Adverse Effects. Am. Fam. Physician 2007, 76, 391–396. [Google Scholar]
  18. Gaedigk, A.; Ingelman-Sundberg, M.; Miller, N.A.; Leeder, J.S.; Whirl-Carrillo, M.; Klein, T.E.; PharmVar Steering Committee. The Pharmacogene Variation (PharmVar) Consortium: Incorporation of the Human Cytochrome P450 (CYP) Allele Nomenclature Database. Clin. Pharmacol. Ther. 2018, 103, 399–401. [Google Scholar] [CrossRef]
  19. Preissner, S.C.; Hoffmann, M.F.; Preissner, R.; Dunkel, M.; Gewiess, A.; Preissner, S. Polymorphic Cytochrome P450 Enzymes (CYPs) and Their Role in Personalized Therapy. PLoS ONE 2013, 8, e82562. [Google Scholar] [CrossRef]
  20. Rajman, I.; Knapp, L.; Morgan, T.; Masimirembwa, C. African Genetic Diversity: Implications for Cytochrome P450-mediated Drug Metabolism and Drug Development. eBioMedicine 2017, 17, 67–74. [Google Scholar] [CrossRef]
  21. Marwa, K.J.; Schmidt, T.; Sjögren, M.; Minzi, O.M.; Kamugisha, E.; Swedberg, G. Cytochrome P450 single nucleotide polymorphisms in an indigenous Tanzanian population: A concern about the metabolism of artemisinin-based combinations. Malar. J. 2014, 13, 420. [Google Scholar] [CrossRef] [Green Version]
  22. World Health Organization. World Malaria Report 2020: 20 Years of Global Progress and Challenges; World Health Organization: Geneva, Switzerland, 2020; Available online: https://reliefweb.int/attachments/5adfc2b9-0e73-3c8b-97cd-c197b6864385/WMR-2020-v5-double-embargoed.pdf (accessed on 17 November 2022).
  23. Meunier, B.; de Visser, S.P.; Shaik, S. Mechanism of Oxidation Reactions Catalyzed by Cytochrome P450 Enzymes. Chem. Rev. 2004, 104, 3947–3980. [Google Scholar] [CrossRef] [PubMed]
  24. Ingelman-Sundberg, M. Human drug metabolising cytochrome P450 enzymes: Properties and polymorphisms. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2004, 369, 89–104. [Google Scholar] [CrossRef]
  25. McDonnell, A.M.; Dang, C.H. Basic Review of the Cytochrome P450 System. J. Adv. Pract. Oncol. 2013, 4, 263–268. [Google Scholar] [CrossRef] [PubMed]
  26. Goh, L.L.; Lim, C.W.; Sim, W.C.; Toh, L.X.; Leong, K.P. Analysis of Genetic Variation in CYP450 Genes for Clinical Implementation. PLoS ONE 2017, 12, e0169233. [Google Scholar] [CrossRef] [PubMed]
  27. Ohkawa, H.; Inui, H. Metabolism of agrochemicals and related environmental chemicals based on cytochrome P450s in mammals and plants. Pest Manag. Sci. 2014, 71, 824–828. [Google Scholar] [CrossRef]
  28. Zhao, M.; Ma, J.; Li, M.; Zhang, Y.; Jiang, B.; Zhao, X.; Huai, C.; Shen, L.; Zhang, N.; He, L.; et al. Cytochrome P450 Enzymes and Drug Metabolism in Humans. Int. J. Mol. Sci. 2021, 22, 12808. [Google Scholar] [CrossRef]
  29. Nebert, D.W.; Adesnik, M.; Coon, M.J.; Estabrook, R.W.; Gonzalez, F.J.; Guengerich, F.P.; Gunsalus, I.C.; Johnson, E.F. The P450 Gene Superfamily: Recommended Nomenclature; Mary Ann Liebert, Inc.: Larchmont, NY, USA, 1987. [Google Scholar]
  30. Nebert, D.W.; Nelson, D.R.; Adesnik, M.; Coon, M.J.; Estabrook, R.W.; Gonzalez, F.J.; Guengerich, F.P.; Gunsalus, I.C.; Johnson, E.F.; Kemper, B.; et al. The P450 Superfamily: Updated Listing of All Genes and Recommended Nomenclature for the Chromosomal Loci; Mary Ann Liebert, Inc.: Larchmont, NY, USA, 1989. [Google Scholar]
  31. Nebert, D.W.; Nelson, D.R.; Coon, M.J.; Estabrook, R.W.; Feyereisen, R.; Fujii-Kuriyama, Y.; Gonzalez, F.J.; Guengerich, F.P.; Gunsalus, I.C.; Johnson, E.F.; et al. The P450 Superfamily: Update on New Sequences, Gene Mapping, and Recommended Nomenclature; Mary Ann Liebert, Inc.: Larchmont, NY, USA, 1991. [Google Scholar]
  32. Nelson, D.; Kamataki, T.; Waxman, D.; Guengerich, F.P.; Estabrook, R.W.; Feyereisen, R.; Gonzalez, F.J.; Coon, M.J.; Gunsalus, I.C.; Gotoh, O.; et al. The P450 Superfamily: Update on New Sequences, Gene Mapping, Accession Numbers, Early Trivial Names of Enzymes, and Nomenclature; Mary Ann Liebert, Inc.: Larchmont, NY, USA, 1993. [Google Scholar]
  33. Nelson, D.R.; Zeldin, D.C.; Hoffman, S.M.; Maltais, L.J.; Wain, H.M.; Nebert, D.W. Comparison of cytochrome P450 (CYP) genes from the mouse and human genomes, including nomenclature recommendations for genes, pseudogenes and alternative-splice variants. Pharmacogenetics 2004, 14, 1–18. [Google Scholar] [CrossRef]
  34. Daly, A.K.; Brockmoller, J.; Broly, F.; Eichelbaum, M.; Evans, W.; Gonzalez, F.J.; Huang, J.-D.; Idle, J.R.; Ingelman-Sundberg, M.; Ishizaki, T.; et al. Nomenclature for human CYP2D6 alleles. Pharmacogenetics 1996, 6, 193–201. [Google Scholar] [CrossRef]
  35. Sim, S.C.; Ingelman-Sundberg, M. The Human Cytochrome P450 (CYP) Allele Nomenclature website: A peer-reviewed database of CYP variants and their associated effects. Hum. Genom. 2010, 4, 278–281. [Google Scholar] [CrossRef]
  36. Sim, S.C.; Ingelman-Sundberg, M. Update on Allele Nomenclature for Human Cytochromes P450 and the Human Cytochrome P450 Allele (CYP-Allele) Nomenclature Database. Methods Mol. Biol. 2013, 987, 251–259. [Google Scholar] [CrossRef]
  37. Danielson, P.B. The Cytochrome P450 Superfamily: Biochemistry, Evolution and Drug Metabolism in Humans. Curr. Drug Metab. 2002, 3, 561–597. [Google Scholar] [CrossRef] [PubMed]
  38. Esteves, F.; Rueff, J.; Kranendonk, M. The Central Role of Cytochrome P450 in Xenobiotic Metabolism—A Brief Review on a Fascinating Enzyme Family. J. Xenobiotics 2021, 11, 94–114. [Google Scholar] [CrossRef] [PubMed]
  39. Nguyen, K.-T.; Nguyen, N.-L.; Tung, N.; Nguyen, H.; Milhim, M.; Le, T.-T.; Lai, T.-H.; Phan, T.-T.; Bernhardt, R. A Novel Thermostable Cytochrome P450 from Sequence-Based Metagenomics of Binh Chau Hot Spring as a Promising Catalyst for Testosterone Conversion. Catalysts 2020, 10, 1083. [Google Scholar] [CrossRef]
  40. Sarparast, M.; Dattmore, D.; Alan, J.; Lee, K.S.S. Cytochrome P450 Metabolism of Polyunsaturated Fatty Acids and Neurodegeneration. Nutrients 2020, 12, 3523. [Google Scholar] [CrossRef] [PubMed]
  41. Machalz, D.; Pach, S.; Bermudez, M.; Bureik, M.; Wolber, G. Structural insights into understudied human cytochrome P450 enzymes. Drug Discov. Today 2021, 26, 2456–2464. [Google Scholar] [CrossRef]
  42. Mustafa, G.; Nandekar, P.P.; Mukherjee, G.; Bruce, N.J.; Wade, R.C. The Effect of Force-Field Parameters on Cytochrome P450-Membrane Interactions: Structure and Dynamics. Sci. Rep. 2020, 10, 1–11. [Google Scholar] [CrossRef]
  43. Poulos, T.L.; Johnson, E.F. Structures of Cytochrome P450 Enzymes. Cytochrome P450: Structure, Mechanism, and Biochemistry; Springer: Cham, Switzerland, 2015; pp. 3–32. [Google Scholar] [CrossRef]
  44. Guengerich, F.P. Cytochrome P450 and Chemical Toxicology. Chem. Res. Toxicol. 2008, 21, 70–83. [Google Scholar] [CrossRef] [PubMed]
  45. Textbook of Biochemistry with Clinical Correlations, 7th Edition|Wiley. In: Wiley.com. Available online: https://www.wiley.com/en-us/Textbook+of+Biochemistry+with+Clinical+Correlations%2C+7th+Edition-p-9780470281734 (accessed on 26 August 2022).
  46. Midlik, A.; Navrátilová, V.; Moturu, T.R.; Koča, J.; Svobodová, R.; Berka, K. Uncovering of cytochrome P450 anatomy by SecStrAnnotator. Sci. Rep. 2021, 11, 1–12. [Google Scholar] [CrossRef] [PubMed]
  47. Williams, P.A.; Cosme, J.; Vinković, D.M.; Ward, A.; Angove, H.C.; Day, P.J.; Vonrhein, C.; Tickle, I.J.; Jhoti, H. Crystal Structures of Human Cytochrome P450 3A4 Bound to Metyrapone and Progesterone. Science 2004, 305, 683–686. [Google Scholar] [CrossRef]
  48. Munro, A.W.; Girvan, H.M.; Mason, A.E.; Dunford, A.J.; McLean, K.J. What makes a P450 tick? Trends Biochem. Sci. 2013, 38, 140–150. [Google Scholar] [CrossRef] [PubMed]
  49. Fishelovitch, D.; Shaik, S.; Wolfson, H.J.; Nussinov, R. Theoretical Characterization of Substrate Access/Exit Channels in the Human Cytochrome P450 3A4 Enzyme: Involvement of Phenylalanine Residues in the Gating Mechanism. J. Phys. Chem. B 2009, 113, 13018–13025. [Google Scholar] [CrossRef] [PubMed]
  50. Cojocaru, V.; Winn, P.J.; Wade, R.C. The ins and outs of cytochrome P450s. Biochim. Biophys. Acta (BBA) 2007, 1770, 390–401. [Google Scholar] [CrossRef] [PubMed]
  51. Dubey, K.D.; Shaik, S. Cytochrome P450—The Wonderful Nanomachine Revealed through Dynamic Simulations of the Catalytic Cycle. Acc. Chem. Res. 2019, 52, 389–399. [Google Scholar] [CrossRef]
  52. Hu, Q.; Hartmann, R.W. The Renaissance of CYP17 Inhibitors for the Treatment of Prostate Cancer. In Cancer Drug Design and Discovery; Elsevier: Amsterdam, The Netherlands, 2014; pp. 319–356. [Google Scholar]
  53. Denisov, I.G.; Makris, T.M.; Sligar, S.G.; Schlichting, I. Structure and Chemistry of Cytochrome P450. Chem. Rev. 2005, 105, 2253–2278. [Google Scholar] [CrossRef]
  54. Coleman, T.; Stok, J.E.; Podgorski, M.N.; Bruning, J.B.; De Voss, J.J.; Bell, S.G. Structural insights into the role of the acid-alcohol pair of residues required for dioxygen activation in cytochrome P450 enzymes. JBIC J. Biol. Inorg. Chem. 2020, 25, 583–596. [Google Scholar] [CrossRef]
  55. Wang, B.; Li, C.; Dubey, K.D.; Shaik, S. Quantum Mechanical/Molecular Mechanical Calculated Reactivity Networks Reveal How Cytochrome P450cam and Its T252A Mutant Select Their Oxidation Pathways. J. Am. Chem. Soc. 2015, 137, 7379–7390. [Google Scholar] [CrossRef]
  56. Wu, J.; Guan, X.; Dai, Z.; He, R.; Ding, X.; Yang, L.; Ge, G. Molecular probes for human cytochrome P450 enzymes: Recent progress and future perspectives. Coord. Chem. Rev. 2020, 427, 213600. [Google Scholar] [CrossRef]
  57. Ueyama, N.; Nishikawa, N.; Yamada, Y.; Okamura, T.-A.; Nakamura, A. Cytochrome P-450 Model (Porphinato)(thiolato)iron(III) Complexes with Single and Double NH···S Hydrogen Bonds at the Thiolate Site. J. Am. Chem. Soc. 1996, 118, 12826–12827. [Google Scholar] [CrossRef]
  58. Ueyama, N.; Terakawa, T.; Nakata, M.; Nakamura, A. Positive shift of redox potential of [Fe4S4(Z-cys-Gly-Ala-OMe)4]2- in dichloromethane. J. Am. Chem. Soc. 1983, 105, 7098–7102. [Google Scholar] [CrossRef]
  59. Matsunaga, I.; Sumimoto, T.; Ueda, A.; Kusunose, E.; Ichihara, K. Fatty acid-specific, regiospecific, and stereospecific hydroxylation by cytochrome P450 (CYP152B1) from Sphingomonas paucimobilis: Substrate structure required for α-hydroxylation. Lipids 2000, 35, 365–371. [Google Scholar] [CrossRef]
  60. Lee, D.-S.; Yamada, A.; Sugimoto, H.; Matsunaga, I.; Ogura, H.; Ichihara, K.; Adachi, S.-I.; Park, S.-Y.; Shiro, Y. Substrate Recognition and Molecular Mechanism of Fatty Acid Hydroxylation by Cytochrome P450 from Bacillus subtilis. J. Biol. Chem. 2003, 278, 9761–9767. [Google Scholar] [CrossRef]
  61. Rendic, S.; Guengerich, F.P. Survey of human oxidoreductases and cytochrome P450 enzymes involved in the metabolism of xenobiotic and natural chemicals. Chem. Res. Toxicol. 2015, 28, 38–42. [Google Scholar] [CrossRef] [PubMed]
  62. Iyanagi, T. Molecular Mechanism of Phase I and Phase II Drug-Metabolizing Enzymes: Implications for Detoxification. Int. Rev. Cytol. 2007, 260, 35–112. [Google Scholar] [CrossRef] [PubMed]
  63. Vasav, A.P.; Barvkar, V.T. Phylogenomic analysis of cytochrome P450 multigene family and their differential expression analysis in Solanum lycopersicum L. suggested tissue specific promoters. BMC Genom. 2019, 20, 116. [Google Scholar] [CrossRef] [PubMed]
  64. Urlacher, V.B.; Girhard, M. Cytochrome P450 monooxygenases: An update on perspectives for synthetic application. Trends Biotechnol. 2012, 30, 26–36. [Google Scholar] [CrossRef]
  65. Van Bogaert, I.N.A.; Groeneboer, S.; Saerens, K.; Soetaert, W. The role of cytochrome P450 monooxygenases in microbial fatty acid metabolism. FEBS J. 2010, 278, 206–221. [Google Scholar] [CrossRef]
  66. Henry, H.L. Cytochrome P450 Structure, Mechanism, and Biochemistry, 2nd ed.; Edited by Paul R. Oritz de Montellano (University of California, San Francisco). Plenum: New York. 1995. xi + 631 pp. $125.00. ISBN 0-306-45141-7. J. Am. Chem. Soc. 1996, 118, 10945. [Google Scholar] [CrossRef]
  67. Sono, M.; Roach, M.P.; Coulter, E.D.; Dawson, J.H. Heme-Containing Oxygenases. Chem. Rev. 1996, 96, 2841–2888. [Google Scholar] [CrossRef]
  68. Poulos, T.L. Heme Enzyme Structure and Function. Chem. Rev. 2014, 114, 3919–3962. [Google Scholar] [CrossRef] [Green Version]
  69. Dawson, J.H.; Sono, M. Cytochrome P-450 and chloroperoxidase: Thiolate-ligated heme enzymes. Spectroscopic determination of their active-site structures and mechanistic implications of thiolate ligation. Chem. Rev. 1987, 87, 1255–1276. [Google Scholar] [CrossRef]
  70. de Montellano, P.R.O.; De Voss, J.J. Oxidizing species in the mechanism of cytochrome P450. Nat. Prod. Rep. 2002, 19, 477–493. [Google Scholar] [CrossRef]
  71. Hoa, G.H.B.; McLean, M.; Sligar, S.G. High pressure, a tool for exploring heme protein active sites. Biochim. Biophys. Acta (BBA) Protein Struct. Mol. Enzym. 2002, 1595, 297–308. [Google Scholar] [CrossRef]
  72. Guengerich, F.P.; Macdonald, T.L. Mechanisms of cytochrome P-450 catalysis. FASEB J. 1990, 4, 2453–2459. [Google Scholar] [CrossRef]
  73. Shaik, S.; Kumar, D.; de Visser, S.P.; Altun, A.; Thiel, W. Theoretical Perspective on the Structure and Mechanism of Cytochrome P450 Enzymes. Chem. Rev. 2005, 105, 2279–2328. [Google Scholar] [CrossRef]
  74. Shaik, S.; Cohen, S.; Wang, Y.; Chen, H.; Kumar, D.; Thiel, W. P450 Enzymes: Their Structure, Reactivity, and Selectivity—Modeled by QM/MM Calculations. Chem. Rev. 2009, 110, 949–1017. [Google Scholar] [CrossRef] [PubMed]
  75. Hayaishi, O. An odyssey with oxygen. Biochem. Biophys. Res. Commun. 2005, 338, 2–6. [Google Scholar] [CrossRef]
  76. Waterman, M.R. Professor Howard Mason and oxygen activation. Biochem. Biophys. Res. Commun. 2005, 338, 7–11. [Google Scholar] [CrossRef]
  77. Mowat, C.G.; Gazur, B.; Campbell, L.P.; Chapman, S.K. Flavin-containing heme enzymes. Arch. Biochem. Biophys. 2009, 493, 37–52. [Google Scholar] [CrossRef]
  78. Guengerich, F.P. Mechanisms of cytochrome P450 substrate oxidation: Mini Review. J. Biochem. Mol. Toxicol. 2007, 21, 163–168. [Google Scholar] [CrossRef]
  79. Zheng, J.; Altun, A.; Thiel, W. Common system setup for the entire catalytic cycle of cytochrome P450cam in quantum mechanical/molecular mechanical studies. J. Comput. Chem. 2007, 28, 2147–2158. [Google Scholar] [CrossRef] [PubMed]
  80. Sligar, S.G.; Gunsalus, I.C. A thermodynamic model of regulation: Modulation of redox equilibria in camphor monoxygenase. Proc. Natl. Acad. Sci. USA 1976, 73, 1078–1082. [Google Scholar] [CrossRef]
  81. Govindaraj, S.; Poulos, T.L. The Domain Architecture of Cytochrome P450BM-3. J. Biol. Chem. 1997, 272, 7915–7921. [Google Scholar] [CrossRef] [PubMed]
  82. Volz, T.J.; Rock, D.A.; Jones, J.P. Evidence for Two Different Active Oxygen Species in Cytochrome P450 BM3 Mediated Sulfoxidation and N-Dealkylation Reactions. J. Am. Chem. Soc. 2002, 124, 9724–9725. [Google Scholar] [CrossRef] [PubMed]
  83. Yu, X.; Cojocaru, V.; Wade, R.C. Conformational diversity and ligand tunnels of mammalian cytochrome P450s. Biotechnol. Appl. Biochem. 2013, 60, 134–145. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, A.; Stout, C.; Zhang, Q.; Johnson, E.F. Contributions of Ionic Interactions and Protein Dynamics to Cytochrome P450 2D6 (CYP2D6) Substrate and Inhibitor Binding. J. Biol. Chem. 2015, 290, 5092–5104. [Google Scholar] [CrossRef]
  85. Dubey, K.D.; Shaik, S. Choreography of the Reductase and P450BM3 Domains toward Electron Transfer Is Instigated by the Substrate. J. Am. Chem. Soc. 2018, 140, 683–690. [Google Scholar] [CrossRef]
  86. Dubey, K.D.; Wang, B.; Vajpai, M.; Shaik, S. MD simulations and QM/MM calculations show that single-site mutations of cytochrome P450BM3 alter the active site’s complexity and the chemoselectivity of oxidation without changing the active species. Chem. Sci. 2017, 8, 5335–5344. [Google Scholar] [CrossRef]
  87. Dubey, K.D.; Wang, B.; Shaik, S. Molecular Dynamics and QM/MM Calculations Predict the Substrate-Induced Gating of Cytochrome P450 BM3 and the Regio- and Stereoselectivity of Fatty Acid Hydroxylation. J. Am. Chem. Soc. 2016, 138, 837–845. [Google Scholar] [CrossRef]
  88. Ramanan, R.; Dubey, K.D.; Wang, B.; Mandal, D.; Shaik, S. Emergence of Function in P450-Proteins: A Combined Quantum Mechanical/Molecular Mechanical and Molecular Dynamics Study of the Reactive Species in the H2O2-Dependent Cytochrome P450SPα and Its Regio and Enantioselective Hydroxylation of Fatty Acids. J. Am. Chem. Soc. 2016, 138, 6786–6797. [Google Scholar] [CrossRef]
  89. Berman, H.M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.N.; Weissig, H.; Shindyalov, I.N.; Bourne, P.E. The Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, A.; Savas, U.; Stout, C.D.; Johnson, E.F. Structural Characterization of the Complex between α-Naphthoflavone and Human Cytochrome P450 1B1. J. Biol. Chem. 2011, 286, 5736–5743. [Google Scholar] [CrossRef]
  91. Urban, P.; Lautier, T.; Pompon, D.; Truan, G. Ligand Access Channels in Cytochrome P450 Enzymes: A Review. Int. J. Mol. Sci. 2018, 19, 1617. [Google Scholar] [CrossRef] [PubMed]
  92. Walsh, A.A.; Szklarz, G.D.; Scott, E.E. Human Cytochrome P450 1A1 Structure and Utility in Understanding Drug and Xenobiotic Metabolism*. J. Biol. Chem. 2013, 288, 12932–12943. [Google Scholar] [CrossRef]
  93. McLaughlin, L.A.; Paine, M.J.I.; Kemp, C.A.; Maréchal, J.-D.; Flanagan, J.U.; Ward, C.J.; Sutcliffe, M.; Roberts, G.; Wolf, C.R. Why Is Quinidine an Inhibitor of Cytochrome P450 2D6? J. Biol. Chem. 2005, 280, 38617–38624. [Google Scholar] [CrossRef] [PubMed]
  94. Hutzler, J.M.; Walker, G.S.; Wienkers, L.C. Inhibition of Cytochrome P450 2D6: Structure−Activity Studies Using a Series of Quinidine and Quinine Analogues. Chem. Res. Toxicol. 2003, 16, 450–459. [Google Scholar] [CrossRef]
  95. Otton, S.; Inaba, T.; Kalow, W. Competitive inhibition of sparteine oxidation in human liver by β-adrenoceptor antagonists and other cardiovascular drugs. Life Sci. 1984, 34, 73–80. [Google Scholar] [CrossRef]
  96. Sansen, S.; Yano, J.K.; Reynald, R.L.; Schoch, G.A.; Griffin, K.J.; Stout, C.D.; Johnson, E.F. Adaptations for the Oxidation of Polycyclic Aromatic Hydrocarbons Exhibited by the Structure of Human P450 1A2. J. Biol. Chem. 2007, 282, 14348–14355. [Google Scholar] [CrossRef]
  97. Huttunen, K.M.; Mahonen, N.; Raunio, H.; Rautio, J. Cytochrome P450-Activated Prodrugs: Targeted Drug Delivery. Curr. Med. Chem. 2008, 15, 2346–2365. [Google Scholar] [CrossRef]
  98. Yano, J.K.; Hsu, M.-H.; Griffin, K.J.; Stout, C.D.; Johnson, E.F. Structures of human microsomal cytochrome P450 2A6 complexed with coumarin and methoxsalen. Nat. Struct. Mol. Biol. 2005, 12, 822–823. [Google Scholar] [CrossRef]
  99. Williams, P.A.; Cosme, J.; Ward, A.; Angove, H.C.; Vinković, D.M.; Jhoti, H. Crystal structure of human cytochrome P450 2C9 with bound warfarin. Nature 2003, 424, 464–468. [Google Scholar] [CrossRef]
  100. Wester, M.R.; Yano, J.K.; Schoch, G.A.; Yang, C.; Griffin, K.J.; Stout, C.D.; Johnson, E.F. The Structure of Human Cytochrome P450 2C9 Complexed with Flurbiprofen at 2.0-Å Resolution. J. Biol. Chem. 2004, 279, 35630–35637. [Google Scholar] [CrossRef] [PubMed]
  101. Dong, D.; Wu, B.; Chow, D.; Hu, M. Substrate selectivity of drug-metabolizing cytochrome P450s predicted from crystal structures and in silico modeling. Drug Metab. Rev. 2012, 44, 192–208. [Google Scholar] [CrossRef] [PubMed]
  102. Porubsky, P.R.; Meneely, K.M.; Scott, E.E. Structures of Human Cytochrome P-450 2E1. J. Biol. Chem. 2008, 283, 33698–33707. [Google Scholar] [CrossRef]
  103. Gay, S.C.; Roberts, A.G.; Halpert, J.R. Structural features of cytochromes P450 and ligands that affect drug metabolism as revealed by x-ray crystallography and NMR. Future Med. Chem. 2010, 2, 1451–1468. [Google Scholar] [CrossRef] [PubMed]
  104. Yano, J.K.; Wester, M.R.; Schoch, G.A.; Griffin, K.J.; Stout, C.D.; Johnson, E.F. The Structure of Human Microsomal Cytochrome P450 3A4 Determined by X-ray Crystallography to 2.05-Å Resolution. J. Biol. Chem. 2004, 279, 38091–38094. [Google Scholar] [CrossRef]
  105. Ekroos, M.; Sjögren, T. Structural basis for ligand promiscuity in cytochrome P450 3A4. Proc. Natl. Acad. Sci. USA 2006, 103, 13682–13687. [Google Scholar] [CrossRef]
  106. Sevrioukova, I.F.; Poulos, T.L. Interaction of human cytochrome P4503A4 with ritonavir analogs. Arch. Biochem. Biophys. 2012, 520, 108–116. [Google Scholar] [CrossRef]
  107. Hsu, M.-H.; Savas, U.; Johnson, E.F. The X-Ray Crystal Structure of the Human Mono-Oxygenase Cytochrome P450 3A5-Ritonavir Complex Reveals Active Site Differences between P450s 3A4 and 3A5. Mol. Pharmacol. 2017, 93, 14–24. [Google Scholar] [CrossRef]
  108. World Health Organization. WHO Guidelines for Malaria, 3 June 2022; World Health Organization: Geneva, Switzerland, 2022; (WHO/UCN/GMP/2022.01 Rev.2); Available online: https://apps.who.int/iris/rest/bitstreams/1427681/retrieve (accessed on 17 November 2022).
  109. AlKadi, H.O. Antimalarial Drug Toxicity: A Review. Chemotherapy 2007, 53, 385–391. [Google Scholar] [CrossRef]
  110. Taylor, W.R.J.; White, N.J. Antimalarial Drug Toxicity. Drug Saf. 2004, 27, 25–61. [Google Scholar] [CrossRef] [PubMed]
  111. Kerb, R.; Fux, R.; Mörike, K.; Kremsner, P.G.; Gil, J.P.; Gleiter, C.H.; Schwab, M. Pharmacogenetics of antimalarial drugs: Effect on metabolism and transport. Lancet Infect. Dis. 2009, 9, 760–774. [Google Scholar] [CrossRef]
  112. Khoo, S.; Back, D.; Winstanley, P. The potential for interactions between antimalarial and antiretroviral drugs. Aids 2005, 19, 995–1005. [Google Scholar] [CrossRef] [PubMed]
  113. Rendic, S.; Guengerich, S.P.R.A.F.P. Metabolism and Interactions of Chloroquine and Hydroxychloroquine with Human Cytochrome P450 Enzymes and Drug Transporters. Curr. Drug Metab. 2020, 21, 1127–1135. [Google Scholar] [CrossRef]
  114. Hodel, E.M.S.; Csajka, C.; Ariey, F.; Guidi, M.; Kabanywanyi, A.M.; Duong, S.; Decosterd, L.A.; Olliaro, P.; Beck, H.-P.; Genton, B. Effect of Single Nucleotide Polymorphisms in Cytochrome P450 Isoenzyme and N-Acetyltransferase 2 Genes on the Metabolism of Artemisinin-Based Combination Therapies in Malaria Patients from Cambodia and Tanzania. Antimicrob. Agents Chemother. 2013, 57, 950–958. [Google Scholar] [CrossRef]
  115. Giao, P.; De Vries, P.J. Pharmacokinetic Interactions of Antimalarial Agents. Clin. Pharmacokinet. 2001, 40, 343–373. [Google Scholar] [CrossRef] [PubMed]
  116. Li, X.-Q.; Björkman, A.; Andersson, T.B.; Ridderström, M.; Masimirembwa, C.M. Amodiaquine Clearance and Its Metabolism to N-Desethylamodiaquine Is Mediated by CYP2C8: A New High Affinity and Turnover Enzyme-Specific Probe Substrate. Experiment 2002, 300, 399–407. [Google Scholar] [CrossRef]
  117. Zuidema, J.; Hilbers-Modderman, E.S.M.; Merkus, F.W.H.M. Clinical Pharmacokinetics of Dapsone. Clin. Pharmacokinet. 1986, 11, 299–315. [Google Scholar] [CrossRef] [PubMed]
  118. Zang, M.; Zhu, F.; Li, X.; Yang, A.; Xing, J. Auto-induction of phase I and phase II metabolism of artemisinin in healthy Chinese subjects after oral administration of a new artemisinin-piperaquine fixed combination. Malar. J. 2014, 13, 214. [Google Scholar] [CrossRef]
  119. Ganesan, S.; Sahu, R.; A Walker, L.; Tekwani, B.L. Cytochrome P450-dependent toxicity of dapsone in human erythrocytes. J. Appl. Toxicol. 2009, 30, 271–275. [Google Scholar] [CrossRef]
  120. Grace, J.M.; Skanchy, D.J.; Aguilar, A.J. Metabolism of artelinic acid to dihydroqinghaosu by human liver cytochrome P4503A. Xenobiotica 1999, 29, 703–717. [Google Scholar] [CrossRef]
  121. Mehlotra, R.K.; Henry-Halldin, C.N.; A Zimmerman, P. Application of pharmacogenomics to malaria: A holistic approach for successful chemotherapy. Pharmacogenomics 2009, 10, 435–449. [Google Scholar] [CrossRef] [PubMed]
  122. Lefèvre, G.; Carpenter, P.; Souppart, C.; Schmidli, H.; McClean, M.; Stypinski, D. Pharmacokinetics and electrocardiographic pharmacodynamics of artemether-lumefantrine (Riamet®) with concomitant administration of ketoconazole in healthy subjects. Br. J. Clin. Pharmacol. 2002, 54, 485–492. [Google Scholar] [CrossRef]
  123. Van Agtmael, M.A.; Gupta, V.; Van Der Wösten, T.H.; Rutten, J.-P.B.; Van Boxtel, C.J. Grapefruit juice increases the bioavailability of artemether. Eur. J. Clin. Pharmacol. 1999, 55, 405–410. [Google Scholar] [CrossRef] [PubMed]
  124. Navaratnam, V.; Mansor, S.M.; Sit, N.W.; Grace, J.; Li, Q.; Olliaro, P. Pharmacokinetics of Artemisinin-Type Compounds. Clin. Pharmacokinet. 2000, 39, 255–270. [Google Scholar] [CrossRef]
  125. Svensson, U.S.; Ashton, M. Identification of the human cytochrome P450 enzymes involved in the in vitro metabolism of artemisinin. Br. J. Clin. Pharmacol. 1999, 48, 528–535. [Google Scholar] [CrossRef]
  126. Woodrow, C.J.; Haynes, R.K.; Krishna, S. Artemisinins. Postgrad. Med. J. 2005, 81, 71–78. [Google Scholar] [CrossRef]
  127. Yusof, W.; Hua, G.S. Gene, ethnic and gender influences predisposition of adverse drug reactions to artesunate among Malaysians. Toxicol. Mech. Methods 2011, 22, 184–192. [Google Scholar] [CrossRef] [PubMed]
  128. Sortica, V.A.; Lindenau, J.D.; Cunha, M.G.; Ohnishi, M.D.; Ventura, A.M.R.; Ribeiro-Dos-Santos, K.; Santos, S.E.; Guimarães, L.S.; Hutz, M.H. The effect of SNPs in CYP450 in chloroquine/primaquine Plasmodium vivax malaria treatment. Pharmacogenomics 2016, 17, 1903–1911. [Google Scholar] [CrossRef]
  129. Projean, D.; Baune, B.; Farinotti, R.; Flinois, J.-P.; Beaune, P.; Taburet, A.-M.; Ducharme, J. In vitro metabolism of chloroquine: Identification of cyp2c8, cyp3a4, and cyp2d6 as the main isoforms catalyzing n-desethylchloroquine formation. Drug Metab. Dispos. 2003, 31, 748–754. [Google Scholar] [CrossRef] [Green Version]
  130. Winte, H.R.; Wang, Y.; Unadkat, J.D. CYP2C8/9 Mediate Dapsone N-Hydroxylation at Clinical Concentrations of Dapsone. Drug Metab. Dispos. 2000, 28, 865–868. [Google Scholar]
  131. May, D.G.; Porter, J.; Wilkinson, G.R.; A Branch, R. Frequency distribution of dapsone N-hydroxylase, a putative probe for P4503A4 activity, in a white population. Clin. Pharmacol. Ther. 1994, 55, 492–500. [Google Scholar] [CrossRef]
  132. A Ward, S.; Sevene, E.J.; Hastings, I.M.; Nosten, F.; McGready, R. Antimalarial drugs and pregnancy: Safety, pharmacokinetics, and pharmacovigilance. Lancet Infect. Dis. 2007, 7, 136–144. [Google Scholar] [CrossRef]
  133. Pernaute-Lau, L.; Camara, M.; de Sousa, T.N.; Morris, U.; Ferreira, M.U.; Gil, J.P. An update on pharmacogenetic factors influencing the metabolism and toxicity of artemisinin-based combination therapy in the treatment of malaria. Expert Opin. Drug Metab. Toxicol. 2022, 18, 39–59. [Google Scholar] [CrossRef] [PubMed]
  134. Fontaine, F.; de Sousa, G.; Burcham, P.; Duchêne, P.; Rahmani, R. Role of cytochrome P450 3A in the metabolism of mefloquine in human and animal hepatocytes. Life Sci. 2000, 66, 2193–2212. [Google Scholar] [CrossRef] [PubMed]
  135. Na Bangchang, K.; Karbwang, J.; Back, D.J. Mefloquine metabolism by human liver microsomes: Effect of other antimalarial drugs. Biochem. Pharmacol. 1992, 43, 1957–1961. [Google Scholar] [CrossRef]
  136. Avula, B.; Tekwani, B.L.; Chaurasiya, N.D.; Fasinu, P.; Nanayakkara, N.P.D.; Herath, H.M.T.B.; Wang, Y.-H.; Bae, J.-Y.; Khan, S.I.; ElSohly, M.A.; et al. Metabolism of primaquine in normal human volunteers: Investigation of phase I and phase II metabolites from plasma and urine using ultra-high performance liquid chromatography-quadrupole time-of-flight mass spectrometry. Malar. J. 2018, 17, 294. [Google Scholar] [CrossRef]
  137. Somogyi, A.A.; Reinhard, H.A.; Bochner, F. Pharmacokinetic evaluation of proguanil: A probe phenotyping drug for the mephenytoin hydroxylase polymorphism. Br. J. Clin. Pharmacol. 1996, 41, 175–179. [Google Scholar] [CrossRef]
  138. Zhang, H.; Coville, P.F.; Walker, R.J.; Miners, J.O.; Birkett, D.J.; Wanwimolruk, S. Evidence for involvement of human CYP3A in the 3-hydroxylation of quinine. Br. J. Clin. Pharmacol. 1997, 43, 245–252. [Google Scholar] [CrossRef]
  139. Nielsen, T.L.; Rasmussen, B.B.; Flinois, J.P.; Beaune, P.; Brosen, K. In Vitro Metabolism of Quinidine: The (3S)-3-Hydroxylation of Quinidine Is a Specific Marker Reaction for Cytochrome P-4503A4 Activity in Human Liver Microsomes. J. Pharmacol. Exp. Ther. 1999, 289, 31–37. [Google Scholar]
  140. Horsburgh, C.R.; Barry, C.E.; Lange, C. Treatment of Tuberculosis. N. Engl. J. Med. 2015, 373, 2149–2160. [Google Scholar] [CrossRef] [PubMed]
  141. Motta, I.; Calcagno, A.; Bonora, S. Pharmacokinetics and pharmacogenetics of anti-tubercular drugs: A tool for treatment optimization? Expert Opin. Drug Metab. Toxicol. 2017, 14, 59–82. [Google Scholar] [CrossRef]
  142. Barozi, V.; Musyoka, T.M.; Sheik Amamuddy, O.; Tastan Bishop, Ö. Deciphering Isoniazid Drug Resistance Mechanisms on Dimeric Mycobacterium tuberculosis KatG via Post-molecular Dynamics Analyses Including Combined Dynamic Residue Network Metrics. ACS Omega 2022, 7, 13313–13332. [Google Scholar] [CrossRef]
  143. Nath, H.; Ryoo, S. First– and Second–Line Drugs and Drug Resistance. Tuberculosis 2013. [Google Scholar] [CrossRef]
  144. World Health Organisation. Rapid Communication: Key Changes to the Treatment of Drug-Resistant Tuberculosis. 2022. Available online: https://apps.who.int/iris/handle/10665/353743 (accessed on 17 November 2022).
  145. Sutezolid. In: TB Alliance. Available online: https://www.tballiance.org/portfolio/trial/12018 (accessed on 25 November 2022).
  146. Zumla, A.; Rao, M.; Wallis, R.S.; E Kaufmann, S.H.; Rustomjee, R.; Mwaba, P.; Vilaplana, C.; Yeboah-Manu, D.; Chakaya, J.; Ippolito, G.; et al. Host-directed therapies for infectious diseases: Current status, recent progress, and future prospects. Lancet Infect. Dis. 2016, 16, e47–e63. [Google Scholar] [CrossRef] [PubMed]
  147. Magis-Escurra, C.; Boogaard, J.V.D.; Ijdema, D.; Boeree, M.; Aarnoutse, R. Therapeutic drug monitoring in the treatment of tuberculosis patients. Pulm. Pharmacol. Ther. 2012, 25, 83–86. [Google Scholar] [CrossRef] [PubMed]
  148. Babalik, A.; Mannix, S.; Francis, D.; Menzies, D. Therapeutic Drug Monitoring in the Treatment of Active Tuberculosis. Can. Respir. J. 2011, 18, 225–229. [Google Scholar] [CrossRef]
  149. Tostmann, A.; Boeree, M.J.; Aarnoutse, R.E.; De Lange, W.C.; van der Ven, A.J.; Dekhuijzen, R. Antituberculosis drug-induced hepatotoxicity: Concise up-to-date review. J. Gastroenterol. Hepatol. 2008, 23, 192–202. [Google Scholar] [CrossRef]
  150. Shao, Q.; Mao, X.; Zhou, Z.; Huai, C.; Li, Z. Research Progress of Pharmacogenomics in Drug-Induced Liver Injury. Front. Pharmacol. 2021, 12, 735260. [Google Scholar] [CrossRef]
  151. Shimada, T.; Yamazaki, H.; Mimura, M.; Inui, Y.; Guengerich, F.P. Interindividual variations in human liver cytochrome P-450 enzymes involved in the oxidation of drugs, carcinogens and toxic chemicals: Studies with liver microsomes of 30 Japanese and 30 Caucasians. J. Pharmacol. Exp. Ther. 1994, 270, 414–423. [Google Scholar]
  152. Liu, K.; Li, F.; Lu, J.; Liu, S.; Dorko, K.; Xie, W.; Ma, X. Bedaquiline Metabolism: Enzymes and Novel Metabolites. Drug Metab. Dispos. 2014, 42, 863–866. [Google Scholar] [CrossRef] [PubMed]
  153. Cuyckens, F.; Balcaen, L.I.L.; De Wolf, K.; De Samber, B.; Van Looveren, C.; Hurkmans, R.; Vanhaecke, F. Use of the bromine isotope ratio in HPLC-ICP-MS and HPLC-ESI-MS analysis of a new drug in development. Anal. Bioanal. Chem. 2008, 390, 1717–1729. [Google Scholar] [CrossRef]
  154. Meermann, B.; Bockx, M.; Laenen, A.; Van Looveren, C.; Cuyckens, F.; Vanhaecke, F. Speciation analysis of bromine-containing drug metabolites in feces samples from a human in vivo study by means of HPLC/ICP-MS combined with on-line isotope dilution. Anal. Bioanal. Chem. 2011, 402, 439–448. [Google Scholar] [CrossRef]
  155. Andries, K.; Verhasselt, P.; Guillemont, J.; Göhlmann, H.W.H.; Neefs, J.-M.; Winkler, H.; Van Gestel, J.; Timmerman, P.; Zhu, M.; Lee, E.; et al. A Diarylquinoline Drug Active on the ATP Synthase of Mycobacterium tuberculosis. Science 2005, 307, 223–227. [Google Scholar] [CrossRef] [PubMed]
  156. Upadhayaya, R.S.; Kulkarni, G.M.; Vasireddy, N.R.; Vandavasi, J.K.; Dixit, S.S.; Sharma, V.; Chattopadhyaya, J. Design, synthesis and biological evaluation of novel triazole, urea and thiourea derivatives of quinoline against Mycobacterium tuberculosis. Bioorganic Med. Chem. 2009, 17, 4681–4692. [Google Scholar] [CrossRef] [PubMed]
  157. Huitric, E.; Verhasselt, P.; Andries, K.; Hoffner, S.E. In Vitro Antimycobacterial Spectrum of a Diarylquinoline ATP Synthase Inhibitor. Antimicrob. Agents Chemother. 2007, 51, 4202–4204. [Google Scholar] [CrossRef]
  158. Obach, R.S. Linezolid Metabolism Is Catalyzed by Cytochrome P450 2J2, 4F2, and 1B1. Drug Metab. Dispos. 2022, 50, 413–421. [Google Scholar] [CrossRef]
  159. Deltyba INN-Delamanid, Summary of Product Characteristics. Available online: https://www.ema.europa.eu/en/documents/product-information/deltyba-epar-product-information_en.pdf (accessed on 28 November 2022).
  160. Ericsson, T.; Sundell, J.; Torkelsson, A.; Hoffmann, K.-J.; Ashton, M. Effects of artemisinin antimalarials on Cytochrome P450 enzymes in vitro using recombinant enzymes and human liver microsomes: Potential implications for combination therapies. Xenobiotica 2013, 44, 615–626. [Google Scholar] [CrossRef]
  161. Nishimura, Y.; Kurata, N.; Sakurai, E.; Yasuhara, H. Inhibitory Effect of Antituberculosis Drugs on Human Cytochrome P450-Mediated Activities. J. Pharmacol. Sci. 2004, 96, 293–300. [Google Scholar] [CrossRef]
  162. Ogu, C.C.; Maxa, J.L. Drug Interactions Due to Cytochrome P450. Bayl. Univ. Med. Cent. Proc. 2000, 13, 421–423. [Google Scholar] [CrossRef]
  163. Shackleford, D.M.; Chiu, F.C.K.; Katneni, K.; Blundell, S.; McLaren, J.; Wang, X.; Zhou, L.; Sriraghavan, K.; Alker, A.M.; Hunziker, D.; et al. Cytochrome P450-Mediated Metabolism and CYP Inhibition for the Synthetic Peroxide Antimalarial OZ439. ACS Infect. Dis. 2021, 7, 1885–1893. [Google Scholar] [CrossRef] [PubMed]
  164. Xing, J.; Kirby, B.J.; Whittington, D.; Wan, Y.; Goodlett, D.R. Evaluation of P450 Inhibition and Induction by Artemisinin Antimalarials in Human Liver Microsomes and Primary Human Hepatocytes. Drug Metab. Dispos. 2012, 40, 1757–1764. [Google Scholar] [CrossRef] [PubMed]
  165. Asimus, S.; Elsherbiny, D.; Hai, T.N.; Jansson, B.; Huong, N.V.; Petzold, M.G.; Simonsson, U.; Ashton, M. Artemisinin antimalarials moderately affect cytochrome P450 enzyme activity in healthy subjects. Fundam. Clin. Pharmacol. 2007, 21, 307–316. [Google Scholar] [CrossRef] [PubMed]
  166. He, F.; Bi, H.-C.; Xie, Z.-Y.; Zuo, Z.; Li, J.-K.; Li, X.; Zhao, L.-Z.; Chen, X.; Huang, M. Rapid determination of six metabolites from multiple cytochrome P450 probe substrates in human liver microsome by liquid chromatography/mass spectrometry: Application to high-throughput inhibition screening of terpenoids. Rapid Commun. Mass Spectrom. 2007, 21, 635–643. [Google Scholar] [CrossRef] [PubMed]
  167. Bapiro, T.E.; Egnell, A.-C.; Hasler, J.A.; Masimirembwa, C.M. Application of higher throughput screening (HTS) inhibition assays to evaluate the interaction of antiparasitic drugs with cytochrome P450s. Drug Metab. Dispos. 2001, 29, 30–35. [Google Scholar]
  168. Bapiro, T.E.; Sayi, J.; Hasler, J.A.; Jande, M.; Rimoy, G.; Masselle, A.; Masimirembwa, C.M. Artemisinin and thiabendazole are potent inhibitors of cytochrome P450 1A2 (CYP1A2) activity in humans. Eur. J. Clin. Pharmacol. 2005, 61, 755–761. [Google Scholar] [CrossRef]
  169. Ericsson, T.; Masimirembwa, C.; Abelo, A.; Ashton, M. The evaluation of CYP2B6 inhibition by artemisinin antimalarials in recombinant enzymes and human liver microsomes. Drug Metab. Lett. 2013, 6, 247–257. [Google Scholar] [CrossRef]
  170. Burk, O.; Arnold, K.A.; Nussler, A.K.; Schaeffeler, E.; Efimova, E.; Avery, B.A.; Avery, M.A.; Fromm, M.F.; Eichelbaum, M. Antimalarial Artemisinin Drugs Induce Cytochrome P450 and MDR1 Expression by Activation of Xenosensors Pregnane X Receptor and Constitutive Androstane Receptor. Mol. Pharmacol. 2005, 67, 1954–1965. [Google Scholar] [CrossRef]
  171. Willson, T.M.; Kliewer, S.A. Pxr, car and drug metabolism. Nat. Rev. Drug Discov. 2002, 1, 259–266. [Google Scholar] [CrossRef]
  172. Haddad, A.; Davis, M.; Lagman, R. The pharmacological importance of cytochrome CYP3A4 in the palliation of symptoms: Review and recommendations for avoiding adverse drug interactions. Support. Care Cancer 2006, 15, 251–257. [Google Scholar] [CrossRef]
  173. Ademisoye, A.A.; Soyinka, J.O.; Olawoye, S.O.; Igbinoba, S.I.; Olowookere, S.A.; Ademisoye, A.T.; Onyeji, C. Induction of Amodiaquine Metabolism by Rifampicin Following Concurrent Administration in Healthy Volunteers. J. Explor. Res. Pharmacol. 2018, 3, 71–77. [Google Scholar] [CrossRef]
  174. Sahasrabudhe, V.; Zhu, T.; Vaz, A.; Tse, S. Drug Metabolism and Drug Interactions: Potential Application to Antituberculosis Drugs. J. Infect. Dis. 2015, 211, S107–S114. [Google Scholar] [CrossRef] [PubMed]
  175. Ramappa, V.; Aithal, G.P. Hepatotoxicity Related to Anti-tuberculosis Drugs: Mechanisms and Management. J. Clin. Exp. Hepatol. 2013, 3, 37–49. [Google Scholar] [CrossRef]
  176. Cho, H.-J.; Koh, W.-J.; Ryu, Y.-J.; Ki, C.-S.; Nam, M.-H.; Kim, J.-W.; Lee, S.-Y. Genetic polymorphisms of NAT2 and CYP2E1 associated with antituberculosis drug-induced hepatotoxicity in Korean patients with pulmonary tuberculosis. Tuberculosis 2007, 87, 551–556. [Google Scholar] [CrossRef] [PubMed]
  177. Sim, E.; Abuhammad, A.; Ryan, A. Arylamine N-acetyltransferases: From drug metabolism and pharmacogenetics to drug discovery. Br. J. Pharmacol. 2014, 171, 2705–2725. [Google Scholar] [CrossRef]
  178. Sandy, J.; Holton, S.; Fullam, E.; Sim, E.; Noble, M. Binding of the anti-tubercular drug isoniazid to the arylamine N-acetyltransferase protein from Mycobacterium smegmatis. Protein Sci. 2005, 14, 775–782. [Google Scholar] [CrossRef] [PubMed]
  179. Klein, D.J.; Boukouvala, S.; McDonagh, E.M.; Shuldiner, S.R.; Laurieri, N.; Thorn, C.F.; Altman, R.; Klein, T.E. PharmGKB summary. Pharmacogenetics Genom. 2016, 26, 436–444. [Google Scholar] [CrossRef] [PubMed]
  180. Wright, J.M.; Timbrell, J.A. Factors affecting the metabolism of [14C] acetylhydrazine in rats. Drug Metab. Dispos. 1978, 6, 561–566. [Google Scholar] [PubMed]
  181. Timbrell, J.A. Young Scientists Award Lecture 1978: Studies on the Role of Acetylhydrazine in Isoniazid Hepatotoxicity. Arch. Toxicol. 1979, 2, 1–8. [Google Scholar] [CrossRef]
  182. Meitei, H.N.; Pandey, A.; Haobam, R. Polymorphisms in drug metabolism genes as a risk factor for first-line anti-tuberculosis drug-induced liver injury. Mol. Biol. Rep. 2022. [Google Scholar] [CrossRef] [PubMed]
  183. Herrera-Rodulfo, A.; Carrillo-Tripp, M.; Yeverino-Gutierrez, M.L.; Peñuelas-Urquides, K.; González-Escalante, L.A.; de León, M.B.; Silva-Ramirez, B. NAT2 polymorphisms associated with the development of hepatotoxicity after first-line tuberculosis treatment in Mexican patients: From genotype to molecular structure characterization. Clin. Chim. Acta 2021, 519, 153–162. [Google Scholar] [CrossRef]
  184. Wang, P.-Y.; Xie, S.-Y.; Hao, Q.; Zhang, C.; Jiang, B.-F. NAT2 polymorphisms and susceptibility to anti-tuberculosis drug-induced liver injury: A meta-analysis [Review article]. Int. J. Tuberc. Lung Dis. 2012, 16, 589–595. [Google Scholar] [CrossRef] [PubMed]
  185. Daly, A.K.; Day, C.P. Genetic association studies in drug-induced liver injury. Drug Metab. Rev. 2012, 44, 116–126. [Google Scholar] [CrossRef] [PubMed]
  186. Cheng, J.; Krausz, K.W.; Li, F.; Ma, X.; Gonzalez, F.J. CYP2E1-dependent elevation of serum cholesterol, triglycerides, and hepatic bile acids by isoniazid. Toxicol. Appl. Pharmacol. 2012, 266, 245–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Sotsuka, T.; Sasaki, Y.; Hirai, S.; Yamagishi, F.; Ueno, K. Association of isoniazid-metabolizing enzyme genotypes and isoniazid-induced hepatotoxicity in tuberculosis patients. Vivo 2011, 25, 803–812. [Google Scholar]
  188. Swaminathan, S.; Ramachandran, G. Role of pharmacogenomics in the treatment of tuberculosis: A review. Pharm. Pers. Med. 2012, 5, 89–98. [Google Scholar] [CrossRef]
  189. Wang, P.; Pradhan, K.; Zhong, X.-B.; Ma, X. Isoniazid metabolism and hepatotoxicity. Acta Pharm. Sin. B 2016, 6, 384–392. [Google Scholar] [CrossRef]
  190. Sarich, T.C.; Adams, S.P.; Petricca, G.; Wright, J.M. Inhibition of isoniazid-induced hepatotoxicity in rabbits by pretreatment with an amidase inhibitor. J. Pharmacol. Exp. Ther. 1999, 289, 695–702. [Google Scholar] [PubMed]
  191. Nicoletti, P.; Devarbhavi, H.; Goel, A.; Venkatesan, R.; Eapen, C.E.; Grove, J.I.; Zafer, S.; Bjornsson, E.; Lucena, M.I.; Andrade, R.J.; et al. Genetic Risk Factors in Drug-Induced Liver Injury Due to Isoniazid-Containing Antituberculosis Drug Regimens. Clin. Pharmacol. Ther. 2020, 109, 1125–1135. [Google Scholar] [CrossRef]
  192. Steele, M.A.; Burk, R.F.; DesPrez, R.M. Toxic Hepatitis with Isoniazid and Rifampin: A Meta-analysis. Chest 1991, 99, 465–471. [Google Scholar] [CrossRef]
  193. Sarma, G.R.; Immanuel, C.; Kailasam, S.; Narayana, A.S.; Venkatesan, P. Rifampin-induced release of hydrazine from isoniazid. A possible cause of hepatitis during treatment of tuberculosis with regimens containing isoniazid and rifampin. Am. Rev. Respir. Dis. 1986, 133. [Google Scholar] [CrossRef]
  194. Abel, L.; Fellay, J.; Haas, D.W.; Schurr, E.; Srikrishna, G.; Urbanowski, M.; Chaturvedi, N.; Srinivasan, S.; Johnson, D.H.; Bishai, W.R. Genetics of human susceptibility to active and latent tuberculosis: Present knowledge and future perspectives. Lancet Infect. Dis. 2018, 18, e64–e75. [Google Scholar] [CrossRef] [PubMed]
  195. Ward, B.A.; Gorski, J.C.; Jones, D.R.; Hall, S.D.; Flockhart, D.A.; Desta, Z. The Cytochrome P450 2B6 (CYP2B6) Is the Main Catalyst of Efavirenz Primary and Secondary Metabolism: Implication for HIV/AIDS Therapy and Utility of Efavirenz as a Substrate Marker of CYP2B6 Catalytic Activity. Experiment 2003, 306, 287–300. [Google Scholar] [CrossRef]
  196. Luetkemeyer, A.F.; Rosenkranz, S.L.; Lu, D.; Grinsztejn, B.; Sanchez, J.; Ssemmanda, M.; Sanne, I.; McIlleron, H.; Havlir, D.V.; Haas, D.W. Combined Effect of CYP2B6 and NAT2 Genotype on Plasma Efavirenz Exposure during Rifampin-based Antituberculosis Therapy in the STRIDE Study. Clin. Infect. Dis. 2015, 60, 1860–1863. [Google Scholar] [CrossRef]
  197. Kwara, A.; Lartey, M.; Sagoe, K.W.; Court, M.H. Paradoxically elevated efavirenz concentrations in HIV/tuberculosis-coinfected patients with CYP2B6 516TT genotype on rifampin-containing antituberculous therapy. Aids 2011, 25, 388–390. [Google Scholar] [CrossRef] [PubMed]
  198. Dooley, K.E.; Denti, P.; Martinson, N.; Cohn, S.; Mashabela, F.; Hoffmann, J.; Haas, D.W.; Hull, J.; Msandiwa, R.; Castel, S.; et al. Pharmacokinetics of Efavirenz and Treatment of HIV-1 among Pregnant Women with and without Tuberculosis Coinfection. J. Infect. Dis. 2014, 211, 197–205. [Google Scholar] [CrossRef]
  199. DI Iulio, J.; Fayet, A.; Arab-Alameddine, M.; Rotger, M.; Lubomirov, R.; Cavassini, M.; Furrer, H.; Günthard, H.; Colombo, S.; Csajka, C.; et al. In vivo analysis of efavirenz metabolism in individuals with impaired CYP2A6 function. Pharm. Genom. 2009, 19, 300–309. [Google Scholar] [CrossRef] [PubMed]
  200. Bertilsson, L.; Dahl, M.-L.; Dalén, P.; Al-Shurbaji, A. Molecular genetics of CYP2D6: Clinical relevance with focus on psychotropic drugs. Br. J. Clin. Pharmacol. 2002, 53, 111–122. [Google Scholar] [CrossRef] [PubMed]
  201. Jarvis, J.P.; Peter, A.P.; Shaman, J.A. Consequences of CYP2D6 Copy-Number Variation for Pharmacogenomics in Psychiatry. Front. Psychiatry 2019, 10, 432. [Google Scholar] [CrossRef]
  202. Zanger, U.M. Genetic variability of CYP2D6: Basic and clinical aspects. Future Med. 2014, 28–41. [Google Scholar] [CrossRef]
  203. Elewa, H.; Wilby, K.J. A Review of Pharmacogenetics of Antimalarials and Associated Clinical Implications. Eur. J. Drug Metab. Pharmacokinet. 2017, 42, 745–756. [Google Scholar] [CrossRef] [PubMed]
  204. Bains, R.K. African variation at Cytochrome P450 genes. Evol. Med. Public Health 2013, 2013, 118–134. [Google Scholar] [CrossRef]
  205. Adehin, A.; Igbinoba, S.I.; Soyinka, J.O.; Onyeji, C.O.; Babalola, C.P.; Bolaji, O.O. Pharmacokinetic Parameters of Quinine in Healthy Subjects and in Patients with Uncomplicated Malaria in Nigeria: Analysis of Data using a Population Approach. Curr. Ther. Res. 2019, 91, 33–38. [Google Scholar] [CrossRef] [PubMed]
  206. Silvino, A.C.R.; Kano, F.S.; Costa, M.A.; Fontes, C.J.F.; Soares, I.S.; de Brito, C.F.A.; Carvalho, L.H.; Sousa, T.N. Novel Insights into Plasmodium vivax Therapeutic Failure: CYP2D6 Activity and Time of Exposure to Malaria Modulate the Risk of Recurrence. Antimicrob. Agents Chemother. 2020, 64. [Google Scholar] [CrossRef]
  207. Mukonzo, J.K.; Waako, P.; Ogwal-Okeng, J.; Gustafsson, L.L.; Aklillu, E. Genetic Variations in ABCB1 and CYP3A5 as well as Sex Influence Quinine Disposition Among Ugandans. Ther. Drug Monit. 2010, 32, 346–352. [Google Scholar] [CrossRef]
  208. Zhou, S.-F.; Liu, J.-P.; Chowbay, B. Polymorphism of human cytochrome P450 enzymes and its clinical impact. Drug Metab. Rev. 2009, 41, 89–295. [Google Scholar] [CrossRef]
  209. Mirghani, R.A.; Sayi, J.; Aklillu, E.; Allqvist, A.; Jande, M.; Wennerholm, A.; Eriksen, J.; Herben, V.M.; Jones, B.C.; Gustafsson, L.L.; et al. CYP3A5 genotype has significant effect on quinine 3-hydroxylation in Tanzanians, who have lower total CYP3A activity than a Swedish population. Pharm. Genom. 2006, 16, 637–645. [Google Scholar] [CrossRef]
  210. Soyinka, J.O.; Nnadi, C.O.; Onyeji, C.O. Insights and Current Perspectives on Pharmacogenomics of Antimalarial Drugs. Res. Sq. 2010. [Google Scholar] [CrossRef]
  211. McLeay, S.C.; Vis, P.; van Heeswijk, R.P.G.; Green, B. Population Pharmacokinetics of Bedaquiline (TMC207), a Novel Antituberculosis Drug. Antimicrob. Agents Chemother. 2014, 58, 5315–5324. [Google Scholar] [CrossRef] [PubMed]
  212. Haas, D.W.; Abdelwahab, M.T.; van Beek, S.W.; Baker, P.; Maartens, G.; Bradford, Y.; Ritchie, M.D.; Wasserman, S.; Meintjes, G.; Beeri, K.; et al. Pharmacogenetics of Between-Individual Variability in Plasma Clearance of Bedaquiline and Clofazimine in South Africa. J. Infect. Dis. 2022, 226, 147–156. [Google Scholar] [CrossRef]
  213. Hustert, E.; Haberl, M.; Burk, O.; Wolbold, R.; He, Y.-Q.; Klein, K.; Nuessler, A.C.; Neuhaus, P.; Klattig, J.; Eiselt, R.; et al. The genetic determinants of the CYP3A5 polymorphism. Pharmacogenetics 2001, 11, 773–779. [Google Scholar] [CrossRef]
  214. Lamba, J.; Hebert, J.M.; Schuetz, E.G.; Klein, T.E.; Altman, R.B. PharmGKB summary. Pharmacogenetics Genom. 2012, 22, 555–558. [Google Scholar] [CrossRef] [PubMed]
  215. Sheng, Y.-J.; Wu, G.; He, H.-Y.; Chen, W.; Zou, Y.-S.; Li, Q.; Zhong, L.; Huang, Y.-M.; Deng, C.-L. The association between CYP2E1 polymorphisms and hepatotoxicity due to anti-tuberculosis drugs: A meta-analysis. Infect. Genet. Evol. 2014, 24, 34–40. [Google Scholar] [CrossRef]
  216. Cai, Y.; Yi, J.; Zhou, C.; Shen, X. Pharmacogenetic Study of Drug-Metabolising Enzyme Polymorphisms on the Risk of Anti-Tuberculosis Drug-Induced Liver Injury: A Meta-Analysis. PLoS ONE 2012, 7, e47769. [Google Scholar] [CrossRef] [PubMed]
  217. Deng, R.; Yang, T.; Wang, Y.; Tang, N. CYP2E1 RsaI/PstI polymorphism and risk of anti-tuberculosis drug-induced liver injury: A meta-analysis [Review article]. Int. J. Tuberc. Lung Dis. 2012, 16, 1574–1581. [Google Scholar] [CrossRef] [PubMed]
  218. Wang, F.-J.; Wang, Y.; Niu, T.; Lu, W.-X.; Sandford, A.J.; He, J.-Q. Update meta-analysis of the CYP2E1 RsaI/PstI and DraI polymorphisms and risk of antituberculosis drug-induced hepatotoxicity: Evidence from 26 studies. J. Clin. Pharm. Ther. 2016, 41, 334–340. [Google Scholar] [CrossRef] [PubMed]
  219. Yang, S.; Hwang, S.J.; Park, J.Y.; Chung, E.K.; I Lee, J. Association of genetic polymorphisms of CYP2E1, NAT2, GST and SLCO1B1 with the risk of anti-tuberculosis drug-induced liver injury: A systematic review and meta-analysis. BMJ Open 2019, 9, e027940. [Google Scholar] [CrossRef]
  220. Huang, Y.; Chern, H.; Su, W.; Wu, J.; Chang, S.; Chiang, C.; Chang, F.; Lee, S. Cytochrome P450 2E1 genotype and the susceptibility to antituberculosis drug-induced hepatitis. Hepatology 2003, 37, 924–930. [Google Scholar] [CrossRef] [PubMed]
  221. Tang, S.-W.; Lv, X.-Z.; Zhang, Y.; Wu, S.-S.; Yang, Z.-R.; Xia, Y.-Y.; Tu, D.-H.; Deng, P.-Y.; Ma, Y.; Chen, D.-F.; et al. CYP2E1, GSTM1 and GSTT1genetic polymorphisms and susceptibility to antituberculosis drug-induced hepatotoxicity: A nested case-control study. J. Clin. Pharm. Ther. 2012, 37, 588–593. [Google Scholar] [CrossRef]
  222. Chamorro, J.G.; Castagnino, J.P.; Musella, R.M.; Nogueras, M.; Aranda, F.M.; Frías, A.; Visca, M.; Aidar, O.; Perés, S.; de Larrañaga, G.F. Sex, ethnicity, and slow acetylator profile are the major causes of hepatotoxicity induced by antituberculosis drugs. J. Gastroenterol. Hepatol. 2013, 28, 323–328. [Google Scholar] [CrossRef]
  223. Xiang, Y.; Ma, L.; Wu, W.; Liu, W.; Li, Y.; Zhu, X.; Ma, J.; Cao, M.; Wang, Q.; Yao, X.; et al. The Incidence of Liver Injury in Uyghur Patients Treated for TB in Xinjiang Uyghur Autonomous Region, China, and Its Association with Hepatic Enzyme Polymorphisms NAT2, CYP2E1, GSTM1 and GSTT1. PLoS ONE 2014, 9, e85905. [Google Scholar] [CrossRef]
  224. Yang, R.; Liu, H.; Chen, Z.; Qi, T.; Zhang, Z.; Qu, Q. Clinical pharmacists use CYP2C19 genotyping test to guide individual medication therapy of Clopidogrel and to evaluate the efficacy of treatment. Chin. J. Clin. Pharmacol. Ther. 2019, 24, 938. [Google Scholar] [CrossRef]
  225. Zhang, J.; Zhu, X.; Li, Y.; Zhu, L.; Li, S.; Zheng, G.; Ren, Q.; Xiao, Y.; Feng, F. Correlation of CpG Island Methylation of the Cytochrome P450 2E1/2D6 Genes with Liver Injury Induced by Anti-Tuberculosis Drugs: A Nested Case-Control Study. Int. J. Environ. Res. Public Health 2016, 13, 776. [Google Scholar] [CrossRef]
  226. Wei, Y.; Huai, C.; Zhou, C.; Gao, Y.; Chen, L.; Zhou, W.; Wei, M.; Qin, S. A methylation functional detection hepatic cell system validates correlation between DNA methylation and drug-induced liver injury. Pharm. J. 2020, 20, 717–723. [Google Scholar] [CrossRef] [PubMed]
  227. Ramachandran, S.; Deshpande, O.; Roseman, C.C.; Rosenberg, N.A.; Feldman, M.W.; Cavalli-Sforza, L.L. Support from the relationship of genetic and geographic distance in human populations for a serial founder effect originating in Africa. Proc. Natl. Acad. Sci. USA 2005, 102, 15942–15947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Alessandrini, M.; Asfaha, S.; Dodgen, T.M.; Warnich, L.; Pepper, M.S. Cytochrome P450 pharmacogenetics in African populations. Drug Metab. Rev. 2013, 45, 253–275. [Google Scholar] [CrossRef] [PubMed]
  229. Dandara, C.; Swart, M.; Mpeta, B.; Wonkam, A.; Masimirembwa, C. Cytochrome P450 pharmacogenetics in African populations: Implications for public health. Expert Opin. Drug Metab. Toxicol. 2014, 10, 769–785. [Google Scholar] [CrossRef] [PubMed]
  230. Zhou, Y.; Lauschke, V.M. The genetic landscape of major drug metabolizing cytochrome P450 genes—An updated analysis of population-scale sequencing data. Pharm. J. 2022, 22, 284–293. [Google Scholar] [CrossRef] [PubMed]
  231. Fukami, T.; Nakajima, M.; Yoshida, R.; Tsuchiya, Y.; Fujiki, Y.; Katoh, M.; McLeod, H.L.; Yokoi, T. A novel polymorphism of human gene has an amino acid substitution (V365M) that decreases enzymatic activity in vitro and in vivo. Clin. Pharmacol. Ther. 2004, 76, 519–527. [Google Scholar] [CrossRef]
  232. Lewis, D.; Dickins, M.; Lake, B.; Eddershaw, P.; Tarbit, M.; Goldfarb, P. Molecular modelling of the human cytochrome P450 isoform CYP2A6 and investigations of CYP2A substrate selectivity. Toxicology 1999, 133, 1–33. [Google Scholar] [CrossRef] [PubMed]
  233. Ho, M.K.; Mwenifumbo, J.C.; Zhao, B.; Gillam, E.; Tyndale, R.F. A novel CYP2A6 allele, CYP2A6*23, impairs enzyme function in vitro and in vivo and decreases smoking in a population of Black-African descent. Pharm. Genom. 2008, 18, 67–75. [Google Scholar] [CrossRef]
  234. Kato, K.; Nakayoshi, T.; Nokura, R.; Hosono, H.; Hiratsuka, M.; Ishikawa, Y.; Kurimoto, E.; Oda, A. Deciphering Structural Alterations Associated with Activity Reductions of Genetic Polymorphisms in Cytochrome P450 2A6 Using Molecular Dynamics Simulations. Int. J. Mol. Sci. 2021, 22, 10119. [Google Scholar] [CrossRef]
  235. Zanger, U.M.; Klein, K. Pharmacogenetics of cytochrome P450 2B6 (CYP2B6): Advances on polymorphisms, mechanisms, and clinical relevance. Front. Genet. 2013, 4, 24. [Google Scholar] [CrossRef]
  236. Honda, M.; Muroi, Y.; Tamaki, Y.; Saigusa, D.; Suzuki, N.; Tomioka, Y.; Matsubara, Y.; Oda, A.; Hirasawa, N.; Hiratsuka, M. Functional Characterization of CYP2B6 Allelic Variants in Demethylation of Antimalarial Artemether. Drug Metab. Dispos. 2011, 39, 1860–1865. [Google Scholar] [CrossRef]
  237. Klein, K.; Lang, T.; Saussele, T.; Barbosa-Sicard, E.; Schunck, W.-H.; Eichelbaum, M.; Schwab, M.; Zanger, U.M. Genetic variability of CYP2B6 in populations of African and Asian origin: Allele frequencies, novel functional variants, and possible implications for anti-HIV therapy with efavirenz. Pharm. Genom. 2005, 15, 861–873. [Google Scholar] [CrossRef]
  238. Arnold, W.R.; Zelasko, S.; Meling, D.D.; Sam, K.; Das, A. Polymorphisms of CYP2C8 Alter First-Electron Transfer Kinetics and Increase Catalytic Uncoupling. Int. J. Mol. Sci. 2019, 20, 4626. [Google Scholar] [CrossRef]
  239. Martiny, V.Y.; Miteva, M.A. Advances in Molecular Modeling of Human Cytochrome P450 Polymorphism. J. Mol. Biol. 2013, 425, 3978–3992. [Google Scholar] [CrossRef]
  240. Parikh, S.J.; Kamat, S.; Phillips, M.; Boyson, S.P.; Yarbrough, T.; Davie, D.; Zhang, Q.; Glass, K.C.; Shah, M.B. Insights into the Genetic Variations of Human Cytochrome P450 2C9: Structural Analysis, Characterization and Comparison. Int. J. Mol. Sci. 2021, 22, 10206. [Google Scholar] [CrossRef]
  241. Gyulkhandanyan, A.; Rezaie, A.R.; Roumenina, L.; Lagarde, N.; Fremeaux-Bacchi, V.; Miteva, M.; Villoutreix, B.O. Analysis of protein missense alterations by combining sequence- and structure-based methods. Mol. Genet. Genom. Med. 2020, 8, e1166. [Google Scholar] [CrossRef]
  242. Blaisdell, J.; Jorge-Nebert, L.F.; Coulter, S.; Ferguson, S.S.; Lee, S.-J.; Chanas, B.; Xi, T.; Mohrenweiser, H.; Ghanayem, B.; A Goldstein, J. Discovery of new potentially defective alleles of human CYP2C9. Pharmacogenetics 2004, 14, 527–537. [Google Scholar] [CrossRef]
  243. Blaisdell, J.; Mohrenweiser, H.; Jackson, J.; Ferguson, S.; Coulter, S.; Chanas, B.; Xi, T.; Ghanayem, B.; A Goldstein, J. Identification and functional characterization of new potentially defective alleles of human CYP2C19. Pharmacogenetics 2002, 12, 703–711. [Google Scholar] [CrossRef] [PubMed]
  244. Wei, D.-Q.; Wang, J.-F.; Chen, C.; Li, Y.; Chou, K.-C. Molecular Modeling of Two CYP2C19 SNPs and Its Implications for Personalized Drug Design. Protein Pept. Lett. 2008, 15, 27–32. [Google Scholar] [CrossRef]
  245. Arendse, L.B. Exploring the Effects of Polymorphic Variation on the Stability and Function of Human Cytochrome P450 Enzymes In Silico And In Vitro. Ph.D. Thesis, University of Cape Town, Cape Town, South Africa, 2014. [Google Scholar]
  246. Dong, A.N.; Ahemad, N.; Pan, Y.; Palanisamy, U.D.; Yiap, B.C.; Ong, C.E. Functional and structural characterisation of common cytochrome P450 2D6 allelic variants—Roles of Pro34 and Thr107 in catalysis and inhibition. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2019, 392, 1015–1029. [Google Scholar] [CrossRef] [PubMed]
  247. Handa, K.; Nakagome, I.; Yamaotsu, N.; Gouda, H.; Hirono, S. In Silieo Study on the Inhibitory Interaction of Drugs with Wild-type CYP2D6.1 and the Natural Variant CYP2D6.17. Drug Metab. Pharmacokinet. 2014, 29, 52–60. [Google Scholar] [CrossRef]
  248. Gotoh, O. Substrate recognition sites in cytochrome P450 family 2 (CYP2) proteins inferred from comparative analyses of amino acid and coding nucleotide sequences. J. Biol. Chem. 1992, 267, 83–90. [Google Scholar] [CrossRef]
  249. Wennerholm, A.; Johansson, I.; Hidestrand, M.; Bertilsson, L.; Gustafsson, L.L.; Ingelman-Sundberg, M. Characterization of the CYP2D6*29 allele commonly present in a black Tanzanian population causing reduced catalytic activity. Pharmacogenetics 2001, 11, 417–427. [Google Scholar] [CrossRef]
  250. Fang, P.; Tang, P.-F.; Xu, R.-A.; Zheng, X.; Wen, J.; Bao, S.-S.; Cai, J.-P.; Hu, G.-X. Functional assessment of CYP3A4 allelic variants on lidocaine metabolism in vitro. Drug Des. Dev. Ther. 2017, 11, 3503–3510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (A) General structure of CYP enzymes. CYP2C8 (PDB ID: 2NNJ) is used to map the 12 major helices denoted A-L in different colors. Substrate entrance site is highlighted in red and an arrow pointing towards the entrance is in black. (B) Conserved catalytically important residues mapped onto the CYP structure, and the cysteine that acts as a proximal axial thiolate ligand for the heme iron is highlighted. Threonine in the conserved motif (A/G)XXT is colored in yellow, while glutamic acid and arginine in the EXXR motif are colored in warm pink. Phenylalanine, glycine and cysteine in the conserved motif FXXGXXXCXG are all colored in purple. (C) CYP2C8 enzyme residue numbers are denoted, as an example, for the helices observed in CYP enzymes.
Figure 1. (A) General structure of CYP enzymes. CYP2C8 (PDB ID: 2NNJ) is used to map the 12 major helices denoted A-L in different colors. Substrate entrance site is highlighted in red and an arrow pointing towards the entrance is in black. (B) Conserved catalytically important residues mapped onto the CYP structure, and the cysteine that acts as a proximal axial thiolate ligand for the heme iron is highlighted. Threonine in the conserved motif (A/G)XXT is colored in yellow, while glutamic acid and arginine in the EXXR motif are colored in warm pink. Phenylalanine, glycine and cysteine in the conserved motif FXXGXXXCXG are all colored in purple. (C) CYP2C8 enzyme residue numbers are denoted, as an example, for the helices observed in CYP enzymes.
Ijms 24 03383 g001
Figure 3. (A) CYP2D6 complexed with quinidine (PDB ID: 4WNU). (B) CYP2D6 complexed with quinine (PDB ID: 4WNV). Enzyme carbons are colored orange and ligand carbons are colored yellow. Oxygen, nitrogen, sulfur, iron, and nickel atoms are colored red, blue, yellow, orange, and green, respectively. Water molecule shown as a red sphere. Distances between the nitrogen in the quinuclidine ring of the bound quinidine with Asp-301 and Glu-21 are shown by dashed lines and distances are given in Å [84].
Figure 3. (A) CYP2D6 complexed with quinidine (PDB ID: 4WNU). (B) CYP2D6 complexed with quinine (PDB ID: 4WNV). Enzyme carbons are colored orange and ligand carbons are colored yellow. Oxygen, nitrogen, sulfur, iron, and nickel atoms are colored red, blue, yellow, orange, and green, respectively. Water molecule shown as a red sphere. Distances between the nitrogen in the quinuclidine ring of the bound quinidine with Asp-301 and Glu-21 are shown by dashed lines and distances are given in Å [84].
Ijms 24 03383 g003
Figure 4. Human pathways of isoniazid metabolism involving NAT2, CYP2E1, and GST enzymes.
Figure 4. Human pathways of isoniazid metabolism involving NAT2, CYP2E1, and GST enzymes.
Ijms 24 03383 g004
Figure 5. Afrocentric missense mutations and their location within the general CYP enzyme structure. (A) Mutations more frequent in African populations and highly restricted to African populations mapped to CYP2D6 (PDB ID 3TDA). Mutations in all Afrocentric alleles were mapped to their corresponding positions in CYP2D6 and colored orange. The (A/G)XXT region is shaded green, while the EXXR region is shown in the color purple, and the FXXGXXXCXG conserved domain is in blue. (B) Mutations in close proximity to known important residues in the active site and in the substrate channel. Active site residues of CYP2D6 shown in red, and residues in the substrate channel shown in cream. Mutation I19L could not be mapped to CYP2C19*15 as this region is not available on all crystalized CYP enzymes to date.
Figure 5. Afrocentric missense mutations and their location within the general CYP enzyme structure. (A) Mutations more frequent in African populations and highly restricted to African populations mapped to CYP2D6 (PDB ID 3TDA). Mutations in all Afrocentric alleles were mapped to their corresponding positions in CYP2D6 and colored orange. The (A/G)XXT region is shaded green, while the EXXR region is shown in the color purple, and the FXXGXXXCXG conserved domain is in blue. (B) Mutations in close proximity to known important residues in the active site and in the substrate channel. Active site residues of CYP2D6 shown in red, and residues in the substrate channel shown in cream. Mutation I19L could not be mapped to CYP2C19*15 as this region is not available on all crystalized CYP enzymes to date.
Ijms 24 03383 g005
Table 1. Some CYP enzyme metabolized antimalarial drugs and their oxidized products.
Table 1. Some CYP enzyme metabolized antimalarial drugs and their oxidized products.
Antimalarial DrugCYP Enzyme MetabolizerMechanism of ActionReferences
AmodiaquineCYP2C8Metabolised to desethylamodiaquine (DEAQ)—likely to proceed in two steps, a hydrogen abstraction and hydroxylation at the adjacent carbon, forming an unstable carbinolamide that rapidly hydrolyzes to DEAQ and acetaldehyde.[116]
ArteetherCYP2B6, CYP3A4, CYP3A5Arteether is deethylated to dihydroartemisinin (DHA), the main bioactive metabolite of atermisinin and its derivatives.[120]
Artelinic acidCYP3A4, CYP3A5Artelinic acid is O-debenzylated to DHA[120,121]
ArtemetherCYP3A4, CYP1A2, CYP2B6Artemether is demethylated to the bioactive metabolite DHA[122,123,124]
ArtemisininCYP2A6, CYP2B6, CYP3A4Artemisinin is not itself metabolized to DHA but acts as the primary antimalarial. Upon reaction with Fe2+ it is converted first into oxygen centered free radicals derived by reductive cleavage of its peroxide bridge, which are then converted into carbon centered free radicals by intramolecular hydrogen abstraction from CH2 groups on the periphery of the artemisinin by the O centered radicals.[116,125,126]
ArtesunateCYP2A6, CYP2B6Rapidly hydrolyzed to the bioactive metabolite DHA[116,125,126,127]
ChloroquineCYP2C8, CYP2C19, CYP3A4, CYP2D6, CYP3A4, CYP3A5Chloroquine is dealkylated into N-desethylchloroquine (DCQ) and N-bis-desethylchloroquine (BDCQ), with DCQ being the major metabolite. [113,128,129]
DapsoneCYP2C9, CYP3A4Unlike other antimalarials, the first metabolizing enzyme of dapsone is N-acetyltransferase which hydrolyses the drug to the active form monoacetyl dapsone. CYP enzymes on the other hand hydrolyze the drug to its N-hydroxy metabolites dapsone hydroxylamine and monoacetyl hydroxylamine which are harmful hemolytic metabolites.[116,117,119,130,131]
Halofantrine and LumefantrineCYP3A4, CYP3A5Halofantrine undergoes desbutylation to N-desbutyl-halofantrine which possesses some antimalarial activity while lumefantrine is metabolized to desbutyl-lumefantrine and excreted via bile and faces.[115,132]
MefloquineCYP1A2, CYP3A4Metabolized into carboxymefloquine metabolite which has little or no antimalarial activity.[121,133,134,135]
PrimaquineCYP1A2, CYP3A4, CYP2D6Three possible pathways exist for primaquine metabolism and that involving CYP enzymes is hydroxylation at different positions on the quinoline ring, with mono-, di-, or even tri-hydroxylations possible, and subsequent glucuronide conjugation of the hydroxylated metabolites. The main metabolite carboxyprimaquine comes about through a monoamine oxidase catalyzed oxidative deamination.[116,136]
ProguanilCYP2C19Oxidative metabolized to cycloguanil, which is the active form of the drug[116,121,137]
Quinine CYP3A4Undergoes hydroxylation to the main metabolite 3-hydroxiquinine[28,121,138,139]
Table 2. CYP alleles with missense mutations that have the highest frequency in African populations and/or are highly restricted to African populations [18,230]. Known structural effects and functional consequences of the mutations are included.
Table 2. CYP alleles with missense mutations that have the highest frequency in African populations and/or are highly restricted to African populations [18,230]. Known structural effects and functional consequences of the mutations are included.
CYP EnzymeAntimalarial/Antituberculosis Drug MetabolizedAlleleFrequency in African Populations (%)Amino acid Mutation PositionFunctional ConsequenceResidue/Mutation Notes
CYP2A6Artemisinin, artesunateCYP2A6*1710.9V365M-Located in CYP2A6 substrate recognition site SRS-5 forming part of the active site region and some side chains that point into the heme pocket, suggesting this residue may be important for substrate specificity. In vitro enzyme assays and metabolism studies showed no effect of the mutation on the stability of the enzyme. Both valine and methionine are non-polar hydrophobic residues; however, methionine is larger and contains a sulfur atom [231].
CYP2A6*231.4R203CDecreasedResidue located on α-helix F in substrate recognition site 2/3. Change from large basic residue to medium sized polar residue. Molecular modeling suggests Arg203 could orient important binding residue Phe209 [232,233].
CYP2A6*251.4F118L-During molecular dynamics simulation, the F118L mutant side chain moves away from the heme and affects secondary structure formation and interaction with heme and substrates [234].
CYP2A6*281.5N418D
E419D
-N418D and E419D cause a structural change in the substrate access channel and the substrate binding site [234].
CYP2B6Artemisinin, artesunate, arteetherCYP2B6*632Q172H; K262RDecreasedQ172H and K262R are not located at the active site and have not been identified in substrate recognition sites [235].
CYP2B6*187I328TInactiveResults in no detectable protein or activity in vitro [236]. Designated as a null allele [237]. The only commonly found inactive CYP2B6 missense mutation, and highly restricted to Africa [230].
CYP2C8Chloroquine, amodiaquine, bedaquilineCYP2C8*215.2I269FDecreasedResidue located on enzyme surface. Larger residue change. Possible effect on enzyme folding and interaction with cytochrome P450 reductase [238].
CYP 2C9DapsoneCYP 2C9*51.1D360EDecreasedIn molecular dynamics simulations, D360E broke the hydrogen bond between D360 and S478 (helix K and loop β4) leading to local structure destabilization. Glutamic acid is larger than aspartic acid [239].
CYP 2C9*86R150HDecreasedR150 is highly conserved. Located on protein surface at D-helix region, away from the active site. The structure of the CYP2C9*8-losartan complex has ~60° rotation of the H150 sidechain in an alternate conformation compared to the sidechain of R150 [240]. Possible involvement of R150H in the salt bridge network with the neighboring residues, suggesting the change to histidine at this solvent-exposed region may no longer coordinate similar ionic and electrostatic interactions and may result in destabilization of the structure [241]. Possible that this change could influence reductase binding [242].
CYP 2C9*97.5H251RNormalH251R is located near the C-terminal end of helix G and hydrogen bonds to D262 on helix H. Arginine has a longer side chain than histidine [242].
CYP 2C19Proguanil, quinine, dapsone, bedaquilineCYP 2C19*91.3R144H
I331V
DecreasedThe R144H mutation could affect enzyme structure and function. The conserved arginine is located in the D helix and is part of a complex salt bridge with conserved Ser180 in the E helix and the backbone of the turn before helix H. The arginine could help stabilize the structure of the enzyme and may also be a part of the hinge for the F–G loop [243]. Docking experiments suggest that I331 is involved in ligand binding. The I331V mutant has differing lipophilicity in the binding pocket or active cavity [239,244].
CYP 2C19*131.8I331V
R410C
NormalResidue 410 located on enzyme surface. Change from large basic residue to medium sized polar residue. Mutation has neutral effect on enzyme activity.
CYP 2C19*151.9I19L
I331V
NormalThe I19L mutation with residue 19 is part of signal-anchor sequence, located in truncated N-terminal region. Mutation has neutral effect [245].
CYP 2D6ChloroquineCYP 2D6*222.5R296C
S486T
NormalR296C is located in substrate recognition site 4 (I-helix) important for catalytic proton delivery. Kinetics data indicate that R296C in CYP2D6*2 causes enhanced ligand binding possibly due to morphological changes Change from medium-sized basic residue to small polar residue. [246]. S486T mutation is located in substrate recognition site (β4-2 sheet) and is very close to active site and heme [239]. Kinetics data indicate that S486T in CYP2D6*2 causes enhanced ligand binding possibly due to morphological changes [246].
CYP 2D6*1720.5T107I
R296C
S486T
DecreasedT107I is located in substrate recognition site 1 (B′-helix). Substitution of hydrophilic Threonine to hydrophobic Isoleucine. Threonine has alcoholic side chain pointing down towards heme, whereas Isoleucine long aliphatic side chain points away from active site [246]. Causes changes in hydrogen bonds with surrounding residues within the active site cavity [247].
CYP 2D6*298.9V136I
R296C
V338M
S486T
DecreasedV136 is located between substrate recognition sites 1 and 2 in helix C [248]. In contact with cytochrome P450 reductase. Neutral effect on protein stability [245]. V338M is located between substrate recognition sites 4 and 5 at the end of helix J [248]. Neutral effect on protein stability [245]. Expression studies reveal that both the V136I and the V338M mutations affect catalytic activity, and the effect is stronger when present together as seen in CYP2D6*29. Both valine and methionine are non-polar hydrophobic residues; however, methionine is larger and contains a sulfur atom [249].
CYP3A4Quinine, quinidine, chloroquine, mefloquine, primaquine, halofantrine, lumefantrine, dapsone, artemisinin, artemether, arteether, artelinic, acid, delaminid, bedaquilineCYP3A4*152.6R162Q-Mutation related to rapid metabolism of quinine in vitro. Change from large basic residue to medium-sized polar hydrophilic residue [250].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chamboko, C.R.; Veldman, W.; Tata, R.B.; Schoeberl, B.; Tastan Bishop, Ö. Human Cytochrome P450 1, 2, 3 Families as Pharmacogenes with Emphases on Their Antimalarial and Antituberculosis Drugs and Prevalent African Alleles. Int. J. Mol. Sci. 2023, 24, 3383. https://doi.org/10.3390/ijms24043383

AMA Style

Chamboko CR, Veldman W, Tata RB, Schoeberl B, Tastan Bishop Ö. Human Cytochrome P450 1, 2, 3 Families as Pharmacogenes with Emphases on Their Antimalarial and Antituberculosis Drugs and Prevalent African Alleles. International Journal of Molecular Sciences. 2023; 24(4):3383. https://doi.org/10.3390/ijms24043383

Chicago/Turabian Style

Chamboko, Chiratidzo R., Wayde Veldman, Rolland Bantar Tata, Birgit Schoeberl, and Özlem Tastan Bishop. 2023. "Human Cytochrome P450 1, 2, 3 Families as Pharmacogenes with Emphases on Their Antimalarial and Antituberculosis Drugs and Prevalent African Alleles" International Journal of Molecular Sciences 24, no. 4: 3383. https://doi.org/10.3390/ijms24043383

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

Article Metrics

Back to TopTop