Biology:CYP3A4

From HandWiki
Short description: Enzyme that metabolizes substances by oxidation
CYP3A4
Identifiers
EC number1.14.14.56
Databases
IntEnzIntEnz view
BRENDABRENDA entry
ExPASyNiceZyme view
KEGGKEGG entry
MetaCycmetabolic pathway
PRIAMprofile
PDB structuresRCSB PDB PDBe PDBsum
A representation of the 3D structure of the protein myoglobin showing turquoise α-helices.
Generic protein structure example

Cytochrome P450 3A4 (abbreviated CYP3A4) (EC 1.14.13.97) is an important enzyme in the body, mainly found in the liver and in the intestine, which in humans is encoded by CYP3A4 gene. It oxidizes small foreign organic molecules (xenobiotics), such as toxins or drugs, so that they can be removed from the body. It is highly homologous to CYP3A5, another important CYP3A enzyme.

While many drugs are deactivated by CYP3A4, there are also some drugs that are activated by the enzyme. Some substances, such as some drugs and furanocoumarins present in grapefruit juice, interfere with the action of CYP3A4. These substances will, therefore, either amplify or weaken the action of those drugs that are modified by CYP3A4.

CYP3A4 is a member of the cytochrome P450 family of oxidizing enzymes. Several other members of this family are also involved in drug metabolism, but CYP3A4 is the most common and the most versatile one. Like all members of this family, it is a hemoprotein, i.e. a protein containing a heme group with an iron atom. In humans, the CYP3A4 protein is encoded by the CYP3A4 gene.[1] This gene is part of a cluster of cytochrome P450 genes on chromosome 7q22.1.[2] Previously another CYP3A gene, CYP3A3, was thought to exist; however, it is now thought that this sequence represents a transcript variant of CYP3A4. Alternatively-spliced transcript variants encoding different isoforms have been identified.[3]

Function

CYP3A4 is a member of the cytochrome P450 superfamily of enzymes. The cytochrome P450 proteins are monooxygenases that catalyze many reactions involved in drug metabolism and synthesis of steroids (including cholesterol), and other lipids.[3]

The CYP3A4 protein localizes to the endoplasmic reticulum, and its expression is induced by glucocorticoids and some pharmacological agents.[3] Cytochrome P450 enzymes metabolize approximately 60% of prescribed drugs, with CYP3A4 responsible for about half of this metabolism;[4] substrates include acetaminophen (paracetamol), codeine, ciclosporin (cyclosporin), diazepam, erythromycin, and chloroquine.[3] The enzyme also metabolizes some steroids and carcinogens.[5] Most drugs undergo deactivation by CYP3A4, either directly or by facilitated excretion from the body. Also, many substances are bioactivated by CYP3A4 to form their active compounds, and many protoxins are toxicated into their toxic forms (see table below for examples).

CYP3A4 also possesses epoxygenase activity in that it metabolizes arachidonic acid to epoxyeicosatrienoic acids (EETs), i.e. (±)-8,9-, (±)-11,12-, and (±)-14,15-epoxyeicosatrienoic acids.[6] EETs have a wide range of activities including the promotion of certain types of cancers (see epoxyeicosatetraenoic acid). CYP3A4 promotes the growth of various types of human cancer cell lines in culture by producing (±)-14,15-epoxyeicosatrienoic acids, which stimulate these cells to grow.[7] The CYP3A4 enzyme is also reported to have fatty acid monooxgenase activity for metabolizing arachidonic acid to 20-Hydroxyeicosatetraenoic acid (20-HETE).[8] 20-HETE has a wide range of activities that include growth stimulation in breast and other types of cancers (see 12-hydroxyeicosatetraenoic acid).

Evolution

The CYP3A4 gene exhibits a much more complicated upstream regulatory region in comparison with its paralogs.[9] This increased complexity renders the CYP3A4 gene more sensitive to endogenous and exogenous PXR and CAR ligands, instead of relying on gene variants for wider specificity.[9] Chimpanzee and human CYP3A4 are highly conserved in metabolism of many ligands, although four amino acids positively selected in humans led to a 5-fold benzylation of 7-BFC in the presence of the hepatotoxic secondary bile acid lithocholic acid.[10] This change in consequence contributes to an increased human defense against cholestasis.[10]

Tissue distribution

Fetuses do not express CYP3A4 in their liver tissue, but rather CYP3A7 (EC 1.14.14.1), which acts on a similar range of substrates. CYP3A4 increases to approximately 40% of adult levels in the fourth month of life and 72% at 12 months.[11][12]

Although CYP3A4 is predominantly found in the liver, it is also present in other organs and tissues of the body, where it may play an important role in metabolism. CYP3A4 in the intestine plays an important role in the metabolism of certain drugs. Often this allows prodrugs to be activated and absorbed, as in the case of the histamine H1-receptor antagonist terfenadine.

Recently CYP3A4 has also been identified in the brain, but its role in the central nervous system is still unknown.[13]

Mechanisms

Cytochrome P450 enzymes perform an assortment of modifications on a variety of ligands, utilizing its large active site and its ability to bind more than one substrate at a time to perform complicated chemical alterations in the metabolism of endogenous and exogenous compounds. These include hydroxylation, epoxidation of olefins, aromatic oxidation, heteroatom oxidations, N- and O- dealkylation reactions, aldehyde oxidations, dehydrogenation reactions, and aromatase activity.[14][15]

Hydroxylation of an sp3 C-H bond is one of the ways in which CYP3A4 (and cytochrome P450 oxygenases) affects its ligand.[16] In fact, hydroxylation is sometimes followed by dehydrogenation, leading to more complex metabolites.[15] An example of a molecule that undergoes more than one reaction due to CYP3A4 includes tamoxifen, which is hydroxylated to 4-hydroxy-tamoxifen and then dehydrated to 4-hydroxy-tamoxifen quinone methide.[15]

Two mechanisms have been proposed as the primary pathway of hydroxylation in P450 enzymes.

Two of the most commonly proposed mechanisms used for the hydroxylation of an sp3 C–H bond

The first pathway suggested is a cage-controlled radical method ("oxygen rebound"), and the second involves a concerted mechanism that does not utilize a radical intermediate but instead acts very quickly via a "radical clock".[16]

Inhibition through fruit ingestion

In 1998, various researchers showed that grapefruit juice, and grapefruit in general, is a potent inhibitor of CYP3A4, which can affect the metabolism of a variety of drugs, increasing their bioavailability.[17][18][19][20][21] In some cases, this can lead to a fatal interaction with drugs like astemizole or terfenadine.[18] The effect of grapefruit juice with regard to drug absorption was originally discovered in 1989. The first published report on grapefruit drug interactions was in 1991 in the Lancet entitled "Interactions of Citrus Juices with Felodipine and Nifedipine", and was the first reported food-drug interaction clinically. The effects of grapefruit last from 3–7 days, with the greatest effects when juice is taken an hour previous to administration of the drug.[22]

In addition to grapefruit, other fruits have similar effects. Noni (Morinda citrifolia), for example, is a dietary supplement typically consumed as a juice and also inhibits CYP3A4.[23] Pomegranate juice has shown some inhibition in limited studies, but has not yet demonstrated the effect in humans.[24][25]

Variability

While over 28 single nucleotide polymorphisms (SNPs) have been identified in the CYP3A4 gene, it has been found that this does not translate into significant interindividual variability in vivo. It can be supposed that this may be due to the induction of CYP3A4 on exposure to substrates.

CYP3A4 alleles that have been reported to have minimal function compared to wild-type include CYP3A4*6 (an A17776 insertion) and CYP3A4*17 (F189S). Both of these SNPs led to decreased catalytic activity with certain ligands, including testosterone and nifedipine in comparison to wild-type metabolism.[26] By contrast, CYP3A4*1G allele has more potent enzymatic activity compared to CYP3A4*1A (the wild-type allele).[27]

Variability in CYP3A4 function can be determined noninvasively by the erythromycin breath test (ERMBT). The ERMBT estimates in vivo CYP3A4 activity by measuring the radiolabelled carbon dioxide exhaled after an intravenous dose of (14C-N-methyl)-erythromycin.[28]

Induction

CYP3A4 is induced by a wide variety of ligands. These ligands bind to the pregnane X receptor (PXR). The activated PXR complex forms a heterodimer with the retinoid X receptor (RXR), which binds to the XREM region of the CYP3A4 gene. XREM is a regulatory region of the CYP3A4 gene, and binding causes a cooperative interaction with proximal promoter regions of the gene, resulting in increased transcription and expression of CYP3A4. Activation of the PXR/RXR heterodimer initiates transcription of the CYP3A4 promoter region and gene. Ligand binding increases when in the presence of CYP3A4 ligands, such as in the presence of aflatoxin B1, M1, and G1. Indeed, due to the enzyme's large and malleable active site, it is possible for the enzyme to bind multiple ligands at once, leading to potentially detrimental side effects.[29]

Induction of CYP3A4 has been shown to vary in humans depending on sex. Evidence shows an increased drug clearance by CYP3A4 in women, even when accounting for differences in body weight. A study by Wolbold et al. (2003) found that the median CYP3A4 levels measured from surgically removed liver samples of a random sample of women exceeded CYP3A4 levels in the livers of men by 129%. CYP3A4 mRNA transcripts were found in similar proportions, suggesting a pre-translational mechanism for the up-regulation of CYP3A4 in women. The exact cause of this elevated level of enzyme in women is still under speculation, however studies have elucidated other mechanisms (such as CYP3A5 or CYP3A7 compensation for lowered levels of CYP3A4) that affect drug clearance in both men and women.[30]

CYP3A4 substrate activation varies amongst different animal species. Certain ligands activate human PXR, which promotes CYP3A4 transcription, while showing no activation in other species. For instance, mouse PXR is not activated by rifampicin and human PXR is not activated by pregnenolone 16α-carbonitrile[31] In order to facilitate study of CYP3A4 functional pathways in vivo, mouse strains have been developed using transgenes in order to produce null/human CYP3A4 and PXR crosses. Although humanized hCYP3A4 mice successfully expressed the enzyme in their intestinal tract, low levels of hCYP3A4 were found in the liver.[31] This effect has been attributed to CYP3A4 regulation by the growth hormone signal transduction pathway.[31] In addition to providing an in vivo model, humanized CYP3A4 mice (hCYP3A4) have been used to further emphasize gender differences in CYP3A4 activity.[31]

CYP3A4 activity levels have also been linked to diet and environmental factors, such as duration of exposure to xenobiotic substances.[32] Due to the enzyme's extensive presence in the intestinal mucosa, the enzyme has shown sensitivity to starvation symptoms and is upregulated in defense of adverse effects. Indeed, in fatheaded minnows, unfed female fish were shown to have increased PXR and CYP3A4 expression, and displayed a more pronounced response to xenobiotic factors after exposure after several days of starvation.[32] By studying animal models and keeping in mind the innate differences in CYP3A4 activation, investigators can better predict drug metabolism and side effects in human CYP3A4 pathways.

Turnover

Estimates of the turnover rate of human CYP3A4 vary widely. For hepatic CYP3A4, in vivo methods yield estimates of the enzyme half-life mainly in the range of 70 to 140 hours, whereas in vitro methods give estimates from 26 to 79 hours.[33] Turnover of gut CYP3A4 is likely to be a function of the rate of enterocyte renewal; an indirect approach based on the recovery of activity following exposure to grapefruit juice yields measurements in the 12- to 33-hour range.[33]

Technology

Due to membrane-bound CYP3A4's natural propensity to conglomerate, it has historically been difficult to study drug binding in both solution and on surfaces. Co-crystallization is difficult since the substrates tend to have a low KD (between 5–150 μM) and low solubility in aqueous solutions.[34] A successful strategy in isolating the bound enzyme is the functional stabilization of monomeric CYP3A4 on silver nanoparticles produced from nanosphere lithography and analyzed via localized surface plasmon resonance spectroscopy (LSPR).[35] These analyses can be used as a high-sensitivity assay of drug binding, and may become integral in further high-throughput assays utilized in initial drug discovery testing. In addition to LSPR, CYP3A4-Nanodisc complexes have been found helpful in other applications including solid-state NMR, redox potentiometry, and steady-state enzyme kinetics.[35]

Substrates, inhibitors and inducers

Following is a table of selected substrates, inducers and inhibitors of CYP3A4. Where classes of agents are listed, there may be exceptions within the class.

Inhibitors of CYP3A4 can be classified by their potency, such as:

  • Strong inhibitor being one that causes at least a 5-fold increase in the plasma AUC values, or more than 80% decrease in clearance.[36]
  • Moderate inhibitor being one that causes at least a 2-fold increase in the plasma AUC values, or 50–80% decrease in clearance.[36]
  • Weak inhibitor being one that causes at least a 1.25-fold but less than 2-fold increase in the plasma AUC values, or 20–50% decrease in clearance.[36]
Selected substrates, inhibitors and inducers of CYP3A4[37]
Substrates Inhibitors Inducers
(not azithromycin)[36]
(not pravastatin)[36]
(not rosuvastatin)[36]

Strong

Moderate

Weak

Unspecified potency

Strong potency

Weak potency

Unspecified potency

Interactive pathway map

See also

  • List of drugs affected by grapefruit

References

  1. "Gene structure of CYP3A4, an adult-specific form of cytochrome P450 in human livers, and its transcriptional control". European Journal of Biochemistry 218 (2): 585–95. December 1993. doi:10.1111/j.1432-1033.1993.tb18412.x. PMID 8269949. 
  2. "Assignment of the human cytochrome P-450 nifedipine oxidase gene (CYP3A4) to chromosome 7 at band q22.1 by fluorescence in situ hybridization". The Japanese Journal of Human Genetics 37 (2): 133–8. June 1992. doi:10.1007/BF01899734. PMID 1391968. 
  3. 3.0 3.1 3.2 3.3 Template:NCBI RefSeq
  4. "Cytochrome P450 enzymes in drug metabolism: regulation of gene expression, enzyme activities, and impact of genetic variation". Pharmacology & Therapeutics 138 (1): 103–41. April 2013. doi:10.1016/j.pharmthera.2012.12.007. PMID 23333322. 
  5. EntrezGene 1576
  6. "Lipid-metabolizing CYPs in the regulation and dysregulation of metabolism". Annual Review of Nutrition 34: 261–79. 2014. doi:10.1146/annurev-nutr-071813-105747. PMID 24819323. http://researchonline.rvc.ac.uk/id/eprint/9922/. 
  7. "The pharmacology of the cytochrome P450 epoxygenase/soluble epoxide hydrolase axis in the vasculature and cardiovascular disease". Pharmacological Reviews 66 (4): 1106–40. October 2014. doi:10.1124/pr.113.007781. PMID 25244930. 
  8. "HET0016, a potent and selective inhibitor of 20-HETE synthesizing enzyme". British Journal of Pharmacology 133 (3): 325–9. June 2001. doi:10.1038/sj.bjp.0704101. PMID 11375247. 
  9. 9.0 9.1 "The unique complexity of the CYP3A4 upstream region suggests a nongenetic explanation of its expression variability". Pharmacogenetics and Genomics 20 (3): 167–78. March 2010. doi:10.1097/FPC.0b013e328336bbeb. PMID 20147837. 
  10. 10.0 10.1 "Ligand diversity of human and chimpanzee CYP3A4: activation of human CYP3A4 by lithocholic acid results from positive selection". Drug Metabolism and Disposition 37 (6): 1328–33. June 2009. doi:10.1124/dmd.108.024372. PMID 19299527. 
  11. "Prediction of the clearance of eleven drugs and associated variability in neonates, infants and children". Clinical Pharmacokinetics 45 (9): 931–56. 2006. doi:10.2165/00003088-200645090-00005. PMID 16928154. 
  12. "Development of CYP2D6 and CYP3A4 in the first year of life". Clinical Pharmacology and Therapeutics 83 (5): 670–1. May 2008. doi:10.1038/sj.clpt.6100327. PMID 18043691. 
  13. "Transgenic mouse models of human CYP3A4 gene regulation". Molecular Pharmacology 64 (1): 42–50. July 2003. doi:10.1124/mol.64.1.42. PMID 12815159. 
  14. "Mechanisms of enhanced oral availability of CYP3A4 substrates by grapefruit constituents. Decreased enterocyte CYP3A4 concentration and mechanism-based inactivation by furanocoumarins". Drug Metabolism and Disposition 25 (11): 1228–33. November 1997. PMID 9351897. 
  15. 15.0 15.1 15.2 "Conformational dynamics of CYP3A4 demonstrate the important role of Arg212 coupled with the opening of ingress, egress and solvent channels to dehydrogenation of 4-hydroxy-tamoxifen". Biochimica et Biophysica Acta (BBA) - General Subjects 1820 (10): 1605–17. October 2012. doi:10.1016/j.bbagen.2012.05.011. PMID 22677141. 
  16. 16.0 16.1 "Mechanism of oxidation reactions catalyzed by cytochrome p450 enzymes". Chemical Reviews 104 (9): 3947–80. September 2004. doi:10.1021/cr020443g. PMID 15352783. 
  17. "Inactivation of cytochrome P450 3A4 by bergamottin, a component of grapefruit juice". Chemical Research in Toxicology 11 (4): 252–9. April 1998. doi:10.1021/tx970192k. PMID 9548795. 
  18. 18.0 18.1 "Grapefruit juice-drug interactions". British Journal of Clinical Pharmacology 46 (2): 101–10. August 1998. doi:10.1046/j.1365-2125.1998.00764.x. PMID 9723817. 
  19. "Effect of grapefruit juice on carbamazepine bioavailability in patients with epilepsy". Clinical Pharmacology and Therapeutics 64 (3): 286–8. September 1998. doi:10.1016/S0009-9236(98)90177-1. PMID 9757152. 
  20. "Interactions between grapefruit juice and cardiovascular drugs". American Journal of Cardiovascular Drugs 4 (5): 281–97. 2004. doi:10.2165/00129784-200404050-00002. PMID 15449971. 
  21. "Grapefruit juice and drug interactions. Exploring mechanisms of this interaction and potential toxicity for certain drugs". Geriatrics 61 (11): 12–8. November 2006. PMID 17112309. 
  22. "Duration of effect of grapefruit juice on the pharmacokinetics of the CYP3A4 substrate simvastatin". Clinical Pharmacology and Therapeutics 68 (4): 384–90. October 2000. doi:10.1067/mcp.2000.110216. PMID 11061578. 
  23. "Integrative Medicine, Noni". Memorial Sloan-Kettering Cancer Center. http://www.mskcc.org/cancer-care/herb/noni. 
  24. "Effects of pomegranate juice on human cytochrome p450 3A (CYP3A) and carbamazepine pharmacokinetics in rats". Drug Metabolism and Disposition 33 (5): 644–8. May 2005. doi:10.1124/dmd.104.002824. PMID 15673597. 
  25. "Pomegranate Juice does not Affect the Bioavailability of Cyclosporine in Healthy Thai Volunteers". Curr Clin Pharmacol 15 (2): 145–151. 2020. doi:10.2174/1574884715666200110153125. PMID 31924158. 
  26. "Functionally defective or altered CYP3A4 and CYP3A5 single nucleotide polymorphisms and their detection with genotyping tests". Pharmacogenomics 6 (4): 357–71. June 2005. doi:10.1517/14622416.6.4.357. PMID 16004554. https://zenodo.org/record/1236271. 
  27. 27.0 27.1 Alkattan A, Alsalameen E. Polymorphisms of genes related to phase-I metabolic enzymes affecting the clinical efficacy and safety of clopidogrel treatment. Expert Opin Drug Metab Toxicol. 2021 Apr 30. doi: 10.1080/17425255.2021.1925249. Epub ahead of print. PMID 33931001.
  28. "Noninvasive tests of CYP3A enzymes". Pharmacogenetics 4 (4): 171–84. August 1994. doi:10.1097/00008571-199408000-00001. PMID 7987401. https://cdr.lib.unc.edu/record/uuid:8d5c6276-d350-4216-b108-276f2be1d59d. 
  29. "Aflatoxins upregulate CYP3A4 mRNA expression in a process that involves the PXR transcription factor". Toxicology Letters 205 (2): 146–53. August 2011. doi:10.1016/j.toxlet.2011.05.1034. PMID 21641981. 
  30. "Sex is a major determinant of CYP3A4 expression in human liver". Hepatology 38 (4): 978–88. October 2003. doi:10.1053/jhep.2003.50393. PMID 14512885. 
  31. 31.0 31.1 31.2 31.3 "CYP3A4 and pregnane X receptor humanized mice". Journal of Biochemical and Molecular Toxicology 21 (4): 158–62. 2007. doi:10.1002/jbt.20173. PMID 17936928. https://zenodo.org/record/1229206. 
  32. 32.0 32.1 "Influence of gender, feeding regimen, and exposure duration on gene expression associated with xenobiotic metabolism in fathead minnows (Pimephales promelas)". Comparative Biochemistry and Physiology. Toxicology & Pharmacology 154 (3): 208–12. September 2011. doi:10.1016/j.cbpc.2011.05.016. PMID 21664292. 
  33. 33.0 33.1 "Cytochrome p450 turnover: regulation of synthesis and degradation, methods for determining rates, and implications for the prediction of drug interactions". Current Drug Metabolism 9 (5): 384–94. June 2008. doi:10.2174/138920008784746382. PMID 18537575. 
  34. "Structural and mechanistic insights into the interaction of cytochrome P4503A4 with bromoergocryptine, a type I ligand". The Journal of Biological Chemistry 287 (5): 3510–7. January 2012. doi:10.1074/jbc.M111.317081. PMID 22157006. 
  35. 35.0 35.1 "Screening of type I and II drug binding to human cytochrome P450-3A4 in nanodiscs by localized surface plasmon resonance spectroscopy". Analytical Chemistry 81 (10): 3754–9. May 2009. doi:10.1021/ac802612z. PMID 19364136. 
  36. 36.000 36.001 36.002 36.003 36.004 36.005 36.006 36.007 36.008 36.009 36.010 36.011 36.012 36.013 36.014 36.015 36.016 36.017 36.018 36.019 36.020 36.021 36.022 36.023 36.024 36.025 36.026 36.027 36.028 36.029 36.030 36.031 36.032 36.033 36.034 36.035 36.036 36.037 36.038 36.039 36.040 36.041 36.042 36.043 36.044 36.045 36.046 36.047 36.048 36.049 36.050 36.051 36.052 36.053 36.054 36.055 36.056 36.057 36.058 36.059 36.060 36.061 36.062 36.063 36.064 36.065 36.066 36.067 36.068 36.069 36.070 36.071 36.072 36.073 36.074 36.075 36.076 36.077 36.078 36.079 36.080 36.081 36.082 36.083 36.084 36.085 36.086 36.087 36.088 36.089 36.090 36.091 36.092 36.093 36.094 36.095 36.096 36.097 36.098 36.099 36.100 36.101 36.102 36.103 36.104 36.105 36.106 36.107 36.108 36.109 36.110 36.111 36.112 36.113 36.114 36.115 36.116 36.117 36.118 36.119 36.120 36.121 36.122 36.123 36.124 36.125 "Drug Interactions: Cytochrome P450 Drug Interaction Table". Indiana University School of Medicine. 2007. http://medicine.iupui.edu/flockhart/table.htm.  Retrieved on 25 December 2008.
  37. Where classes of agents are listed, there may be exceptions within the class
  38. 38.00 38.01 38.02 38.03 38.04 38.05 38.06 38.07 38.08 38.09 38.10 38.11 38.12 38.13 38.14 38.15 38.16 38.17 38.18 38.19 38.20 38.21 38.22 38.23 38.24 38.25 38.26 38.27 38.28 38.29 38.30 38.31 38.32 38.33 38.34 38.35 38.36 38.37 38.38 38.39 38.40 38.41 38.42 38.43 38.44 38.45 38.46 38.47 38.48 38.49 38.50 38.51 38.52 38.53 38.54 38.55 38.56 38.57 38.58 38.59 38.60 38.61 38.62 38.63 38.64 38.65 38.66 38.67 38.68 FASS (drug formulary): Swedish environmental classification of pharmaceuticals Facts for prescribers (Fakta för förskrivare). Retrieved July 2011
  39. 39.0 39.1 "Rinvoq: EPAR – Public assessment report" (PDF). European Medicines Agency. 5 March 2020. Archived (PDF) from the original on 21 July 2020. Retrieved 21 July 2020.
  40. 40.0 40.1 Austria-Codex (in German). Vienna: Österreichischer Apothekerverlag. 2020. Rinvoq 15 mg Retardtabletten.
  41. "Erlotinib". https://www.drugs.com/ppa/erlotinib.html. "Metabolized primarily by CYP3A4 and, to a lesser degree, by CYP1A2 and the extrahepatic isoform CYP1A1" 
  42. "Cyclobenzaprine". DrugBank. http://www.drugbank.ca/drugs/DB00924. 
  43. "Lurasidone". StatPearls. Treasure Island (FL): StatPearls Publishing. 2022. http://www.ncbi.nlm.nih.gov/books/NBK541057/. Retrieved 14 October 2022. 
  44. "Effect of rifampin and nelfinavir on the metabolism of methadone and buprenorphine in primary cultures of human hepatocytes". Drug Metabolism and Disposition 37 (12): 2323–9. December 2009. doi:10.1124/dmd.109.028605. PMID 19773542. 
  45. "CYP2D6 and CYP3A4 involvement in the primary oxidative metabolism of hydrocodone by human liver microsomes". British Journal of Clinical Pharmacology 57 (3): 287–97. Mar 2004. doi:10.1046/j.1365-2125.2003.02002.x. PMID 14998425. 
  46. "Clinically significant pharmacokinetic drug interactions with benzodiazepines". Journal of Clinical Pharmacy and Therapeutics 24 (5): 347–355. October 1999. doi:10.1046/j.1365-2710.1999.00247.x. PMID 10583697. 
  47. "Role of CYP3A4 in human hepatic diltiazem N-demethylation: inhibition of CYP3A4 activity by oxidized diltiazem metabolites". The Journal of Pharmacology and Experimental Therapeutics 282 (1): 294–300. July 1997. PMID 9223567. 
  48. "Drug Development and Drug Interactions: Table of Substrates, Inhibitors and Inducers". 25 June 2009. https://www.fda.gov/drugs/developmentapprovalprocess/developmentresources/druginteractionslabeling/ucm093664.htm#table2-2. 
  49. "Grapefruit juice increases felodipine oral availability in humans by decreasing intestinal CYP3A protein expression". The Journal of Clinical Investigation (American Society for Clinical Investigation) 99 (10): 2545–53. May 1997. doi:10.1172/jci119439. PMID 9153299. 
  50. "Erythromycin-felodipine interaction: magnitude, mechanism, and comparison with grapefruit juice". Clinical Pharmacology and Therapeutics (Springer Nature) 60 (1): 25–33. July 1996. doi:10.1016/s0009-9236(96)90163-0. PMID 8689808. 
  51. "Oxidation of dihydropyridine calcium channel blockers and analogues by human liver cytochrome P-450 IIIA4". Journal of Medicinal Chemistry 34 (6): 1838–44. June 1991. doi:10.1021/jm00110a012. PMID 2061924. 
  52. "Inhibitory effects of CYP3A4 substrates and their metabolites on P-glycoprotein-mediated transport". European Journal of Pharmaceutical Sciences 12 (4): 505–13. February 2001. doi:10.1016/s0928-0987(00)00215-3. PMID 11231118. 
  53. "Selection of alternative CYP3A4 probe substrates for clinical drug interaction studies using in vitro data and in vivo simulation". Drug Metabolism and Disposition (American Society for Pharmacology & Experimental Therapeutics (ASPET)) 38 (6): 981–7. June 2010. doi:10.1124/dmd.110.032094. PMID 20203109. 
  54. "Grapefruit juice-nifedipine interaction: possible involvement of several mechanisms". Journal of Clinical Pharmacy and Therapeutics 30 (2): 153–8. April 2005. doi:10.1111/j.1365-2710.2004.00618.x. PMID 15811168. 
  55. "NIFEDIPINE EXTENDED RELEASE- nifedipine tablet, extended release". 29 November 2012. https://dailymed.nlm.nih.gov/dailymed/drugInfo.cfm?setid=4617417a-08df-4417-a944-dfc30de183db. "Drug Interactions: Nifedipine is mainly eliminated by metabolism and is a substrate of CYP3A. Inhibitors and inducers of CYP3A can impact the exposure to nifedipine and, consequently, its desirable and undesirable effects. In vitro and in vivo data indicate that nifedipine can inhibit the metabolism of drugs that are substrates of CYP3A, thereby increasing the exposure to other drugs. Nifedipine is a vasodilator, and coadministration of other drugs affecting blood pressure may result in pharmacodynamic interactions." 
  56. "Overlapping substrate specificities of cytochrome P450 3A and P-glycoprotein for a novel cysteine protease inhibitor". Drug Metabolism and Disposition 26 (4): 360–6. April 1998. PMID 9531525. 
  57. "The effect of three different oral doses of verapamil on the disposition of theophylline". European Journal of Clinical Pharmacology 43 (1): 35–8. 1992. doi:10.1007/bf02280751. PMID 1505606. 
  58. "Verapamil-induced inhibition of theophylline elimination in healthy humans". Pharmacology & Toxicology 66 (2): 101–3. February 1990. doi:10.1111/j.1600-0773.1990.tb00713.x. PMID 2315261. 
  59. "The effect of verapamil on the pharmacokinetic disposition of theophylline in cigarette smokers". Journal of Clinical Pharmacology 29 (8): 728–32. August 1989. doi:10.1002/j.1552-4604.1989.tb03407.x. PMID 2778093. 
  60. "Effect of calcium channel blockers on theophylline disposition". Clinical Pharmacology and Therapeutics 44 (1): 29–34. July 1988. doi:10.1038/clpt.1988.108. PMID 3391002. 
  61. "Selective inhibitory effects of nifedipine and verapamil on oxidative metabolism: effects on theophylline". British Journal of Clinical Pharmacology 25 (3): 397–400. March 1988. doi:10.1111/j.1365-2125.1988.tb03319.x. PMID 3358901. 
  62. "Substrate-selective inhibition by verapamil and diltiazem: differential disposition of antipyrine and theophylline in humans". The Journal of Pharmacology and Experimental Therapeutics 244 (3): 994–9. March 1988. PMID 3252045. 
  63. 63.0 63.1 "Inhibitory potencies of 1,4-dihydropyridine calcium antagonists to P-glycoprotein-mediated transport: comparison with the effects on CYP3A4". Pharmaceutical Research 17 (10): 1189–97. October 2000. doi:10.1023/a:1007568811691. PMID 11145223. 
  64. "Active ingredient: Tadalafil - Brands, Medical Use, Clinical Data". Druglib.com. http://www.druglib.com/activeingredient/tadalafil/. 
  65. "Bicalutamide: clinical pharmacokinetics and metabolism". Clinical Pharmacokinetics 43 (13): 855–78. 2004. doi:10.2165/00003088-200443130-00003. PMID 15509184. 
  66. 66.0 66.1 "[Adrenal crisis associated with modafinil use]" (in es). Medicina 81 (5): 846–849. 2021. PMID 34633961. 
  67. "Iatrogenic Cushing's syndrome with inhaled fluticasone". Australian Prescriber 42 (4): 139–140. August 2019. doi:10.18773/austprescr.2019.040. PMID 31427846. 
  68. "Involvement of multiple human cytochromes P450 in the liver microsomal metabolism of astemizole and a comparison with terfenadine". British Journal of Clinical Pharmacology 51 (2): 133–42. February 2001. doi:10.1046/j.1365-2125.2001.01292.x. PMID 11259984. 
  69. Enzyme 1.14.13.32 at KEGG
  70. "Showing Protein Cytochrome P450 3A4 (HMDBP01018)". Human Metabolome Database. http://www.hmdb.ca/proteins/HMDBP01018. 
  71. "Pharmacogenetics of oral anticoagulants". Pharmacogenetics 13 (5): 247–52. May 2003. doi:10.1097/00008571-200305000-00002. PMID 12724615. 
  72. "Atorvastatin reduces the ability of clopidogrel to inhibit platelet aggregation: a new drug-drug interaction". Circulation 107 (1): 32–7. January 2003. doi:10.1161/01.CIR.0000047060.60595.CC. PMID 12515739. 
  73. "Ketamine-derived designer drug methoxetamine: metabolism including isoenzyme kinetics and toxicological detectability using GC-MS and LC-(HR-)MSn". Analytical and Bioanalytical Chemistry 405 (19): 6307–21. July 2013. doi:10.1007/s00216-013-7051-6. PMID 23774830. 
  74. "LOSARTAN- losartan potassium tablet, film coated". 26 December 2018. https://dailymed.nlm.nih.gov/dailymed/drugInfo.cfm?setid=a98a821c-7b81-4f9b-9801-1a16d71871ce. 
  75. 75.0 75.1 75.2 75.3 75.4 75.5 "Drug Development and Drug Interactions: Table of Substrates, Inhibitors and Inducers". FDA. 26 May 2021. https://www.fda.gov/drugs/drug-interactions-labeling/drug-development-and-drug-interactions-table-substrates-inhibitors-and-inducers. 
  76. 76.0 76.1 76.2 76.3 76.4 76.5 Rang & Dale's pharmacology. Edinburgh: Churchill Livingstone. 2007. ISBN 978-0-443-06911-6. [page needed]
  77. "Evidence-Based Guidelines for Drug Interaction Studies: Model-Informed Time Course of Intestinal and Hepatic CYP3A4 Inhibition by Clarithromycin". AAPS J 23 (5): 104. August 2021. doi:10.1208/s12248-021-00632-7. PMID 34467456. 
  78. "Dose-dependent inhibition of CYP3A activity by clarithromycin during Helicobacter pylori eradication therapy assessed by changes in plasma lansoprazole levels and partial cortisol clearance to 6beta-hydroxycortisol". Clin Pharmacol Ther 72 (1): 33–43. July 2002. doi:10.1067/mcp.2002.125559. PMID 12152002. 
  79. "Drug Interactions of Tetrahydrocannabinol and Cannabidiol in Cannabinoid Drugs: Recommendations for Clinical Practice". Dtsch Arztebl Int 120 (49): 833–840. December 2023. doi:10.3238/arztebl.m2023.0223. PMID 37874128. 
  80. 80.0 80.1 80.2 "Interaction potential between clarithromycin and individual statins-A systematic review". Basic Clin Pharmacol Toxicol 126 (4): 307–317. April 2020. doi:10.1111/bcpt.13343. PMID 31628882. "Erythromycin 500 mg three-four times daily for 6-7 days markedly increased lovastatin exposure (≈6-fold increase in AUC)". 
  81. "Chloramphenicol is a potent inhibitor of cytochrome P450 isoforms CYP2C19 and CYP3A4 in human liver microsomes". Antimicrobial Agents and Chemotherapy 47 (11): 3464–9. November 2003. doi:10.1128/AAC.47.11.3464-3469.2003. PMID 14576103. 
  82. 82.00 82.01 82.02 82.03 82.04 82.05 82.06 82.07 82.08 82.09 82.10 82.11 82.12 82.13 82.14 82.15 82.16 Center for Drug Evaluation and Research. "Drug Interactions & Labeling - Drug Development and Drug Interactions: Table of Substrates, Inhibitors and Inducers". https://www.fda.gov/drugs/developmentapprovalprocess/developmentresources/druginteractionslabeling/ucm093664.htm. 
  83. Darweesh, Ruba S.; El-Elimat, Tamam; Zayed, Aref; Khamis, Tareq N.; Babaresh, Wahby M.; Arafat, Tawfiq; Al Sharie, Ahmed H. (2020). "The effect of grape seed and green tea extracts on the pharmacokinetics of imatinib and its main metabolite, N-desmethyl imatinib, in rats". BMC Pharmacology and Toxicology 21 (1): 77. doi:10.1186/s40360-020-00456-9. PMID 33198812. 
  84. Nishikawa, Masataka; Ariyoshi, Noritaka; Kotani, Akira; Ishii, Itsuko; Nakamura, Hiroyoshi; Nakasa, Hiromitsu; Ida, Mayuri; Nakamura, Hideo et al. (2004). "Effects of Continuous Ingestion of Green Tea or Grape Seed Extracts on the Pharmacokinetics of Midazolam". Drug Metabolism and Pharmacokinetics 19 (4): 280–289. doi:10.2133/dmpk.19.280. PMID 15499196. http://www.sciencedirect.com/science/article/pii/S1347436715308302. 
  85. Wanwimolruk, S.; Wong, K.; Wanwimolruk, P. (2009). "Variable inhibitory effect of different brands of commercial herbal supplements on human cytochrome P-450 CYP3A4". Drug Metabolism and Drug Interactions 24 (1): 17–35. doi:10.1515/dmdi.2009.24.1.17. PMID 19353999. http://pubmed.ncbi.nlm.nih.gov/19353999/. 
  86. Francis Carballo-Arce, Ana; Raina, Vikrant; Liu, Suqi; Liu, Rui; Jackiewicz, Victoria; Carranza, David; Arnason, John Thor; Durst, Tony (2019). "Potent CYP3A4 Inhibitors Derived from Dillapiol and Sesamol". ACS Omega 4 (6): 10915–10920. doi:10.1021/acsomega.9b00897. PMID 31460189. 
  87. Briguglio, M.; Hrelia, S.; Malaguti, M.; Serpe, L.; Canaparo, R.; Dell'Osso, B.; Galentino, R.; De Michele, S. et al. (2018). "Food Bioactive Compounds and Their Interference in Drug Pharmacokinetic/Pharmacodynamic Profiles". Pharmaceutics 10 (4): 277. doi:10.3390/pharmaceutics10040277. PMID 30558213. 
  88. Kondža, M.; Bojić, M.; Tomić, I.; Rezić, V.; Ćavar, I. (2021). "Characterization of the CYP3A4 Enzyme Inhibition Potential of Selected Flavonoids". Molecules (Basel, Switzerland) 26 (10): 3018. doi:10.3390/molecules26103018. PMID 34069400. 
  89. "Modulatory effects of rutin on the expression of cytochrome P450s and antioxidant enzymes in human hepatoma cells". Acta Pharmaceutica 66 (4): 491–502. December 2016. doi:10.1515/acph-2016-0046. PMID 27749250. 
  90. "Inhibition of Cytochrome P450 (CYP3A4) Activity by Extracts from 57 Plants Used in Traditional Chinese Medicine (TCM)". Pharmacognosy Magazine 13 (50): 300–308. 2017. doi:10.4103/0973-1296.204561. PMID 28539725. 
  91. Product Information: ORAVIG(R) buccal tablets, miconazole buccal tablets. Praelia Pharmaceuticals, Inc (per FDA), Cary, NC, 2013.
  92. "Bergamottin a CYP3A inhibitor found in grapefruit juice inhibits prostate cancer cell growth by downregulating androgen receptor signaling and promoting G0/G1 cell cycle block and apoptosis". PLOS ONE 16 (9): e0257984. 2021. doi:10.1371/journal.pone.0257984. PMID 34570813. Bibcode2021PLoSO..1657984V. 
  93. "Valerian: Health Benefits, Side Effects, Uses, Dose & Precautions". http://www.rxlist.com/valerian-page3/supplements.htm#Interactions. 
  94. "The Intracellular Mechanism of Berberine-Induced Inhibition of CYP3A4 Activity". Curr Pharm Des 27 (40): 4179–4185. 2021. doi:10.2174/1381612827666210715155809. PMID 34269665. 
  95. "An Integrative Approach to Elucidate Mechanisms Underlying the Pharmacokinetic Goldenseal-Midazolam Interaction: Application of In Vitro Assays and Physiologically Based Pharmacokinetic Models to Understand Clinical Observations". J Pharmacol Exp Ther 387 (3): 252–264. December 2023. doi:10.1124/jpet.123.001681. PMID 37541764. 
  96. "Clinical evidence of herbal drugs as perpetrators of pharmacokinetic drug interactions". Planta Medica 78 (13): 1458–77. September 2012. doi:10.1055/s-0032-1315117. PMID 22855269. 
  97. "The enhancement of cardiotoxicity that results from inhibiton of CYP 3A4 activity and hERG channel by berberine in combination with statins". Chemico-Biological Interactions 293: 115–123. September 2018. doi:10.1016/j.cbi.2018.07.022. PMID 30086269. 
  98. "Interaction of buprenorphine and its metabolite norbuprenorphine with cytochromes p450 in vitro". Drug Metabolism and Disposition 31 (6): 768–72. June 2003. doi:10.1124/dmd.31.6.768. PMID 12756210. 
  99. "Interaction of coffee diterpenes, cafestol and kahweol, with human P-glycoprotein". AAPS Journal. The American Association of Pharmaceutical Scientists. 2009. http://www.aapsj.org/abstracts/AM_2009/AAPS2009-001235.PDF. 
  100. "Determination of effective concentrations of drug absorption enhancers using in vitro and ex vivo models". Eur J Pharm Sci 167: 106028. December 2021. doi:10.1016/j.ejps.2021.106028. PMID 34601070. 
  101. "In vitro evaluation of valproic acid as an inhibitor of human cytochrome P450 isoforms: preferential inhibition of cytochrome P450 2C9 (CYP2C9)". Br J Clin Pharmacol 52 (5): 547–53. 2001. doi:10.1046/j.0306-5251.2001.01474.x. PMID 11736863. 
  102. "Potent inhibition of human cytochrome P450 3A isoforms by cannabidiol: role of phenolic hydroxyl groups in the resorcinol moiety". Life Sciences 88 (15–16): 730–6. April 2011. doi:10.1016/j.lfs.2011.02.017. PMID 21356216. 
  103. 103.0 103.1 Non-nucleoside reverse-transcriptase inhibitors have been shown to both induce and inhibit CYP3A4.
  104. "Potent inhibition by star fruit of human cytochrome P450 3A (CYP3A) activity". Drug Metabolism and Disposition 32 (6): 581–3. June 2004. doi:10.1124/dmd.32.6.581. PMID 15155547. 
  105. "HCVadvocate.org". Archived from the original on 5 March 2010. https://web.archive.org/web/20100305175124/http://www.hcvadvocate.org/hepatitis/hepC/mthistle.html. 
  106. "Inhibition of human P450 enzymes by nicotinic acid and nicotinamide". Biochemical and Biophysical Research Communications 317 (3): 950–6. May 2004. doi:10.1016/j.bbrc.2004.03.137. PMID 15081432. 
  107. "Inhibitory effects of polyphenols on human cytochrome P450 3A4 and 2C9 activity". Food and Chemical Toxicology 48 (1): 429–35. January 2010. doi:10.1016/j.fct.2009.10.041. PMID 19883715. "Ginko Biloba has been shown to contain the potent inhibitor amentoflavone". 
  108. "Sesamin: A Naturally Occurring Lignan Inhibits CYP3A4 by Antagonizing the Pregnane X Receptor Activation". Evidence-Based Complementary and Alternative Medicine 2012: 242810. 2012. doi:10.1155/2012/242810. PMID 22645625. 
  109. "Piperine, a major constituent of black pepper, inhibits human P-glycoprotein and CYP3A4". The Journal of Pharmacology and Experimental Therapeutics 302 (2): 645–50. August 2002. doi:10.1124/jpet.102.034728. PMID 12130727. 
  110. "Isoniazid is a mechanism-based inhibitor of cytochrome P450 1A2, 2A6, 2C19 and 3A4 isoforms in human liver microsomes". European Journal of Clinical Pharmacology 57 (11): 799–804. January 2002. doi:10.1007/s00228-001-0396-3. PMID 11868802. 
  111. Ekstein, D.; Schachter, S. C. (2010). "Natural Products in Epilepsy—the Present Situation and Perspectives for the Future". Pharmaceuticals (Basel, Switzerland) 3 (5): 1426–1445. doi:10.3390/ph3051426. PMID 27713311. 
  112. "Highlights of Prescribing Information: XTANDI (enzalutamide) capsules for oral use". Astellas Pharma US, Inc.. U.S. Food and Drug Administration. August 2012. https://www.accessdata.fda.gov/drugsatfda_docs/label/2012/203415lbl.pdf. 
  113. "Antiepileptic drug interactions - principles and clinical implications". Current Neuropharmacology 8 (3): 254–67. September 2010. doi:10.2174/157015910792246254. PMID 21358975. 
  114. "Dose-dependent induction of cytochrome P450 (CYP) 3A4 and activation of pregnane X receptor by topiramate". Epilepsia 44 (12): 1521–8. December 2003. doi:10.1111/j.0013-9580.2003.06203.x. PMID 14636322. 
  115. "Capsaicin induces CYP3A4 expression via pregnane X receptor and CCAAT/enhancer-binding protein β activation". Molecular Nutrition & Food Research 56 (5): 797–809. May 2012. doi:10.1002/mnfr.201100697. PMID 22648626. 

External links

This article incorporates text from the United States National Library of Medicine, which is in the public domain.