Abstract
Information is scarce regarding pharmacokinetic-based herb-drug interactions (HDI) with trans-cinnamaldehyde (CA) and 2-methoxycinnamaldehyde (MCA), components of cinnamon. Given the presence of cinnamon in food and herbal treatments for various diseases, HDIs involving the CYP2A6 substrates nicotine and letrozole with MCA (KS = 1.58 µM; Hill slope = 1.16) and CA were investigated. The time-dependent inhibition (TDI) by MCA and CA of CYP2A6-mediated nicotine metabolism is a complex process involving multiple mechanisms. Molecular dynamic simulations showed that CYP2A6’s active site accommodates two dynamic ligands. The preferred binding orientations for MCA and CA were consistent with the observed metabolism: epoxidation, O-demethylation, and aromatic hydroxylation of MCA and cinnamic acid formation from CA. The percent remaining activity plots for TDI by MCA and CA were curved, and they were analyzed with a numerical method using models of varying complexity. The best-fit models support multiple inactivator binding, inhibitor depletion, and partial inactivation. Deconvoluted mass spectra indicated that MCA and CA modified CYP2A6 apoprotein with mass additions of 156.79 (142.54–171.04) and 132.67 (123.37–141.98), respectively, and it was unaffected by glutathione. Heme degradation was observed in the presence of MCA (48.5% ± 13.4% loss; detected by liquid chromatography–tandem mass spectrometry). In the absence of clinical data, HDI predictions were made for nicotine and letrozole using inhibition parameters from the best-fit TDI models and parameters scaled from rats. Predicted area under the concentration-time curve fold changes were 4.29 (CA-nicotine), 4.92 (CA-letrozole), 4.35 (MCA-nicotine), and 5.00 (MCA-letrozole). These findings suggest that extensive exposure to cinnamon (corresponding to ≈ 275 mg CA) would lead to noteworthy interactions.
Significance Statement Human exposure to cinnamon is common because of its presence in food and cinnamon-based herbal treatments. Little is known about the risk for cinnamaldehyde and methoxycinnamaldehyde, two components of cinnamon, to interact with drugs that are eliminated by CYP2A6-mediated metabolism. The interactions with CYP2A6 are complex, involving multiple-ligand binding, time-dependent inhibition of nicotine metabolism, heme degradation, and apoprotein modification. An herb-drug interaction prediction suggests that extensive exposure to cinnamon would lead to noteworthy interactions with nicotine.
Introduction
Cinnamon has been used for centuries as a spice and in traditional medicine (Ranasinghe et al., 2013). Large doses of cinnamon powder (1–10 g) have been reported for treating diabetes (Pham et al., 2007; Crawford, 2009; Kirkham et al., 2009; Akilen et al., 2010), and the components of cinnamon are under investigation for a wide variety of other maladies, including inflammation (Hagenlocher et al., 2016; Schink et al., 2018; Cao et al., 2019; Cheng et al., 2020), multiple sclerosis (Pahan, 2015; Kundu et al., 2016), and pain (Burgess and Williams, 2010; Yokoyama et al., 2011; Weyer-Menkhoff and Lötsch, 2018, 2019). Considering its use in food (Friedman et al., 2000; Rietjens et al., 2020) and as a complementary treatment, there is potential for components in cinnamon to act as precipitants for herb-drug interactions (HDIs).
trans-Cinnamaldehyde (CA) and 2-methoxycinnamaldehyde (MCA) (Fig. 1) are present in cinnamon and are used for flavoring in the food industry (Rietjens et al., 2020). CA is also often present in fluids for electronic nicotine delivery systems (Behar et al., 2016, 2018). Previously, we observed that CA is a time-dependent inhibitor of CYP2A6 (Chan et al., 2016). Compared with reversible inhibition, time-dependent inhibition (TDI) is especially concerning for HDI and drug-drug interactions (DDIs) because it can result in prolonged or irreversible inhibition due to long-lived metaboliteintermediate complexes (MICs) or enzyme inactivation by metabolites that covalently bind to either the apoprotein or heme prosthetic group essential for enzyme function (Mohutsky and Hall, 2014).
We hypothesized that MCA would inhibit CYP2A6 and investigated the kinetics and mechanisms of inhibition in comparison with CA. Traditionally, the replot method has been used to determine the TDI parameters, KI and kinact, from in vitro experiments. This involves graphing the percent remaining activity (PRA) and inactivation rates (kobs) as a function of inhibitor concentration. KI and kinact are determined by fitting to the equation kobs = (kinact• [I])/(KI + [I]) (Silverman, 1995). The approach is based on several assumptions: Michaelis-Menten/steady-state kinetics, irreversible inactivation, measurement of initial rates, and binding of a single substrate. In vitro–in vivo extrapolation (IVIVE) models use KI and kinact to assess the potential for DDIs and HDIs.
Previously, TDI parameters generated using the replot method and a static IVIVE model predicted a 3-fold change in nicotine AUC, with CA as a precipitant (Chan et al., 2016). However, the replot method overlooks several mechanistic complexities observed with P450 kinetics (e.g., multiple-ligand binding) and TDI (partial inactivation, reversible MIC formation). Oftentimes, only the linear portion of PRA plots is used to estimate KI and kinact, as longer incubations can result in nonlinear plots, which is indicative of mechanistic complexity (e.g., non–Michaelis-Menten kinetics), but also makes interpretation challenging with the replot method. Thus, the replot method can overpredict the likelihood for clinically relevant DDIs as a result of inaccuracies in parameter estimates, at least partially due to overlooking P450 TDI mechanistic complexity (Korzekwa et al., 2014). Numerical analysis was shown to lower error in parameter estimates in comparison with the replot method (Nagar et al., 2014).
Numerical analysis is a more recently developed approach for evaluating TDIs for DDI predictions (Yadav et al., 2020). It can account for diverse mechanistic complexities associated with P450 TDI and other factors (lipid partitioning). The mechanistic processes and kinetic models of varying complexity are described using ordinary differential equations, which are solved simultaneously to estimate inhibition parameters by fitting kinetic data to the models. Fit quality for individual models can provide evidence to support proposed mechanisms that contribute to a specific DDI/HDI scenario. Coupling information from numerical analysis with results from experiments designed to discover structure-based mechanistic information can further strengthen support for a proposed mechanism. Previously, using the standard replot method to analyze the TDI of CYP2A6 by MCA, the kobs data fit best to a sigmoidal model, which provides evidence for the involvement of more complex mechanisms and atypical kinetics (manuscript under review). Here, we aimed to determine TDI parameters for the inhibition of nicotine metabolism by CA and MCA using numerical analysis, to assess potential mechanisms of inactivation, and to predict the potential for an HDI using IVIVE. We also evaluated the mechanisms of CYP2A6 interactions with CA and MCA using structure-based approaches. We analyzed binding affinity, inhibitor orientation, and inhibitor dynamics using spectral binding assays and molecular dynamic computational approaches. The inhibitory mechanism was investigated by measuring heme loss, which was observed previously with CA (manuscript under review), and apoprotein modification and by identifying CYP2A6-generated metabolites of MCA.
Materials and Methods
Materials.
(−)-Cotinine, (±)-cotinine-D3, trans-cinnamaldehyde, 2-methoxycinnamaldehyde, and reduced glutathione were obtained from Sigma-Aldrich. S-(−)-nicotine and NADPH tetrasodium salt were purchased from Alfa Aesar and EMD Chemicals, respectively. Pooled (N = 50 livers; 28 male and 22 female) human liver microsomes (HLMs) were obtained from Gibco. Human liver cytosol and NADPH solution A and solution B were obtained from Corning Life Sciences. High performance liquid chromatography columns were obtained from Phenomenex and Waters, as indicated below.
Expression and Purification of CYP2A6 and Rat Cytochrome P450 Reductase.
CYP2A6 was expressed and purified as previously reported, with some slight modifications (Stephens et al., 2012; Chan et al., 2016). Spectral ligand binding assays, metabolite identification, apoprotein, and heme analysis experiments were conducted using modified recombinant human CYP2A6, which had an N-terminal transmembrane sequence truncation (∆2–23) and a C-terminal His4-tag. The recombinant protein was expressed from pKK2A6dH in DH5α Escherichia coli with a 72-hour induction time in δ-amino levulinic acid–supplemented Terrific broth media. The protein was purified using Ni affinity chromatography, followed by carboxymethyl Sepharose cation exchange chromatography (Stephens et al., 2012). The detergent included in purification buffers was Brij L23. P450 content was determined from the carbon monoxide difference spectrum (Omura and Sato, 1962, 1964; Guengerich et al., 2009). Rat P450 reductase was expressed and purified as previously reported (Shen et al., 1989).
Molecular Dynamic Studies of CYP2A6-CA and CYP2A6-MCA Interactions.
The three-dimensional structures of the ligand molecules CA and MCA were built, minimized, and geometry optimized in Molecular Operating Environment (MOE) (2019) from Chemical Computing Group (https://www.chemcomp.com/Products.htm). Charge calculation and atom types were assigned by the CHARMM General Force Field small-molecule parameterization module.
Molecular docking and dynamics studies were conducted using the published X-ray crystal structure of coumarin-bound CYP2A6 (PDB: 1Z10) (Yano et al., 2005); resolution 1.9 Å). Missing transmembrane helix residues 1–29 were obtained as the FASTA sequence from UniProt (P11509-1) and constructed through homology modeling with MODELLER 9.19 (Šali and Blundell, 1993; Webb and Sali, 2016) using the intact X-ray crystal structure of the transmembrane domain of CYP51 [PDB: 5EQB (Monk et al., 2014)]. Of 100 models successfully generated by the software, the top-scoring model with the most favorable discrete optimized protein energy score was chosen for simulation. The fully constructed protein was prepared using the “Structure Preparation” module in MOE (missing atoms, residues, and H atoms were added; protonation states of residues were assigned using Protonate 3D within MOE) and then inserted and oriented within the 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine membrane using the Orientations of Proteins in Membranes web server (Lomize et al., 2012). The reoriented structure was measured to have a heme tilt angle and a transmembrane helix angle of 67.495° and 35.876° versus the bilayer normal (z-axis), respectively.
As guidance for docking CA and MCA to CYP2A6, 28 known CYP2A6 substrates or inhibitors with reported Ki values and binding modes were docked to identify the optimal combination of parameter settings within MOE for placement and refinement scoring (i.e., by reproducing the greatest number of crystal poses) for an induced-fit docking process. This pool included reversible inhibitors coumarin (PDB: 1Z10) and pilocarpine (PDB: 3T3Q), as well as irreversible inhibitors methoxsalen (PDB: 1Z11), aldrithiol (4,4′-dipyridyl disulfide, PDB: 2FDY), N,N-dimethyl(5-(pyridin-3-yl)furan-2-yl)methanamine (PDB: 2FDU), 5-(pyridin-3-yl)furan-2-yl)methanamine (PDB: 2FDW), and N-methyl(5-(pyridin-3-yl)furan-2-yl)methanamine (PDB: 2FDV). MMFF94X optimized force field for small molecules was chosen for docking within MOE. Usage of the London dG scoring function, the Triangle Matcher placement method, and Generalized-Born Volume Integral/Weighted Surface Area dG refinement scoring methods was found to reproduce the best poses for the studied molecules, as determined based on overlay root-mean-square deviation (RMSD) and adoption of the expected binding modes. Both CA and MCA were docked using the above set of parameters, and docked structures were used as the starting points for running all-atom MD simulations. In addition, a structure-based pharmacophore model (Fig. 2) was generated using the superpositioned crystal structures to obtain essential functional features. This model was used for selecting the best poses in addition to the docking scores.
All MD simulations were performed using the NAMD package (Phillips et al., 2005) under isothermal/isobaric NPT (number of particles; pressure constant; temperature constant) ensemble with periodic boundary conditions. The CHARMM36 force field (Huang and MacKerell, 2013), CHARMM General Force Field (Vanommeslaeghe et al., 2010, 2012; Vanommeslaeghe and MacKerell, 2012), and transferable intermolecular potential with 3 points (TIP3P) water model (Jorgensen et al., 1983) were employed for the protein and membrane, including monovalent ions, ligands, and solvent molecules, respectively. The CHARMM-GUI (Lee et al., 2016) membrane builder plug-in was used to construct the model membrane, which consisted of 90 and 105 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine lipids and 10 cholesterol molecules in the upper and lower leaflets, respectively. The CYP2A6-ligand complex oriented bilayer (Supplemental Fig. 1) was solvated with a minimum water thickness of 22.5 Å on each side.
The net charge of the membrane-CYP2A6-ligand complex systems was kept neutral (with 0.15 M KCl, i.e., 52 K+ and 60 Cl− ions) and subjected to a multistage minimization and equilibration protocol. In the first stage, the solvent molecules and ions were briefly minimized (2500 steps of steepest descent minimization followed by another 2500 steps of conjugate gradient method) to remove any bad contacts with the complexes. The rest of the system was position-restrained using a force constant of 100 kcal/(mol × Å2). This step was followed by another 5000 steps of minimization of the whole solvated complex. The readjustment of membrane and solvent molecules to the potential field of the protein-ligand complex was achieved by subsequent heating and equilibration steps. The system was heated to a final temperature of 310 K in three steps; Langevin dynamics (Feller et al., 1995) was used in all stages to control the temperature using a collision frequency of 1.0 picoseconds−1. The final solvent equilibration step was performed for 50 picoseconds under NPT conditions as above. The SHAKE algorithm (Ryckaert et al., 1977) was used to constrain bonds involving hydrogen, allowing a time step of 2 femtoseconds. Lennard-Jones and electrostatic interactions were calculated explicitly within a cutoff of 1.2 nm, and long-range electrostatic interactions were calculated by particle mesh Ewald summation (Ulrich et al., 1995). For both CA and MCA, the production simulation was carried out for 150 nanoseconds, and the trajectory files were written for every 5000 steps, resulting in 15,000 frames for each molecule.
The resulting trajectories were analyzed for potential H-bond interactions as well as the distance between ligand atoms and the heme iron atom using Hbonds plug-in Visual Molecular Dynamics software with the following criteria for the H-bonds: the distance between the donor atom (D) and the acceptor atom (A) is less than the cutoff distance of 3.5 Å, and the angle D-H-A is less than the cutoff angle of 40°.
Spectral Binding Studies with MCA and Purified CYP2A6.
Samples of 1 µM purified rCYP2A6 in 50 mM potassium phosphate buffer (pH = 7.4) were titrated with aliquots of MCA dissolved in DMSO. The final DMSO concentration in the titrations was kept below 1%. During the titration, solvent-only additions of equal volume (that is, equal to the volume added to the sample cuvette containing MCA) were added to a reference cuvette containing an identical protein sample. Difference spectra were recorded on a Shimadzu UV-visible 1601 spectrophotometer, and the ∆absorbance between the minimum and maximum was measured. Equilibrium dissociation constants (KS) were determined from nonlinear least-squares fits to two equations: one-site specific ligand binding and specific binding with Hill slope using GraphPad Prism 8 (La Jolla, CA). Fits were compared using Akaike’s Information Criterion (AICc). The experiment was repeated on three different days (N = 3), and each individual experiment was analyzed independently for comparison of model fits, as well as KS and Hill coefficient determination.
CYP2A6-Mediated Metabolism of 2-Methoxycinnamaldehyde and Metabolite Identification by LC-MS.
Purified rCYP2A6 (1 µM) was reconstituted with rat cytochrome P450 reductase (1 µM) and incubation buffer (100 mM potassium phosphate buffer; pH = 7.4) in microcentrifuge tubes. The incubations were preheated at 37°C for 5 minutes after the addition of MCA (200 µM) and reduced glutathione (5 mM) to trap epoxide metabolites that could form. A 100 mM solution of MCA (in DMSO) was diluted in incubation buffer to make a 2 mM working solution, which was further diluted to the final concentration used in the incubations. The working stock of reduced glutathione was made in incubation buffer. Incubations (100 µl final volume) were initiated with NADPH (1 mM), heated at 37°C for 1 hour in a water bath, and terminated with an equal volume of acetonitrile:methanol (1:1) and centrifuged. The supernatants were removed for LC-MS analysis, and those samples were analyzed using an AB Sciex 5600 TripleTOFTM system and mass defect filtering. Details of the method and data analysis are described in Supplemental Material 4.
Time-Dependent Inhibition of Nicotine Metabolism by CA and MCA in Human Liver Microsomes and Cytosol and LC-MS/MS Analysis of Nicotine and Cotinine.
Time-dependent inhibition of CYP2A6 was tested using a standard two-step approach using pooled (N = 50) human liver microsomes (Yadav et al., 2018) with slight modifications. Briefly, 12 concentrations of either CA (0–1000 μM) or MCA (0–250 μM) were incubated at 37°C with a 5 mg/ml suspension of HLM in 100 mM potassium phosphate buffer (pH = 7.4) as a primary incubation. After 7 minutes of warming the primary incubation, the reaction was initiated by addition of NADPH regenerating system (final concentration 1.3 mM NADP+, 3.3 mM glucose-6-phosphate, 0.4 U/ml glucose-6-phosphate dehydrogenase, and 3.3 mM magnesium chloride). The NADPH regenerating system was used rather than NADPH salt to avoid the inhibition of P450 reductase at later time points by NADP+ accumulation. At specific time points, an aliquot (10 μl) of the primary incubation was added to 190 μl of secondary incubation. The secondary incubation consisted of nicotine (75 μM), NADPH (1 mM), and human liver cytosol (1 mg/ml). The conditions for the secondary incubation were developed based on previous studies of nicotine metabolism (Nakajima et al., 1996; Messina et al., 1997; Yamazaki et al., 1999; Le Gal et al., 2003). Since nicotine is metabolized by cytochrome P450 to nicotine-Δ1′(5′) -iminium ion, which is further metabolized to cotinine by aldehyde oxidase (Murphy, 1973; Brandänge and Lindblom, 1979; Hukkanen et al., 2005);, cytosol was added to the secondary incubation to facilitate cotinine formation. The primary incubation was run for 0–90 minutes, and the secondary incubation was allowed to run for 10 minutes, followed by quenching with ice-cold acidified acetonitrile containing cotinine-D3 as the internal standard. Each assay was conducted in triplicate. After centrifugation at 2000 rpm for 10 minutes, the supernatant was analyzed for cotinine by LC-MS/MS.
LC-MS/MS Analysis of Cotinine.
Calibration curves were prepared in 0.25 mg/ml HLM in phosphate buffer (pH = 7.4) spiked with analyte standards, followed by precipitation with acetonitrile containing the internal standard cotinine-D3. The supernatant was analyzed with LC-MS/MS. For chromatographic separation of cotinine, a Phenomenex Synergi 4-µm Polar-RP 80-Å column (50 × 2 mm) was used. The sample volume was 10 µl. An AB Sciex API 5500 LC-MS/MS system was used to analyze samples in positive ion mode using the following multiple reaction monitoring transitions: 176.979–80.016 m/z for cotinine and 179.985–80.022 m/z for cotinine-D3. LC-MS solvents consisted of 10 mM ammonium formate in water (pH = 4.0) as aqueous mobile phase (A) and 0.1% formic acid in acetonitrile as an organic mobile phase (B). The flow rate was 1 ml/min, with a gradient elution programmed from 5% to 95% B in 0.20 minutes maintained at 90% until 0.4 minutes and returned to initial condition at 0.8 minutes and maintained until 1.1 minutes. The total run time was 1.1 minutes, and the retention time was 0.61 minutes.
Microsomal Partitioning of CA and MCA.
Nonspecific microsomal partitioning and human plasma protein binding were estimated using the model developed by Gao et al. (2008). The model was developed using molecular descriptors including calculated log P, calculated log D, pKa, Kier connectivity, shape, and E-state indices, a subset of MOE descriptors, and a set of absorption, distribution, metabolism, and excretion (ADME) features. The statistical model was derived using the Cubist committee plus composite models, which is a tool for generating rule-based predictive models (Gao et al., 2008).
Model Fitting Using the Numerical Method and Estimate of TDI Parameters for CA and MCA.
Cotinine concentrations obtained from the in vitro TDI experiments were converted to log (PRA) plots and further evaluated for model development. The concave upward curvature is indicative of either quasi-irreversible or partial inactivation, as shown previously (Nagar et al., 2014). Using the numerical method (Korzekwa et al., 2014; Nagar et al., 2014), kinetic models were fit to the data, and kinetic parameters were estimated. The initial estimates of the rate constants were obtained from analyzing the data as detailed in previous publications (Korzekwa et al., 2014; Nagar et al., 2014; Barnaba et al., 2016; Yadav et al., 2018, 2020). Association rate constants were fixed at 270 μM−1 min−1 (Barnaba et al., 2016). For nicotine, the association (k1) and dissociation rate constants (k2) were fixed at 270 μM−1 min−1 and 4050 min−1, respectively, based on the reported KM of nicotine (Yamazaki et al., 1999; Le Gal et al., 2003). Lipid partitioning was also incorporated in the models to account for microsomal partitioning (Yadav et al., 2019). The association rate constant for lipid was set at 2000 μM min−1, and the dissociation rate constant was calculated using the following equation (Nagar and Korzekwa, 2017):where kon is the association rate constant, koff is the dissociation rate constant, and fu,mic is the unbound fraction in the microsomes.
As previously mentioned, curvature in the PRA plots indicates either MIC formation or partial inactivation (Nagar et al., 2014). Curved PRA plots are also observed in cases of inhibitor depletion (Yadav et al., 2020). Several models were developed to capture the inactivation kinetics of CA. Further, a competitive inhibition model was fit to 0- and 60-minute time point data to obtain an initial estimate for KI. A difference in initial estimates of KI from 0 versus 60 minutes was indicative of multiple binding. As shown previously (Barnaba et al., 2016), MIC formation is a complex multistep process involving the formation of Fe+3:carbene and Fe2+:carbene. Hence, enzyme inactivation was modeled with three types of rate constants: e.g., in Fig. 8A, k6 and k11 for Fe+3:carbene formation, k7 for the reformation of active enzyme, and k8 for Fe2+:carbene formation. The second type of quasi-irreversible model was also evaluated, which involved two steps denoted as k6 and k11 for Fe+2:carbene formation and k7 for the reformation of active enzyme (Rodgers et al., 2018). Partial inactivation (PI) models were also tested, including partial inactivation with inhibitor depletion, double binding with partial inactivation, and the sequential formation of an intermediate leading to partial inactivation. The kinetic scheme for the simplest partial inactivation model is shown in Fig. 8C (Nagar et al., 2014).
Assuming rapid equilibrium, KI values were estimated from ratios of association and dissociation rate constants. KI obtained from the numerical method is the same as unbound KI,u (KI,u = KI) since lipid partitioning was incorporated in the model. Net kinact was calculated using the net rate constant concept (Cleland, 1975). For MCA, net kinact was calculated by using the following equation, which was derived using the net rate constant concept for the EII-PI-M-IL model:whereWhereas for CA, kinact was calculated by using the following equation for EII-MIC-M-IL:where
The replot method was also used to analyze the in vitro TDI data sets. The linear portion of the PRA plots was used to calculate the slopes (kobs). The estimates of KI and kinact were obtained using the following replot equation:Although post hoc correction for lipid partitioning can lead to errors (Yadav et al., 2019), for comparison of replot and numerical method, unbound KI was obtained by post hoc multiplying the KI obtained from the above equation.
Parameter errors for net rate constants were calculated with error propagation for individual rate constants. AICc (Akaike, 1974) and adjusted R2 were used to compare different models for each data set. Model fitting was conducted with Mathematica 11.3.0.0 (Wolfram Research, Champagne, IL) using the NonlinearModelFit function to fit the model to the data with PrecisionGoal = 10, finite difference derivatives with an order of 4, and 1/Y weighting.
Herb-Drug Interaction Prediction Using In Vitro–In Vivo Extrapolation.
HDI predictions using KI,u and kinact obtained from the numerical method were performed using the following mechanistic static equations (https://www.fda.gov/media/134582/download). HDI predictions were performed assuming no substrate-dependent interactions by using nicotine and letrozole as CYP2A6 substrates. The estimated exposure of CA is between 8 and 275 mg (Friedman et al., 2000; Kirkham et al., 2009; Chan et al., 2016). Hence, a 275-mg dose was used to predict all the CA and MCA interactions with a dosing interval of 24 hours. For HDI predictions with MCA, KIu2 and the net kinact calculated (as shown above) were used. CYP2A6 is not known to be present in gut; therefore, only the systemic interaction was predicted for oral dosing of nicotine and letrozole. The following equations were used to predict AUC changes for the probe substrate dosed through the oral route:where AUCi and AUC are the areas under the plasma concentration–time curve of the probe substrate in the presence and absence of the inactivator, respectively; fm,CYP2A6 is the fraction of nicotine metabolized by CYP2A6;[I]h is the inactivator concentration in the hepatic portal vein (defined below); and kdeg,h is the degradation rate constant of hepatic CYP2A6 (Yang et al., 2008). Kiu is the unbound competitive inhibition constant. Kiu was obtained from the literature by assuming competitive inhibition (Chan et al., 2016). Ih,u was given by the following equations (Rostami-Hodjegan and Tucker, 2004; Obach et al., 2007):where Imax is the maximum concentration of the inactivator in plasma after oral dosing, D is the dose, Fa is the fraction absorbed, Fg is the fraction escaping gut metabolism, ka is the first-order absorption rate constant, Qh (1.5 l/h) is the hepatic blood flow, fu,p is the unbound fraction in the plasma, fu,g (assumed to be equal to 1) is the unbound fraction in the enterocyte, and BP is the blood-to-plasma partition ratio. Maximum unbound plasma concentration Imax,u was given by the following equationswhere τ is the dosing interval, and k is elimination rate constant, which is calculated by the following equation:where CLs is the systemic clearance, and Vss is the volume of distribution at steady state.
Average unbound systemic concentration Isys,u is the unbound average systemic inactivator concentration:where F is the bioavailability.
Human pharmacokinetic parameters for CA were predicted from rats to humans. Rat clearance and volume of distribution were obtained from the literature (Yuan et al., 1992). Human clearance was predicted using the following equation (Tang et al., 2007):
Volume from rats was scaled to human by using the following method (Caldwell et al., 2004):
For MCA, in vitro CLint was predicted using an internal in silico Cubist-based model (Wenzel et al., 2019). The predicted CLint was then scaled using biologic scaling factors [microsomal protein per gram of liver = 45; liver weight = 1800 g (Davies and Morris, 1993) for a 70-kg human] assuming no extra hepatic clearance. In vivo CL for MCA was estimated using the well stirred model (Pang and Rowland, 1977) and the predicted in vitro CLint. Volume of distribution for MCA was predicted using the Rodgers model (Rodgers and Rowland, 2006) using log P of 2.35, fu,p of 0.261, and pka of 14.
Inactivator pharmacokinetic parameters used for calculation of in vivo plasma inactivator concentrations are listed in Table 1.
LC-MS Search for Heme and Protoporphryin IX Adducts with MCA and Metabolites of MCA.
Purified rCYP2A6 (1 µM) was reconstituted with rat cytochrome P450 reductase (1 µM) and incubation buffer (100 mM potassium phosphate buffer; pH 7.4) in microcentrifuge tubes. The P450:reductase ratio for heme adduct analysis and heme degradation was based on the range of ratios used in previous studies (Raner et al., 1997; Kuo et al., 1999; Lin et al., 2005; Foti et al., 2011). The incubations were preheated at 37°C for 5 minutes after the addition of 200 µM MCA, which was estimated to be a saturating concentration based on preliminary TDI experiments. Stock and working solutions of MCA were prepared as above for metabolite identification studies. Incubations were initiated with NADPH (1 mM final concentration) and heated at 37°C for 1 hour in the dark. The incubations were terminated with an equal volume of acetonitrile:methanol (1:1) with 1% formic acid, vortexed, and centrifuged. The supernatants were removed for LC-MS analysis, and those samples were analyzed using an AB Sciex 5600 TripleTOFTM system. Details of the method and analysis are described in Supplemental Material 5.
LC-MS/MS Analysis of Heme Degradation.
Incubations were performed in microcentrifuge tubes in a 37°C water bath. Purified rCYP2A6 (400 pmol) and rat P450 reductase (400 pmol) were combined and left at room temperature for 10 minutes. Buffer (100 mM potassium phosphate; pH = 7.4) and MCA (80 µM) were added and preincubated at 37°C for 5 minutes. The MCA concentration was selected for the purposes of comparison with CA from a previous study (Chan et al., 2016); also, it was estimated that 80 µM approached enzyme saturation based on preliminary TDI experiments. NADPH (1 mM) was added to initiate the incubations, and the samples were heated at 37°C for 5 minutes. Incubations were terminated with an equal volume of acetonitrile, mixed with a vortex mixer, and centrifuged at 11,000g for 5 minutes. Enzyme-only controls (without MCA or NADPH) and controls without NADPH were prepared similarly. The supernatant was removed and was injected (50 µl) onto a Phenomenex Kinetix C18 column (50 × 2.1 mm; 2.6 μm) using an AB Sciex Exion high performance liquid chromatograpy system interfaced with a 3500 triple quadrupole mass spectrometer. Heme was eluted using a gradient of 30% B to 95% B over 6 minutes (A = water/0.1% formic acid; B = acetonitrile/0.1% formic acid; 0.6 ml/min). Heme was detected by UV absorption at 405 nm and by ion intensity using multiple reaction monitoring at m/z = 616/557.2 (parent mass/transition state) in positive mode. Instrument parameters were as follows: ionspray voltage, 5500; declustering potential, 190 V; entrance potential, 10 V; collision energy, 54 eV; collision cell exit potential, 6 V; curtain gas, 30 psi; collision gas, 9 psi; ion source gas 1 and 2, 60 psi; temperature, 550°C. Nitrogen was used as the nebulizer, heater, and auxiliary gas. Heme loss was determined by comparing the change in peak area relative to the enzyme-only or enzyme and NADPH control areas, which were set to 100% of heme content. Statistical significance between individual data sets was determined using a two-tailed t test.
LC-MS/MS Analysis of the CYP2A6 Apoprotein.
rCYP2A6 (60 pmol) and reductase (600 pmol) were combined and preincubated at room temperature for 10 minutes. The P450:reductase ratio was based on previous LC-MS/MS analyses of P450 apoprotein adducts with time-dependent inhibitors (Lin et al., 2005; Baer et al., 2007; Harrelson et al., 2012). Preliminary experiments showed that the detection and deconvolution of the protein envelope for rCYP2A6 was substantially better than reductase, so the relative amount of reductase used here was higher than the previous reports to improve the detection of reductase. Buffer (100 mM potassium phosphate, pH = 7.4) and MCA or CA (200 µM) were added and preincubated at 37°C for 5 minutes. NADPH (1 mM) was added, and the samples were heated at 37°C for 20 minutes. A positive control containing 80 µM 8-methoxypsoralen (8-MOP) was prepared in a similar fashion. All samples were compared with negative controls containing no inhibitor or no NADPH. A portion (10 µl) of the incubation was analyzed on the aforementioned AB Sciex 3500 triple quadrupole LC-MS/MS system using a POROS 10 R1 column (2.1 mm internal diameter × 100 mm; Thermoscientific). Protein was eluted using a gradient of 35% B to 55% B over 4 minutes (A = water/0.1% formic acid; B = acetonitrile/0.1% formic acid; 1.2 ml/min). Protein envelopes were detected using a Q1 scan in positive mode from 1000 to 2000 m/z. Instrument parameters were as follows: ionspray voltage, 4500; declustering potential, 170 V; entrance potential, 3 V; curtain gas, 30 psi; ion source gas 1 and 2, 60 psi; temperature, 25°C. Nitrogen was used as the nebulizer, heater, and auxiliary gas. After the spectra were obtained, deconvolution of the ion envelopes was undertaken by averaging the peak area corresponding to the retention time of CYP2A6 using the Biotool Kit plug-in in the AB Sciex Peakview software suite.
Results
Molecular Dynamics of CYP2A6-CA and CYP2A6-MCA Interactions.
We performed molecular docking to predict the preferred binding modes of CA and MCA within the CYP2A6 binding site, followed by all-atom MD simulations to assess the dynamics and stability of the critical protein-ligand interactions. To guide the docking process, a structure-based pharmacophore model was developed using the available structural data on published crystal structures of several inhibitors bound to CYP2A6 (see Materials and Methods section). The sequence alignment and structure-based superposition revealed important structural features shared among these ligands (Fig. 2). Specifically, all bound substrates were found to overlay in a similar way, in which one end of the molecule occupied the distal axial heme coordination position, and the other was in close proximity to the N297 side chain amido group nitrogen. The majority of the ligands carried a complementary H-bond acceptor group, and all bound substrates contained an aromatic ring, likely to be essential for engaging in π-π interactions with the aromatic side chains of several phenylalanine residues, including F118, F107, F209, F480, and F111. Our docking procedure used the coordinates of these superpositioned ligands as well as the consensus features as a template for guided docking. The best-docked poses generated for CA and MCA were in close agreement with the expected pose, with their aldehyde groups pointing toward the heme iron atom and the benzene ring in close proximity to N297.
The trajectory analyses of 150-nanosecond-long MD simulations for each ligand revealed the major binding modes, critical interactions, and dynamic nature of these molecules within the binding pocket. After approximately 2 nanoseconds into the simulation, CA changed from its initial docked pose and reoriented to a more stable pose in which its carbonyl oxygen engaged in H-bonding with N297 (Fig. 3A). This binding orientation of CA lasted for 76.6 nanoseconds (51.07% of 150 nanoseconds total). The H-bond interaction between the oxygen atom of CA and the side chain amido group of N297 was characterized by an average distance of ∼2.93 Å and an angle of ∼142° (Supplemental Fig. 2A). CA was also stabilized in this pose by a T-shaped π stacking with F209 (an average distance of ∼3.48 Å). The H-bond with N297 was maintained for the remainder of the simulation, although the entry and binding of a water molecule that began to coordinate with the heme iron (at ∼77 nanoseconds) pushed CA’s aromatic end toward the phenylalanine-rich region (Fig. 3B). The π stacking interaction with F107 stabilized CA in this configuration for the rest of the simulation. The dynamic nature of CA within the binding pocket was characterized by the RMSD values in reference to the initial docked pose, calculated for the entire simulation time (Supplemental Fig. 2B).
Interestingly, the MCA molecule also shifted from its docked pose to an entirely different pose after approximately 8 nanoseconds into the simulation. Initially, the aldehyde group was pointing toward the heme iron, and the methoxyl oxygen was in contact with N297 engaging in an H-bond interaction (Fig. 4A). As with CA, the rearrangement was prompted by the entry of several water molecules near N297. This caused MCA to reorient itself to a different pose (Fig. 4B). In this stable pose, which lasted for more than 80% of the total simulation, the oxygen atom of the aldehyde group was engaged in a water-mediated interaction with N297, whereas the phenyl ring was in close proximity to the heme iron atom (∼2.96 Å, Supplemental Fig. 3A). Likewise, the overall dynamics of MCA within the binding pocket was calculated as RMSD values in reference to the initial docked pose for the entire simulation time (Supplemental Fig. 3B).
Multiple-ligand binding and cooperative kinetics (i.e., non–Michaelis-Menten) are widely observed with xenobiotic-metabolizing P450s (Atkins, 2005). Homotropic and heterotropic cooperative kinetic behavior has been observed for CYP2A6-mediated metabolism; for example, the binding of multiple xylene ligands modulated substrate motion and product selectivity (Harrelson et al., 2008). To investigate the possible cooperativity effects in binding for each of the two ligands, we prepared an analogous system with a second copy of the molecule (ligand 2) docked in close proximity to the already-docked CYP2A6-ligand complex. This was possible because of the volume of the binding pocket, which allowed enough space left after the first molecule (ligand 1) was docked. We performed MD simulations on these systems to investigate the effect of dual ligands on the binding pose and interactions of CA and MCA to CYP2A6. For the cinnamaldehyde simulation, the second ligand molecule (CA-2) is less stable than the first one (CA-1) and experienced more dynamic behavior, retaining its initial pose for only 3.24 nanoseconds before its carbonyl oxygen began to respond to the presence of a cluster of water molecules near an access channel by the residue T482 (Fig. 5). At the same time, CA-1 was bound in a highly similar pose to the docked pose, with its aromatic group oriented toward the phenylalanine-rich region. This pose was maintained for 64 nanoseconds (∼43% of the total simulation time) by a network of π stacking interactions from F118, F107, and F209. The aromatic ring of CA-2 was oriented such that it made an edge-edge π-π interaction with F209 (Fig. 5). After ∼65 nanoseconds, the entry of water molecules propelled CA-2 toward the heme, its aromatic ring colliding with the polar end of CA-1. The torque felt by CA-1 displaced the aldehyde group toward N297, and the electrostatic force eventually stabilized the CA-1 in this pose for the remainder of the simulation (86 nanoseconds, 57.48% of the total simulation time) (Supplemental Fig. 2D).
In the case of MCA, the docked pose of MCA-1 was maintained for a much longer duration, held for 58.84 nanoseconds (39% of the simulation time), with supportive contribution from the second molecule, MCA-2 (Supplemental Fig. 3, C and D). Interestingly, the slight rearrangement involves a rotation that places its methoxyl oxygen in an optimal position for hydrogen bonding with N297, while its aldehyde group continued to maintain its positioning above the heme iron (Fig. 6A). This interaction was maintained for 108 nanoseconds until the long-lasting water-mediated interaction between MCA-2 with T309 and T305 finally broke, freeing the MCA-2 carbonyl oxygen, which became attracted to the water cluster at the opposing end of the site (Fig. 6). Despite a slight displacement, MCA-1 continued to be stabilized by the electrostatic interaction of the methoxyl oxygen with N297 and the π interaction of the aromatic benzene ring with F118 (Fig. 6B). Increased dynamics of both ligands in the site were seen shortly after 108 nanoseconds and until 135 nanoseconds, after which MCA-1 steadily reoriented back to the favorable predominant pose (pose assumed between 23 and 100 nanoseconds), which was maintained through 150 nanoseconds (Supplemental Fig. 3D).
Spectral Analysis of MCA Binding to Human CYP2A6.
Spectral binding studies were conducted on three different days with rCYP2A6 to evaluate the binding affinity between MCA and CYP2A6 (Fig. 7). The spectra indicated that MCA is a type I ligand of CYP2A6, consistent with a compound deficient in sp2 hybridized nitrogen atoms. The data were fit to two different models in Prism: one-site specific ligand binding and specific binding with Hill slope. Specific binding with Hill slope was the preferred model for all experiments based on differences in AICc. The probabilities for correctness in comparison with the one-site specific binding model were 99.93%, 62.97%, and 99.99%, with AICc differences of 14.11, 1.06, and 17.73, respectively. Fits to the model using specific binding with Hill slope generated a KS = 1.58 µM (S.D. = 0.03; 95% CI = 1.50–1.66; N = 3), a Hill slope = 1.16 (S.D. = 0.26; 95% CI = 0.50–1.81), and R2 values of 0.9993, 0.9971, and 0.9994. R2 values for the one-site specific binding model were 0.9949, 0.9970, and 0.9958.
CYP2A6-Mediated Metabolism of 2-Methoxycinnamaldehyde and Metabolite Identification by LC-MS.
The metabolism of MCA was investigated using a reconstituted system with rCYP2A6 and P450 reductase (i.e., the same system used for heme analysis). Based on parent masses, fragmentation patterns, and mass defect analysis, three metabolites were detected from CYP2A6-mediated oxidative metabolism: demethylation of the 2-methoxy group, trapped as a glutathione conjugate (m/z = 456.1428 and 456.1436); epoxidation of the double bond between the α- and β-carbons (m/z = 178.0630); and ring hydroxylation (m/z = 178.0630). Epoxidation and ring hydroxylation were clearly delineated from each other through their respective fragmentation patterns and the mass defect of their fragments. Epoxidation resulted in a unique 146.0362 m/z fragment corresponding to a 3,6-dioxatricyclo[5.4.0.02,4]undeca-1(7),4,8,10-tetraene, and the tandem mass spectrometry spectra were devoid of the acrolein fragment seen in the spectra of both MCA and the ring hydroxylated metabolite, indicating side chain modification. The epoxide was also found trapped as a glutathione conjugate, resulting in ring opening and side chain hydroxyl formation (m/z = 486.1531). The ring hydroxylation metabolite was deduced from several fragments, which clearly eliminated the possibility of the presence of hydroxylation on either the side chain or the methyl group. Discerning fragments included the parent minus acetaldehyde (m/z = 136.052), the parent minus a methyl group (m/z = 163.0388), and acrolein (m/z = 55.0171). Representative liquid chromatography traces and mass spectra with fragment structures are available in the Supplemental Material 6.
Time-Dependent Inhibition of CYP2A6-Mediated Nicotine Metabolism by CA and MCA.
Several kinetic models were developed for both of the inactivators to capture the inactivation kinetics. The best-fit model was chosen based on AICc values, R2, and residual analysis. Curved PRA plots indicated either formation of MIC or partial inactivation; hence, both types of models were tested for CA. Although both quasi-irreversible and partial inactivation models result in curved PRA plots, the shapes for both of the mechanisms are different. For quasi-irreversible inactivation, the terminal plateau region is inhibitor concentration–dependent and concentration-independent for partial inactivation (Nagar et al., 2014). Notably, at later time points, activity recovery at lower inhibitor concentrations suggested significant inhibitor loss during the preincubation phase. Graphs of the predicted depletion of each inhibitor are available in Supplemental Figs. 4 and 5.
As expected, Michaelis-Menten models with and without inhibitor depletion gave a poor fit (data not shown, AICc for the Michaelis-Menten model with inhibitor depletion model was −337, and AICc for the Michaelis-Menten model without inhibitor deletion was −295). Preliminary analysis suggested partial inactivation as a better model than a quasi-irreversible model (AICc −536 vs. −511). The single-binding quasi-irreversible with inhibitor depletion model (MIC-M-IL) did not result in a convergence, because of a lack of a reasonable estimate of the second step of inactivation (represented as k8 in Fig. 8A). Therefore, a sensitivity analysis was also performed with a range of k8 values between 0 and 0.005 minutes−1. Equally good fits were observed when values were fixed at ≤0.005 minutes−1. Hence, 0.005 minutes−1 was considered as an upper limit and 0 minutes−1 as the lower limit for k8. Since the model fitting for an EII-MIC-M-IL model showed that k8 → 0, the second type of MIC model, in which inactivation was modeled as two steps, was also tested (Fig. 8B). The modified quasi-irreversible model gave a reasonable fit with an AICc = −495 (data not shown).
Double-binding models were also investigated both with quasi-irreversible and partial inactivation models. A double-binding model along with inhibitor depletion (EII-MIC-M-IL) improved the fit as compared with a single-binding quasi-irreversible model, but not for a partial inactivation model. Addition of a second binding to the partial inactivation model did not improve the model fit in comparison with the single-binding model PI-M-IL (data not shown). AICc and adjusted R2 for the EII-MIC-M-IL model were −528.557 and 0.99872, respectively, whereas for PI-M-IL, the AICc and adjusted R2 were −536.417 and 0.99881, respectively. The fits for all three models are shown in Fig. 8. Figure 8D shows that the EII-MIC-M-IL model fit the CA data set assuming k8 = 0.005 minutes−1. The kinact for CA with an upper limit of k8 as 0.005 minutes−1 was estimated to be 0.0032 ± 0.001 minutes−1 for the EII-MIC-M-IL model (Table 2). The kinact for CA was found to be 0.023 ± 0.012 and 0.1 ± 0.01 minutes−1 for the EII-MIC-M-IL model and PI-M-IL model, respectively. The estimated rate constants for all the models were well defined (Supplemental Table 1). However, the propagation of errors to calculate the net kinact value resulted in a high error (0.023 ± 0.012 and 0.0032 ± 0.001 minutes−1). The estimates of KI,u for CA with different models are shown in Table 2. The estimates of individual rate constants for each model fit are provided in Supplemental Table 1. All three models were able to capture the concave upward curvature observed in the PRA plots reasonably well (Fig. 8, D–F). However, AICc differences suggested PI-M-IL as the better model.
The MCA data set also gave nonlinear PRA plots suggesting either quasi-irreversible or partial inactivation. Preliminary results suggested that inhibitor loss was required to capture the inactivation kinetics. Hence, partial inactivation with inhibitor depletion (PI-M-IL) and quasi-irreversible with inhibitor depletion models (MIC-M-IL) were tested. It was found that a partial inactivation model with inhibitor depletion (PI-M-IL) gave a better fit as compared with a quasi-irreversible model with inhibitor depletion (MIC-M-IL) (Table 3; AICc −608 vs. −583). Further, analysis of the data set suggested a KI value of ∼125 μM, which is inconsistent with the competitive inhibition observed after inhibitor dilution (Ki ∼2.5 μM). Therefore, double-binding models with either quasi-irreversible or partial inactivation were also evaluated.
Biphasic kinetics is observed when kinact2 > kinact1, and inhibition of inactivation is observed when kinact1 > kinact2 (Nagar et al., 2014). Preliminary analysis of MCA data sets suggested KI2 > KI1 and kinact2 > kinact1; however, fitting with constraints on KI1 and kinact2 did not result in model convergence. Using the standard replot method and linear sections of the PRA plot suggested biphasic kinetics with lack of saturation for the second binding. In general, it is difficult to estimate net inactivation and KI2 if kinact1 < < kinact2 and KI2 > > KI1 and/or inhibitor concentrations do not cover the saturation range for KI2. As suggested previously, the slower kinact is more difficult to parameterize (Nagar et al., 2014). Therefore, optimization steps were followed (Nagar et al., 2014). Briefly, KI1 was estimated from the 0-minute incubation time. The value of kinact2 was estimated from the standard replot method. KI2 and kinact1 were optimized while constraining kinact2 and KI1. However, the model was unable to parameterize the second binding KI. Manual optimization showed that the objective function decreased as KI2 became smaller such that KI2 < < KI1. The double-binding model with partial inactivation and inhibitor depletion gave a better fit than the double-binding model with quasi-irreversible and inhibitor depletion (data not shown). The results for three model fits (MIC-M-IL, PI-M-IL, and EII-PI-M-IL) are shown in Table 3 and Fig. 9.
In summary, for CA, partial inactivation with inhibitor depletion (PI-M-IL) was the best-fit model, followed by double-binding models with MIC and inhibitor depletion (EII-MIC-M-IL), in which PI indicates partial inactivation. For MCA, a double-binding model with partial inactivation along with inhibitor depletion gave the best fit (EII-PI-M-IL). PI occurs when modified enzymes retain residual activity (Crowley and Hollenberg, 1995). EII is double-inhibitor binding, and MIC represents metabolite intermediate complex formation. M and IL indicate inhibitor metabolism and lipid partitioning, respectively.
Microsomal Partitioning of CA and MCA.
The predicted fu,mic (0.25 mg/ml) was 0.88 and 0.86 for CA and MCA, respectively. The predicted fu,mic was scaled to 5 mg/ml using an equation described previously (Kalvass and Maurer, 2002):where fu,mic (5 mg/ml) is the scaled unbound fraction at 5 mg/ml microsomal protein, and fu,mic (0.25 mg/ml) is the unbound fraction experimentally measured at 0.25 mg/ml. D is the dilution factor of microsomal concentration between 0.25 and 5 mg/ml, which is 20. The predicted fu,p was 0.19 and 0.26 for CA and MCA, respectively.
Herb-Drug Interaction Prediction for Nicotine.
HDI predictions were performed using the standard static model as recommended by the FDA (https://www.fda.gov/media/134582/download). KI,u and kinact parameters obtained from the PI-M-IL model for CA and EII-PI-M-IL model for MCA were used for HDI predictions. The CYP2A6 half-life of 37 hours was used for the HDI predictions (Renwick et al., 2000). The results for oral dosing are shown in Table 4. The predicted AUC ratios for the CA-nicotine and CA-letrozole interactions were 4.29 and 4.92, respectively. The predicted AUC ratios were 4.35 and 5.00 for the MCA-nicotine and MCA-letrozole interactions, respectively. The predicted interaction suggested that substantial exposure to CA and MCA through cinnamon-containing food, cinnamon-flavored products, and the use of cinnamon powder as an herbal treatment can lead to significant interactions with CYP2A6 substrates. Moreover, CYP2A6 is known to be polymorphic (Raunio et al., 2001; McDonagh et al., 2012; Hosono et al., 2017; Tanner and Tyndale, 2017), which can increase the variability and significance in HDIs (https://www.pharmvar.org/gene/CYP2A6). Notably, the fraction metabolized by CYP2A6 can range from 0.60 to 0.77 for nicotine (Benowitz and Jacob, 1994) and ≈0.80 for letrozole (Sioufi et al., 1997; Murai et al., 2009; Chan et al., 2016).
Analysis of CYP2A6 Heme and Search for Heme/Protoporphryin IX Adducts with MCA and Metabolites of MCA.
To investigate the effect of MCA on heme stability, heme was analyzed after incubation with rCYP2A6 in the presence and absence of MCA and directly measured by LC-MS/MS/UV (616/557 amu; 405 nm). When incubated with MCA and NADPH, 48.5% of the heme peak was lost relative to enzyme-plus-NADPH controls (S.D. = 13.4%; 95% CI = 34.4%–62.6%; P = 0.0082; N = 6) (Fig. 10). No isolable heme adducts were found with UV or by using a full Q1 scan. In an attempt to identify potential adducts of MCA and its metabolites to heme or protoporphryin IX, samples were subjected to mass defect filtering. We have used this approach previously to successfully identify benzyne adducts of CYP2A6 from incubations with 1-aminobenztriazole, a pan inhibitor of cytochrome P450 enzymes (manuscript under review). No adducts of MCA (or metabolites) to the heme or protoporphyrin IX were detected using this approach.
Apoprotein Analysis.
The potential for MCA, CA, and/or respective metabolites to form an adduct with CYP2A6 apoprotein was investigated using purified reconstituted enzyme. As 8-MOP is known to form a covalent adduct with the apoprotein of CYP2A6 (Koenigs and Trager, 1998), it was used as a positive control. A comparison of the deconvoluted spectra from each condition indicated that CA, MCA, and 8-MOP all formed a second detectable ion envelope of CYP2A6 (Fig. 11), which indicated that apoprotein modifications were present in each case (Table 5). Incubations with the inhibitors without NADPH also yielded adduct peaks, although smaller in area in comparison with incubations including NADPH. Incubations without inhibitor yielded a single dominant mass for CYP2A6 of 54,467.47 ± 1.23 Da, corresponding to the theoretical mass of rCYP2A6. Reductase yielded a main peak of 77,724.02 ± 4 Da.
Incubations with MCA yielded an additional mass of 54,622.85 ± 11.89 Da near the CYP2A6 peak, whereas incubations with CA yielded an additional mass of 54,600.05 ± 9.59 Da. There were no statistically significant additions to the reductase peaks in any case. Incubations were also attempted with 5 mM glutathione (N = 6 for each inhibitor). Incubations from these experiments did not yield statistically significant differences from incubations without glutathione, signifying that the reactive metabolites are likely adducting CYP2A6 without leaving the active site. In an attempt to determine whether adduct formation was NADPH-dependent, area ratios of modified CYP2A6 and unmodified CYP2A6 peaks were determined for each incubation condition (Table 5). In the case of both MCA and CA, the peak area ratio—that is, peak area(CYP2A6 mass + inhibitor mass)/peak area(CYP2A6 mass)—was higher in samples containing NADPH than controls without NADPH.
Discussion
Interactions between CYP2A6, MCA, and CA were investigated using a multifaceted approach that was intended to discover mechanistic insights to inform HDI predictions. New information includes a description of MCA-CYP2A6 binding, evidence for the dynamic nature of CA/MCA ligands bound to CYP2A6, and evidence that at least two mechanisms contribute to TDI. Analyzing TDI data with the numerical analysis method allowed for consideration of multiple mechanisms in estimating kinact and KI, which were used for HDI predictions. The results are discussed in order: 1) binding interactions, 2) TDI mechanisms and parameter estimates, and 3) HDI predictions.
Atypical kinetics (i.e., non–Michaelis-Menten) are widely observed with P450s (Korzekwa et al., 1998; Atkins, 2005; Korzekwa, 2014; Leow and Chan, 2019), which have mechanistically been attributed to multiple binding (Korzekwa et al., 1998), protein heterogeneity (Davydov and Halpert, 2008), enzyme oligomerization, protein-protein interactions (Davydov et al., 2017; Davydova et al., 2019), or a combination. Atypical kinetics can profoundly impact TDI kinetics (Korzekwa et al., 2014; Nagar et al., 2014), and evidence suggests that DDI predictions can be improved by accounting for mechanistic complexities (Yadav et al., 2018).
CYP2A6 and CYP2E1 have more restricted active site volumes among human P450s (DeVore et al., 2008; Porubsky et al., 2008); yet there is evidence both enzymes accommodate multiple ligands simultaneously and exhibit homotropic and heterotropic allosteric behavior (Harrelson et al., 2008; Hartman et al., 2012, 2013, 2015). Spectral binding analysis can provide evidence for multiple-ligand binding; it may also overlook some multiple-ligand binding scenarios, as the binding of a second ligand may not always generate changes in the spin state of the heme iron (Davydov et al., 2002; Atkins, 2005). Here, the spectral binding data for MCA fit best to a model that allows for multiple-ligand binding (Hill coefficient >1). The binding data were consistent with results from the numerical method, in which a double-inhibitor model (E•I•I) fit best to the kinetic data. MCA (KS = 1.6 µM) binds with greater affinity than CA to CYP2A6 (KS = 14.9 µM), which was found previously to fit best to a one-site binding model (Chan et al., 2016). Notably, here, the numerical method analysis of the CA TDI data set fit best to a single-inhibitor model.
Although MCA binds with greater affinity than CA, there were some similarities in binding dynamics. MD simulations investigated both single- (E•I) and double-ligand (E•I•I) binding scenarios. In both scenarios, the ligand closest to the heme exhibited two predominant orientations: one with the aldehyde carbonyl oriented to the heme iron (position I) and another with the aromatic ring near the heme iron (position II). Hydrogen bonds with N297 were important in stabilizing the ligand orientations, similar to previous observations (DeVore et al., 2008, 2009). The MD results showed that the CYP2A6 active site is adequately voluminous to accommodate two ligands and that simultaneous binding of two CA or two MCA molecules is possible. The MD results also showed that both inhibitors are dynamic within the active site, and the binding of a second ligand to form the E•I•I complex stabilized the structure of the ligand closest to the heme.
The overall observation that ligand binding is dynamic suggests multiple atoms on the same ligand will be exposed to the active oxidizing species and is consistent with the CYP2A6-mediated metabolism of CA and MCA. For MCA, three metabolic routes were identified: O-demethylation, aromatic hydroxylation, and epoxidation. A predominant binding position for MCA places the aromatic ring in close proximity to the heme iron (i.e., position II; Fig. 4B; Fig. 6B), a favorable position for aromatic hydroxylation and demethylation. Transitory positions that occur as MCA reorients between positions I and II expose the double bond, between the α,β-carbons, to the active oxidizing species, allowing for epoxidation. For CA and MCA, both single-ligand binding and E•I•I also result in orientations in which the carbonyl carbon is in close proximity to the heme (position I). For CA, this is consistent with metabolism studies in which cinnamic acid was the only detectable metabolite (manuscript under review). The metabolism studies were conducted at a saturating substrate concentration to maximize metabolite production. Now that multiple MCA metabolites have been identified, it affords an opportunity to further investigate multiple-ligand binding by evaluating product ratios as a function of MCA concentration.
The accumulated mass spectral evidence supports at least two mechanisms for CYP2A6 inactivation by both CA and MCA: heme degradation and apoprotein modification. Simultaneous modification of heme and apoprotein was observed previously with bergamottin (He et al., 1998; Lin et al., 2005), mifepristone (Lin et al., 2009), and clopidogrel (Zhang et al., 2011). Substantial NADPH-dependent heme loss was observed in the presence of MCA (Fig. 10). Studies of CYP2B4 inactivation by 3-phenylpropionaldehyde, a structural analog of CA, showed evidence of heme adducts with a deformylated metabolite of 3-phenylpropionaldehyde (Kuo et al., 1999). If heme adducts are formed with MCA metabolites, they may be unstable with the conditions used here. Heme loss without detectable heme adducts has also been observed with CA (manuscript under review) as well as loss of CO binding (Chan et al., 2016). The previous evidence that heme adduction/degradation involves the carbonyl of phenylpropanoids may explain the absence of 2-methoxycinnamic acid as a detectable metabolite of MCA. When MCA is metabolized by CYP2A6 at the carbonyl position, heme degradation may be more favored than carboxylic acid metabolite formation.
CA and MCA form covalent adducts with CYP2A6 apoprotein through catalytic and noncatalytic processes, although metabolism enhanced adduction, as modification was maximal with NADPH. Although the variability for the deconvoluted spectra precludes determining exact masses of the adducts, the measured masses and 95% CI (Table 5) suggest structures with molecular weights that are close to the parent masses, CA (mol. wt. 132.16) and MCA (mol. wt. 162.19), without the addition of oxygen. Noncatalytic modification through Michael addition and/or Schiff base formation (Prakash et al., 2008; Chan et al., 2016) would generate adducts similar to the detected masses. For catalytically driven modification, a multistep process would result in the observed adducts: 1) epoxidation between the α,β-carbons; 2) reaction with a nucleophilic residue (e.g., cysteine) to open the epoxide; and 3) dehydration to eliminate water, regenerating the double bond (Niklasson et al., 2014). Notably, epoxide metabolites trapped as glutathione conjugates were identified for MCA (mol. wt. 162) and demethylated MCA (mol. wt. 150), intimating that epoxides are formed from both the parent compound and demethylated metabolite. If both epoxides react with CYP2A6 residues similar to reactions with glutathione, this could explain both why the detected adduct mass was ≈156 (the average of 162 and 150) and the greater variability observed for MCA versus CA. Although glutathione did not change the formation of apoprotein adducts for either inhibitor, it is possible that other nucleophilic trapping agents (e.g., methoxylamine, N-acetyllysine, semicarbizide, potassium cyanide) may capture the reactive metabolites.
The numerical method allows for diverse mechanistic factors to be considered when estimating TDI parameters. CYP2A6 inactivation data were evaluated using several models (>15) that combined different kinetic events, such as inhibitor depletion, quasi-irreversible (MIC) formation, partial inactivation, lipid partitioning, and multiple binding. Consistent with MD simulations, results from the numerical method analysis exhibited support for multiple inactivator binding to CYP2A6 for both CA and MCA. Further, the concave upward curvature in the PRA plots suggested either partial inactivation or quasi-irreversible inactivation. The mass spectral evidence showed heme loss and apoprotein modification as potential contributors to inhibition for both compounds. Apoprotein modification can lead to partial inactivation wherein the enzyme still possesses activity, albeit lower than the unmodified enzyme (Crowley and Hollenberg, 1995).
For both CA and MCA, quasi-irreversible (MIC formation) and partial inactivation models were evaluated. The partial inactivation models with inhibitor depletion (PI-M-IL) resulted in better fits for both inhibitors. CA and MCA do not possess functional groups typically associated with MIC (e.g., amines, methylenedioxy aromatics) (Kalgutkar et al., 2007; Kamel and Harriman, 2013; Mohutsky and Hall, 2014; Limban et al., 2018). For MCA, a divergence between KI and Ki suggested multiple binding kinetics (Korzekwa et al., 2014; Nagar et al., 2014). Analysis of the PRA graphs (linear portions) using the standard replot method suggested biphasic kinetics with lack of saturation for the second binding. In general, it is difficult to estimate KI2 when inhibitor solubility prevents enzyme saturation. The kinetics of MCA inactivation are complex and might require more data for complete parameterization. In summary, the combined results point toward an inactivation process that involves multiple inhibitor binding, inhibitor depletion, and partial inactivation.
Although clinical data are lacking for CA and MCA, an effort was made to predict an HDI with parameters scaled from rats. The exposure range for CA is estimated to be between 8 and 275 mg (Friedman et al., 2000; Kirkham et al., 2009; Chan et al., 2016). The prediction used a virtual study design of 275 mg dosed every 24 hours for 10 days. The prediction suggests that extensive exposure to cinnamon—for example, through food or large doses of cinnamon powder as a complementary treatment—would lead to noteworthy interactions (AUC changes > 4-fold) with CYP2A6 substrates (e.g., nicotine, letrozole, and metronidazole).
In summary, the evidence from multiple approaches indicates that the time-dependent inhibition of CYP2A6 by MCA and CA is a complex process involving multiple factors and mechanisms. Both agents are dynamic within the CYP2A6 active site, reorienting between a few preferred orientations. The CYP2A6 active site appears to be sufficiently voluminous and flexible to accommodate two MCA or CA ligands: for MCA, this is especially consistent with results from ligand binding and numerical method analysis. Apoprotein modification, heme degradation, and partial inactivation are observed with both agents. Although MCA is more potent than CA, both are present in cinnamon, and questions remain as to how one agent may modulate the inhibition potency and metabolism of the other (e.g., heterotropic effects). Furthermore, factors that influence the relative partitioning between heme degradation, apoprotein modification, and metabolite formation are unknown. Based on this study, individuals taking CYP2A6 substrates, such as nicotine (e.g., self-administration through smoking, vaping, nicotine replacement therapy) and ingesting large quantities of cinnamon may be at risk of increased exposure to the CYP2A6 substrate through HDI resulting from CA and/or MCA.
Authorship Contributions
Participated in research design: Espiritu, Yadav, Pelletier, Chan, Natesan, Harrelson.
Conducted experiments: Espiritu, Chen, Yadav, Larkin, Pelletier, Harrelson.
Performed data analysis: Espiritu, Chen, Yadav, Larkin, Pelletier, Chan, GC, Natesan, Harrelson.
Wrote or contributed to the writing of the manuscript: Espiritu, Chen, Yadav, Larkin, Pelletier, Chan, GC, Natesan, Harrelson.
Footnotes
- Received April 7, 2020.
- Accepted July 15, 2020.
This work was supported by National Institutes of Health National Institute on Drug Abuse [Grant R15-DA042341] (to J.P.H. and J.M.C.) and National Institute of General Medical Sciences [Grant R15-GM131293] (to S.N.), the Pacific University School of Pharmacy Research Incentive grant to J.P.H., and Pacific Research Institute for Science and Mathematics grant to J.M.C.
↵This article has supplemental material available at dmd.aspetjournals.org.
Abbreviations
- AICc
- Akaike’s Information Criterion
- AUC
- area under the concentration-time curve
- CA
- trans-cinnamaldehyde CHARMM Chemistry at Harvard Molecular Mechanics
- CI confidence internal DDI
- drug-drug interaction
- EII
- multiple-inhibitor binding (double binding)
- fm,CYP2A6
- fraction of nicotine metabolized by CYP2A6
- fu,mic
- unbound fraction in microsomes
- fu,p
- unbound fraction in plasma
- HDI
- herb-drug interaction
- HLM
- human liver microsome
- IL
- lipid partitioning
- IVIVE
- in vitro–in vivo extrapolation
- LC-MS
- liquid chromatography–mass spectrometry
- LC-MS/MS
- liquid chromatography–tandem mass spectrometry
- M
- indicates inhibitor depletion
- MCA
- 2-methoxycinnamaldehyde
- MD
- molecular dynamics
- MIC
- metabolite intermediate complex
- MOE
- Molecular Operating Environment
- 8-MOP
- 8-methoxypsoralen m/z mass to charge ratio
- P450
- cytochrome P450 PDB Protein Data Bank
- PI
- partial inactivation
- PRA
- percent remaining activity
- rCYP2A6
- recombinant CYP2A6
- RMSD
- root-mean-square deviation
- TDI
- time-dependent inhibition
- Copyright © 2020 by The American Society for Pharmacology and Experimental Therapeutics