Skip to main content
Drug Metabolism and Disposition logoLink to Drug Metabolism and Disposition
. 2012 Dec;40(12):2324–2331. doi: 10.1124/dmd.112.047308

Preferred Binding Orientations of Phenacetin in CYP1A1 and CYP1A2 Are Associated with Isoform-Selective Metabolism

Qingbiao Huang 1, Rahul S Deshmukh 1, Spencer S Ericksen 1, Youbin Tu 1, Grazyna D Szklarz 1,
PMCID: PMC3500547  PMID: 22949628

Abstract

Human cytochromes P450 1A1 and 1A2 play important roles in drug metabolism and chemical carcinogenesis. Although these two enzymes share high sequence identity, they display different substrate specificities and inhibitor susceptibilities. In the present studies, we investigated the structural basis for these differences with phenacetin as a probe using a number of complementary approaches, such as enzyme kinetics, stoichiometric assays, NMR, and molecular modeling. Kinetic and stoichiometric analyses revealed that substrate specificity (kcat/Km) of CYP1A2 was approximately 18-fold greater than that of CYP1A1, as expected. Moreover, despite higher H2O2 production, the coupling efficiency of reducing equivalents to acetaminophen formation in CYP1A2 was tighter than that in CYP1A1. CYP1A1, in contrast to CYP1A2, displayed much higher uncoupling, producing more water. The subsequent NMR longitudinal (T1) relaxation studies with the substrate phenacetin and its product acetaminophen showed that both compounds displayed similar binding orientations within the active site of CYP1A1 and CYP1A2. However, the distance between the OCH2 protons of the ethoxy group (site of phenacetin O-deethylation) and the heme iron was 1.5 Å shorter in CYP1A2 than in CYP1A1. The NMR findings are thus consistent with our kinetic and stoichiometric results, providing a likely molecular basis for more efficient metabolism of phenacetin by CYP1A2.

Introduction

Cytochromes P450 (P450s) comprise a superfamily of heme-containing enzymes that have a ubiquitous presence across the biological kingdom. These enzymes are associated with the oxidative metabolism of a large number of structurally diverse chemicals, including drugs, carcinogens, environmental chemicals, and endogenous compounds such as steroids and fatty acids. Each P450 isoform usually has the ability to metabolize a variety of substrates, and different P450s often display overlapping substrate specificities and inhibitor susceptibilities.

The human P450 1A subfamily consists of two members, CYP1A1 and CYP1A2, with CYP1A2 being one of the major P450s in the human liver (∼13–15%) (Shimada et al., 1994; Guengerich, 2005). Apart from their role in the metabolism of drugs such as phenacetin, caffeine, and warfarin, these two enzymes are also involved in the metabolic activation of procarcinogenic chemicals and toxins (Androutsopoulos et al., 2009; Zhou et al., 2009). CYP1A1 and CYP1A2 share 72% amino acid sequence identity but exhibit different substrate specificities and inhibitor susceptibilities (Kawajiri and Fujii-Kuriyama, 1991; Guengerich, 2005). For example, CYP1A1 metabolizes benzo[a]pyrene and polycyclic aromatic hydrocarbons to their toxic derivatives (Kawajiri and Fujii-Kuriyama, 1991; Shou et al., 1996), whereas CYP1A2 preferentially oxidizes heterocyclic aromatic amines (Hammons et al., 1997; Turesky et al., 1998). The two enzymes are also different in their tissue distribution and expression. Constitutively expressed CYP1A2 is particularly abundant in hepatic tissue, whereas CYP1A1 is distributed throughout many tissues, including the lungs, and is highly inducible. However, both enzymes are regulated, in part, by the aryl hydrocarbon receptor system, which is known to be induced by a variety of chemicals (Kawajiri and Fujii-Kuriyama, 1991; Shimada et al., 2002).

Despite many differences, CYP1A enzymes share functional similarities as well. Therefore, many of the substrate and inhibitor probes do not clearly differentiate between the 1A1 and 1A2 isoforms (Tassaneeyakul et al., 1993). For example, caffeine has been shown to be a nonspecific in vitro probe for CYP1A2, because it is also metabolized by CYP1A1 (Tassaneeyakul et al., 1992). Likewise, the resorufin substrates, 7-ethoxyresorufin and 7-methoxyresorufin, undergo O-dealkylation by CYP1A1 and CYP1A2, although the efficiencies of the reactions are different (Nerurkar et al., 1993). In the case of phenacetin and acetaminophen, both compounds have been well documented as substrates for CYP1A1 and CYP1A2 and are known to be metabolized by these enzymes with different efficiencies. Phenacetin is primarily metabolized to acetaminophen via O-deethylation, although the formation of a minor metabolite acetol (∼5%) via acetyl hydroxylation has also been reported (Yun et al., 2000). Acetaminophen can be further metabolized by CYP1A1 and CYP1A2 to a potent cytotoxin, N-acetyl-p-benzoquinone imine (NAPQI), along with 3-hydroxy acetaminophen.

Of interest, in each of the cases above, the substrates are nonspecific, yet demonstrably selective with respect to the two isoforms in terms of reaction rates, efficiency, and Km. It is likely that the subtle differences in oxidative catalysis arise from isoform-specific substrate binding position and mobility in the active site.

Despite the fact that the crystal structure of CYP1A2 and the homology model of CYP1A1 have been available, there has been no direct physicochemical evidence to assess the binding position(s) of phenacetin within the active sites of CYP1A1 and CYP1A2. One experimental method used to indirectly determine substrate position with respect to the heme iron is NMR T1 relaxation measurement, which was first used to study the orientation of a substrate (xylidine) in a P450 active site by Novak et al. (1977). More recently, this method has been successfully applied toward several ligand-P450 complexes (Myers et al., 1994; Regal and Nelson, 2000; Hummel et al., 2004, 2005; Cameron et al., 2005; Mosher et al., 2008; Gay et al., 2010). On the basis of these findings, T1 relaxation experiments are also suitable to study phenacetin/acetaminophen-CYP1A complexes because these compounds have reasonable numbers of protons with few overlapping NMR signals, characteristics necessary to resolve proton distances from the heme iron.

Therefore, in the current study, we chose phenacetin and acetaminophen to evaluate and compare the substrate binding characteristics in CYP1A1 and CYP1A2. The distance restraints based on NMR T1 relaxation experiments were used to dock the substrates in the active sites of the enzymes. The relationship between substrate binding orientation and CYP1A1 and CYP1A2 substrate selectivity has been further investigated using molecular modeling, enzyme kinetics, and stoichiometry studies. The results help to better rationalize the molecular basis for differences in substrate specificity between these two closely related P450 isoforms.

Materials and Methods

Materials.

Phenacetin, acetaminophen, 2-hydroxy acetanilide, aniline, sodium dithionite, dilauroyl-l-3-phosphatidylcholine (DLPC), and PVP were purchased from Sigma-Aldrich (St. Louis, MO). Potassium phosphate and EDTA were obtained from Thermo Fisher Scientific (Waltham, MA). Emulgen and Chaps were purchased from the Chemical Division of the KAO Corporation (Tokyo, Japan) and EMD Biosciences (La Jolla, CA), respectively. Deuterium oxide (D2O, 99.9%) was from Cambridge Isotope Laboratory (Andover, MA). Centricon 10-kDa molecular mass cutoff filters were obtained from Millipore Corporation (Billerica, MA). Carbon monoxide was purchased from Mountain State Airgas (Morgantown, WV).

Expression and Purification of CYP1A1 and CYP1A2.

The enzymes were expressed and purified as described previously (Liu et al., 2003, 2004; Tu et al., 2008; Huang and Szklarz, 2010). In brief, His-tag-containing CYP1A1 and CYP1A2 were expressed in Escherichia coli DH5α cells, and the enzymes were purified by affinity chromatography with nickel-nitrilotriacetic acid agarose. During purification of CYP1A2, 5 mM caffeine was added to all of the buffers to stabilize the enzyme. Caffeine was removed completely from the enzyme preparation during the subsequent ultrafiltration stage (Huang and Szklarz, 2010). Rat P450 reductase was expressed and purified as described earlier (Liu et al., 2003, 2004). The purified enzyme fractions were stored at −80°C until further use.

Enzyme Kinetics Assay.

Phenacetin O-dealkylase activities of CYP1A1 and CYP1A2 in vitro were determined as described previously (Huang and Szklarz, 2010) by quantification of product acetaminophen using high-performance liquid chromatography. The incubation mixture contained 0.5 μM CYP1A enzyme, 1 μM P450 reductase, and 45 μM DLPC in 100 mM potassium phosphate buffer (pH 7.5). After a 5-min preincubation, reactions were initiated by the addition of NADPH (1 mM final concentration). Reactions were conducted for 30 min, quenched by addition of 5 μl of 60% HClO4, and cooled on ice for 10 min. Then 10 μl of 100 μM 2-hydroxy acetanilide was added as internal standard, and the mixture was spun in a centrifuge for 5 min. Acetaminophen in the supernatant was quantified by high-performance liquid chromatography using a C18 column (Alltech Associates, Deerfield, IL). The mobile phase consisted of methanol-0.1% acetic acid (20:80, v/v; flow rate, 1 ml/min), and the eluent was monitored at 254 nm. The product was quantified by comparison with acetaminophen standards. Kinetic parameters (kcat and Km) were calculated using nonlinear regression with GraphPad Prism software (GraphPad Software Inc., San Diego, CA).

Stoichiometry Studies.

The methods for NADPH oxidation, H2O2 production, and oxygen consumption were the same as those described previously (Huang and Szklarz, 2010). The reaction mixtures were the same as those used for the enzyme kinetics assay described above. The reaction was initiated by addition of NADPH and conducted at 37°C. NADPH oxidation was determined spectrophotometrically at 340 nm, H2O2 production was assayed using the xylenol orange iron(III) assay, and oxygen consumption was measured using a Mitocell (Strathkelvin Instruments, Ltd., Glasgow, UK). The ratios of product/NADPH, H2O2/O2, and H2O/products were calculated to evaluate the coupling efficiency and uncoupling rates at three branch points of CYP1A1 and CYP1A2 catalytic cycles (Huang and Szklarz, 2010).

Spin State Determination.

The percentages of high and low spin in CYP1A1 and CYP1A2 enzymes with phenacetin/acetaminophen were determined as described previously (Locuson et al., 2006). Spectral titrations with phenacetin or acetaminophen in CYP1A1 and CYP1A2 enzymes were performed by spectral scanning between 320 and 500 nm, with the concentrations of phenacetin or acetaminophen identical to those in NMR studies. Absorbance spectra were deconvoluted into three components: a low-spin component, a high-spin component, and the broad δ-band using multiple Gaussian curve fitting with the OriginPro v8 package (OriginLab Corporation, Northampton, MA). To estimate percentages of low-spin components, the three components (δ-bands, low spin, and high spin) were used to fit data for phenacetin or acetaminophen with CYP1A1 and CYP1A2 enzymes.

Enzyme Preparation and Incubation for NMR.

Deuterated potassium phosphate buffer (1 M, pH 7.5) was prepared by multiple rounds of resuspension in D2O and lyophilization. A solution of 1% (w/v) polyvinyl pyrrolidone in 100 mM deuterated phosphate buffer was prepared and applied to the surface of the 10-kDa molecular mass cutoff Centricon filter. The enzyme preparation, initially containing 100 mM phosphate buffer and 20% glycerol, was exchanged and concentrated against 100 mM deuterated phosphate buffer over the Centricon filter by centrifugation (1000g, 4°C) three to four times. The final enzyme preparation contained less than 1% glycerol and approximately 55 μM CYP1A1 or 20 μM CYP1A2.

For studies involving cytochrome P450 reductase, the enzyme-reductase complex in the ratio of 1:2 was incubated at room temperature in the presence of 45 μM DLPC. Before the NMR experiments, all solutions were treated with Chelex 100 to remove any metal ion contaminants. Because the contributions of dioxygen were supposed to remain constant during the experiments, the samples were not degassed.

NMR Spectroscopy.

NMR T1 relaxation studies were performed on a Varian Inova NMR spectrometer operating at 600 MHz, internally locked on the deuterium signal of the solvent, D2O, as described previously (Hummel et al., 2004, 2005). Signals were internally referenced to the HDO peak at 4.8 ppm. A standard inversion recovery sequence (d1-180°-d2-90°) was used, along with presaturation of the residual HDO signal. The 90° pulse width was calibrated on each sample. The preacquisition delay (d1) was set to 10 × T1 (40 s) of the longest relaxation time. The spectra associated with at least 10 d2 values were acquired. Line broadening and gaussian fitting were applied for precise calculations of T1 values for the protons. The value of the longitudinal relaxation time was obtained by a nonlinear least-squares fitting of the peak height as a function of the delay d2, using Varian software. T1 values were measured on the substrate solutions without addition of enzyme and after the addition of enzyme to these solutions. The relaxation times T1 (ferrous-CO) of the ligand protons were measured after in situ conversion of the enzyme to its diamagnetic ferrous carbonyl complex (Fe2+-CO) by bubbling carbon monoxide for 15 min, followed by an addition of a small amount of sodium dithionite (∼1 mg). The sample was then allowed to equilibrate 30 min before to T1 (ferrous-CO) measurement. The final T1 (ferrous-CO) measurement of the ligands in presence of the Fe2+-CO complex and the T1 of the enzyme-free ligand were almost identical, which suggests very little paramagnetic contribution due to impurities. The integrity of the enzyme was assayed by measuring the UV-visible spectra of samples maintained under similar conditions. No significant cytochrome P420 formation was observed over the course of the experiment.

To validate the fast-exchange conditions, the temperature dependence of T1P was studied. Experiments involving enzyme titration studies with CYP1A1 and CYP1A2 were performed with phenacetin or acetaminophen as a substrate. Concentrated enzyme was added to the substrate in small increments, and the T1 measurements were performed after each addition.

Proton-Heme Iron Distance Calculations.

A more precise method for the distance (r) calculation using spin-state data was adopted as described in detail earlier (Hummel et al., 2008). In brief, eq. 1 to calculate the distance can be written as

graphic file with name zdd01212-5778-m01.jpg

The distance estimated by NMR is represented by r. The tumbling coefficient τc represents the correlation time of the dipolar interactions of the protein in solution and can be calculated by measuring T1P at several magnetic field strengths (Mildvan and Gupta, 1978). An estimate for τc of P450 1A2 is 3.38·10−10 s−1, as reported previously (Regal and Nelson, 2000). T1P is the portion of T1 due to paramagnetic effects alone and is given by eq. 2:

graphic file with name zdd01212-5778-m02.jpg

assuming that all of the diamagnetic contribution is represented by 1/T1, Fe2+−CO (Regal and Nelson, 2000). The parameter αm, the fractional binding coefficient, is obtained from the equation αm = [P450]/(KS + [Substrate]) under conditions of fast exchange when only one substrate species is present in solution (Regal and Nelson, 2000). KS values determined from visible spectroscopy were used for the distance calculations rather than KD determined by NMR for phenacetin/acetaminophen because, as reported by Regal and Nelson (2000), KS and KD values were either equal or differed only slightly on the basis of Michaelis-Menten kinetics. The KS values for phenacetin binding to CYP1A1 and CYP1A2 were 57 and 17 μM, respectively (Huang and Szklarz, 2010). In the case of acetaminophen, KS values reported previously in the literature for CYP1A1 and CYP1A2 were 870 and 108 μM, respectively (Myers et al., 1994) and were used for the purpose of distance calculation for acetaminophen. The S(S+1) term was simplified by eq. 3:

graphic file with name zdd01212-5778-m03.jpg

where fHS and fLS refer to the fractions of the high-spin and low-spin iron, respectively (Hummel et al., 2008).

Molecular Modeling: General.

All molecular modeling was performed on a Silicon Graphics Octane workstation using Insight II software (Accelrys, San Diego, CA) for molecular dynamics simulations, ligand docking, ligand construction, and minimizations, and Modeler 8v2 was used for homology modeling. Phenacetin and acetaminophen were constructed and minimized using the Insight II/Builder module. Energy minimization and MD simulations were performed using the Insight II/Discover module with the consistent valence force field. The parameters for heme containing Fe+3 were as described previously (Paulsen and Ornstein, 1991, 1992).

Construction of the Homology Model of CYP1A1.

A homology model of CYP1A1 was made with Modeler 8v2 (Sali and Blundell, 1993) using the CYP1A2 crystal structure as the template. CYP1A2 (Protein Data Bank 2HI4) coordinates were obtained courtesy of Dr. Eric F. Johnson (The Scripps Research Institute, La Jolla, CA) (Sansen et al., 2007). The CYP1A1 model consisted of core residues 32 to 512, and the N-terminal transmembrane region was disregarded. Modeler's conjugate gradients optimization method was applied to build five models based on the template-based spatial restraints. The resulting model with the best objective function value was selected for further study. Heme cofactor coordinates for this model were obtained directly from CYP1A2 after structural alignment to the apo-CYP1A1 homology model.

MD Docking with Distance Restraints.

Phenacetin and acetaminophen were initially placed into the active sites of both CYP1A1 and CYP1A2 manually to avoid steric overlaps. Automated docking of ligands was then performed with the Insight II/Affinity module using default parameters, as described previously (Liu et al., 2004; Ericksen and Szklarz, 2005; Huang and Szklarz, 2010), except that the ferric form of the heme was used. After 20 positions (or poses) were obtained, the most energetically favorable complex was subjected to MD simulations and minimization with NMR-based distance restraints imposed. To reduce possible protein deformation resulting from restraint forces, the protein backbone was tethered to its initial coordinates by a harmonic restraint force. Substrate protons were guided to NMR-derived distances from the heme iron by a gradually strengthening harmonic restraint (k = 2–32 kcal · mol−1 · Å−1) over 5 ps of MD. The nonbond cutoff was set at 15 Å, and a screened Coulomb potential with a distance-dependent dielectric was used to simulate an aqueous solvent environment. After MD simulations, the final frame of the trajectory was minimized to convergence while restraints were maintained.

Results

Purification of CYP1A1 and CYP1A2.

The overall yield of purified enzyme was approximately 20 to 40%, similar to that reported previously (Liu et al., 2004; Tu et al., 2008; Huang and Szklarz, 2010). The purity of CYP1A1 and CYP1A2 verified by SDS-polyacrylamide gel electrophoresis and Western blots indicated that the proteins were at least 95% pure. The UV-visible spectrum of the purified enzymes was that of a typical P450, mainly purified in its low-spin form, whereas the spectrum of the Fe2+-CO complex exhibited the characteristic peak at 450 nm, with little or no P420 formation. The holoenzyme content of the enzymes was consistently in the range of 40 to 60%, as observed previously in our laboratory.

Kinetic Parameters of Purified CYP1A1 and CYP1A2 Enzymes.

Both CYP1A1 and CYP1A2 catalyze the O-deethylation of phenacetin to the product acetaminophen (Fig. 1), but with different efficiencies. The Michaelis-Menten plots for CYP1A enzymes are shown in Fig. 2. Kinetic parameters, kcat and Km, and catalytic efficiencies, kcat/Km, for phenacetin O-deethylation catalyzed by CYP1A1 and CYP1A2, are shown in Table 1. The kcat for phenacetin oxidation by CYP1A2 was ∼2.5-fold higher than that for CYP1A1, whereas the Km was ∼8-fold lower in CYP1A2 than in CYP1A1, probably reflecting better substrate affinity of CYP1A2. Overall, the catalytic efficiency for phenacetin metabolism by CYP1A2 was approximately 18-fold higher than that measured in CYP1A1.

Fig. 1.

Fig. 1.

Structures of phenacetin and acetaminophen with protons numbered as they are referenced in the text. The hydrogens in the OCH2 group of phenacetin are the sites of oxidation.

Fig. 2.

Fig. 2.

Kinetics of conversion of substrate phenacetin to product acetaminophen by CYP1A1 (A) and CYP1A2 (B).

TABLE 1.

Kinetic parameters for phenacetin O-deethylation by CYP1A1 and CYP1A2

Enzymes Phenacetin O-Deethylation
kcata Kma kcat/Kmb
min1 μM μM1 · min1
CYP1A1 0.84 ± 0.08 803 ± 137 0.001 ± 0.001
CYP1A2 2.20 ± 0.10 120 ± 14 0.018 ± 0.002
a

Data are means of triplicate determinations.

b

kcat/Km (catalytic efficiency) is used here to compare substrate specificities.

Coupling of CYP1A Reactions.

The stoichiometry of CYP1A reactions was studied in the presence of phenacetin to compare the coupling efficiencies of the enzymes. The rates of NADPH oxidation, O2 consumption, product (acetaminophen) formation, and H2O2 and excess water production during phenacetin oxidation by CYP1A1 and CYP1A2 are shown in Table 2. Oxygen consumption was similar for both enzymes, but the consumption rate of NADPH was higher in CYP1A2 than in CYP1A1. Likewise, the formation of product acetaminophen, as well as by-product H2O2, was greater in CYP1A2. Of note, CYP1A2 produced ∼50% less water than CYP1A1. Table 3 presents coupling efficiencies of CYP1A1 and CYP1A2, both overall and at specific P450 uncoupling branching points. The coupling efficiency of the catalytic cycle to product formation was ∼2-fold higher in CYP1A2 than in CYP1A1 on the basis of the ratio of product formation to NADPH oxidation. The uncoupling of the reaction toward H2O2 production, measured as the H2O2/O2 ratio, occurs in the initial stages of the catalytic cycle during the formation of the [Fe2+-O2] and [Fe2+-OOH] species. This type of uncoupling was also ∼2-fold higher in CYP1A2. The uncoupling toward H2O formation, calculated as the ratio of H2O/product, takes place in the later stage of the reaction, after the formation of the [Fe3+-O-substrate] complex. In contrast to the previous uncoupling indicators, unproductive water formation was ∼4-fold lower in CYP1A2 than in CYP1A1.

TABLE 2.

Rates determined for phenacetin metabolism by CYP1A1 and CYP1A2

Data are means of triplicate determinations.

Enzyme NADPH Oxidized O2 Consumed Product Formed H2O2 Produced Excess H2Oa Excess H2Ob
nmol · min1 · (nmol of P450)1
CYP1A1 33 ± 4 22.4 ± 3.0 0.84 ± 0.08 8 ± 0 24 ± 0 26 ± 0
CYP1A2 43 ± 4 22.7 ± 0.9 2.20 ± 0.10 14 ± 2 15 ± 2 15 ± 2
a

Excess water (H2O) was calculated from the equation: H2O = NADPH − H2O2 − product.

b

Excess water (H2O) was calculated from the equation: H2O = 2 (O2 − H2O2 − product).

TABLE 3.

Coupling efficiency of CYP1A1 and CYP1A2 at different branching points of the P450 cycle

Ratios were calculated from the parameters in Table 2.

Enzyme Product/NADPHa H2O2/O2b H2O/Productc
CYP1A1 0.025 0.37 28.71
CYP1A2 0.051 0.62 6.82
a

Efficiency of the coupling of reducing equivalents to product formation.

b

Reflects uncoupling at the first or second branch point.

c

Reflects uncoupling at the third branch point.

Determination of Spin State in CYP1A1 and CYP1A2.

The percentages of high- and low-spin CYP1A1 and CYP1A2 enzymes, with no substrate, in the presence of phenacetin and in the presence of acetaminophen were determined for NMR distance calculations. The UV/visible spectra (wavelength 320–500 nm) of the enzymes were deconvoluted into a low-spin component (∼416–420 nm), a high-spin component (∼390–405 nm), and δ-bands (∼360 nm). The percentages of low and high spin calculated from the Soret bands of CYP1A1 and CYP1A2 in the absence and presence of substrates are shown in Table 4 and Fig. 3. In the absence of substrate, CYP1A1 and CYP1A2 existed primarily in the low-spin state (95 and 88% low spin, respectively). The addition of phenacetin or acetaminophen caused only a slight increase (2–3%) in the proportion of the high-spin state of CYP1A1. In contrast, the occupancy of the high-spin state in CYP1A2 increased by ∼9 and 17% upon binding of acetaminophen and phenacetin, respectively. Thus, phenacetin induced a stronger shift toward the high spin than acetaminophen in the case of CYP1A2, whereas both phenacetin and acetaminophen had similar small effects with respect to the induction of the high-spin state in CYP1A1.

TABLE 4.

Percentages of low spin and high spin in CYP1A1 and CYP1A2 in the absence and the presence of phenacetin or acetaminophen at 27°C

Enzyme (+Substrate) Low Spin High Spin
%
CYP1A1 (no substrate) 95 5
    + Phenacetin 93 7
    + Acetaminophen 92 8
CYP1A2 (no substrate) 88 12
    + Phenacetin 71 29
    + Acetaminophen 79 21
Fig. 3.

Fig. 3.

The percentages of high-spin (385 nm) and low-spin (418 nm) components of CYP1A2 with the absence and presence of phenacetin at 27°C. Abs, absorbance.

NMR T1 Experiments.

The spectral binding constant (KS) of phenacetin for CYP1A1 and CYP1A2 was calculated to be 57.4 and 17.9 μM, respectively, as reported previously (Huang and Szklarz, 2010). The KS values of 870 and 108 μM for acetaminophen reported previously in the literature for CYP1A1 and CYP1A2, respectively, were used for the purpose of distance calculation for acetaminophen (Myers et al., 1994).

The validity of distances derived from T1 relaxation times of substrate protons is based on the assumption of substrate binding under fast-exchange conditions, meaning that substrate molecules in the active site must be rapidly exchanging with those in the bulk solution on the NMR time scale. To verify fast-exchange conditions, the temperature dependence of the T1 relaxation times of the substrate protons was measured. A positive slope in the double reciprocal plot of T1P versus temperature indicates that the fast-exchange condition is being met. The double reciprocal plot of T1P versus temperature for CYP1A1 with phenacetin shown in Fig. 4 demonstrates that the fast-exchange requirement has also been satisfied. Similar results were obtained for CYP1A1 in the presence of acetaminophen and CYP1A2 in the presence of either phenacetin or acetaminophen (data not shown).

Fig. 4.

Fig. 4.

Temperature dependence of T1P of phenacetin protons in the presence of CYP1A1. Positive slopes indicate that the substrate bound in the active site is in fast exchange with the surroundings. Similar results were obtained for CYP1A2 in the presence of phenacetin and acetaminophen. ●, phenacetin protons H2/6; ○, H3/5; ▾, CH2; ▿, COCH3; ■, CH3.

The mean distances between the substrate protons and the heme iron for phenacetin and acetaminophen with CYP1A1 and CYP1A2 obtained from T1 relaxation experiments are shown in Tables 5 and 6. In general, there was a substantial difference in T1 relaxation times between the samples with and without carbon monoxide. T1 values for the CO-reduced samples were larger for all protons of both phenacetin and acetaminophen. The distance of the OCH2 proton(s) of phenacetin bound to CYP1A1, which appears very close to the water signal on the NMR spectrum, could not be measured accurately because of inadequate water suppression during the experiment. However, with accurate measurements for the other protons of phenacetin, and the structural model, we were able to infer their positions. All the aromatic protons of phenacetin/acetaminophen were nearly equidistant from the heme iron. In the case of phenacetin, the mean distances between the heme and the protons in H2/6, H3/5, COCH3, and CH3 were slightly longer in CYP1A2 than in CYP1A1. However, the mean distance between the OCH2 protons of the ethoxy group (site of phenacetin oxidation) and the heme iron was ∼1.5 Å shorter in CYP1A2 (7.27 Å) than in the complex with CYP1A1 (8.84 Å). In the case of acetaminophen, the distances of protons in H2/6, H3/5, and COCH3 were almost identical between the isoforms. Of interest, in both enzymes, phenacetin was positioned farther from the heme than acetaminophen.

TABLE 5.

T1 relaxation rate-estimated distances of phenacetin protons from the heme iron of CYP1A1 and CYP1A2

SEs for measurements are shown in parentheses. Errors in the T1 values were those reported by the fitting routine. Errors in the reported distances (r) were determined by propagation of error from the T1 calculation. Generally, the error is <10%.

Protona CYP1A1b
CYP1A2c
T1, Fe3+ T1, Fe −CO2+ rd Re T1, Fe3+ T1, Fe −CO2+ rd Re
Å Å
2,6 2.73 (0.25) 3.53 (0.21) 6.10 (0.24) 7.97 2.35 (0.18) 3.0 (0.21) 7.07 (0.52) 8.25
3,5 2.993 (0.1) 2.37 (0.12) 6.13 (0.41) 8.03 1.85 (0.1) 2.13 (0.11) 7.37 (0.65) 8.32
OCH2 1.62 (0.12) 1.65 (0.19) 8.84 (9.02)f 10.21 1.58 (0.15) 1.8 (0.18) 7.27 (1.32) 8.14
COCH3 1.33 (0.07) 1.44 (0.04) 6.46 (0.83) 9.21 1.35 (0.07) 1.45 (0.06) 7.75 (1.23) 9.44
CH3 1.6 (0.06) 1.89 (0.06) 5.94 (0.32) 9.95 1.44 (0.07) 1.61 (0.06) 7.34 (0.68) 8.64
a

See Fig. 1 for the numbering scheme of the protons for phenacetin.

b

[CYP1A1] = 0.778 μM; [phenacetin] = 7.9 mM; KS (CYP1A1) = 57 μM.

c

[CYP1A2] = 0.833 μM; [phenacetin] = 7.4 mM; KS (CYP1A2) = 17 μM.

d

Values were calculated by r = [9.78 × 1016T1Pm S(S + 1)τc)]1/6.

e

Values were obtained by averaging the 20 lowest-energy conformations selected by the docking program (see MD Docking with Distance Restraints).

f

Distance could not be determined with accuracy.

TABLE 6.

T1 relaxation rate-estimated distances of acetaminophen protons from the heme iron of CYP1A1 and CYP1A2

SEs for measurements are shown in parentheses. Errors in the T1 values were those reported by the fitting routine. Errors in the reported distances (r) were determined by propagation of error from the T1 calculation. Generally, the error is <10%.

Protona CYP1A1b
CYP1A2c
T1, Fe3+ T1, Fe −CO2+ rd Re T1, Fe3+ T1, Fe2+−CO rd Re
Å Å
2,6 3.58 (0.13) 4.31 (0.20) 5.74 (0.27) 6.97 3.57 (0.17) 4.38 (0.20) 5.79 (0.32) 6.74
3,5 3.67 (0.16) 4.45 (0.22) 5.81 (0.34) 6.95 3.69 (0.22) 4.51 (0.21) 5.85 (0.38) 6.88
COCH3 1.53 (0.05) 1.60 (0.06) 6.34 (1.33) 7.53 1.54 (0.07) 1.58 (0.06) 7.06 (4.39) 7.84
a

See Fig. 1 for numbering scheme of the protons for acetaminophen.

b

[CYP1A1] = 1.006 μM; [acetaminophen] = 24.5 mM; KS (CYP1A1) = 870 μM.

c

[CYP1A2] = 0.570 μM; [acetaminophen] = 23.6 mM; KS (CYP1A2) = 108 μM.

d

Values were calculated by r = [9.78 × 1016T1Pm S(S + 1)τc)]1/6.

e

Values were obtained by averaging the 20 lowest-energy conformations selected by the docking program (see MD Docking with Distance Restraints).

Position of Phenacetin/Acetaminophen Relative to the Heme Based on NMR Data and Molecular Modeling.

Models of Michaelis complexes of acetaminophen and phenacetin bound to CYP 1A1 and CYP1A2 with poses in agreement with the measured proton-heme distances (Tables 5 and 6) are shown in Fig. 5. The heme iron-substrate proton distances obtained from modeling simulations are included in Tables 5 and 6 and are, in general, somewhat longer than those from NMR. Phenacetin and acetaminophen display similar binding orientations in both CYP1A1 and CYP1A2. However, the protons of the OCH2 group in phenacetin, the site of metabolism, are closer to the heme iron in CYP1A2 than in CYP1A1.

Fig. 5.

Fig. 5.

Models of phenacetin (yellow) and acetaminophen (green) bound in CYP1A1 (A) and CYP1A2 (B) were constructed on the basis of NMR-derived distance restraints. The most favorably ranked orientations for each docked substrate were refined by restrained MD simulations using our measured distances between protons and iron. The substrate positions appear to overlap on the distal side of the heme (red), adjacent to the I helix (cyan helix to the right).

Discussion

The objective of the present study was to explore the structural basis for the differences in substrate specificity between CYP1A1 and CYP1A2 using phenacetin as a probe. These two enzymes are closely related to each other, with 72% amino acid sequence identity. The differences in activities probably arise from structural differences between CYP1A1 and CYP1A2. Our previous studies indicated that five key residues in the active site that differ between the enzymes play a role in the determination of substrate specificity with alkoxyresorufins (Liu et al., 2003, 2004; Tu et al., 2008) and phenacetin (Huang and Szklarz, 2010). In CYP1A1, the unique active site residues are Ser122, Asn221, Gly225, Leu312, and Val382, which correspond to Thr124, Thr223, Val227, Asn312, and Leu382 of CYP1A2, respectively. These residues line the active-site cavity of the enzymes and may directly interact with substrates and thus influence binding orientations.

The relationship between enzyme specificity and binding orientation of the substrate phenacetin within the active site of CYP1A1 and CYP1A2 was investigated using a number of complementary approaches, such as enzyme kinetics, stoichiometry, NMR, and molecular modeling. Kinetic studies have shown that CYP1A2 displayed a higher turnover rate and substrate specificity (kcat/Km) with phenacetin than CYP1A1 (Table 1), as expected. Of interest, the Km value that we obtained here is 10-fold higher than the value reported previously, whereas kcat and KS values are similar (Yun et al., 2000), in contrast to our recent studies (Huang and Szklarz, 2010), which corroborate the current data. Moreover, as indicated by KS, the affinity for phenacetin is approximately three times higher with CYP1A2 than with CYP1A1, consistent with their respective Km values. Stoichiometry studies revealed that CYP1A2 displayed higher coupling efficiency than CYP1A1 because of the formation of less water, despite higher production of H2O2 (Tables 2 and 3). Thus, it appears that CYP1A2 is consuming reducing equivalents at a faster rate than CYP1A1 for this particular substrate.

The differences in coupling efficiencies and catalytic rates of P450 enzymes may be related to the position of the substrate in the active site with respect to the heme cofactor. On the basis of NMR data, the distance between protons of phenacetin at the oxidation site (OCH2) and heme iron is shorter in CYP1A2 than in CYP1A1 (Table 5), consistent with higher catalytic efficiency and coupling in CYP1A2. A similar correlation between coupling efficiency and the proximity of the substrate oxidation sites to heme was also reported for CYP2C9 (Hummel et al., 2004; Mosher et al., 2008).

The spin transition indicates the likelihood of displacement of the water ligand by substrate and therefore suggests the time-averaged, temperature-dependent proximity of substrate to the distal face of the heme (Jovanovic et al., 2005). Binding of either phenacetin or acetaminophen produced a slight high-spin shift, although the enzymes continued to predominantly occupy the low-spin state, with CYP1A2 showing a somewhat higher proportion of the high-spin enzyme (Table 4). This result suggests that these substrates are located close to heme and have a strong ability to replace the water in the reduced state of P450. Thus, this difference in spin state distribution between the enzymes is consistent with previous kinetic and stoichiometric data (Huang and Szklarz, 2010), as well as the current NMR results.

NMR studies with CYP1A isoforms have been successfully performed in the past and included binding of acetaminophen (Myers et al., 1994) and fluoromethylanilines (Koerts et al., 1995) to CYP1A1 and binding of caffeine to CYP1A2 (Regal and Nelson, 2000). In this study, we evaluated the binding interactions of substrates phenacetin and acetaminophen with CYP1A1 and CYP1A2. In general, an increase in T1 values for both substrates was observed after the reduction to the CO-complexed ferrous form of the enzyme (Tables 5 and 6) due to a higher diamagnetic contribution from the protein. This is in agreement with previous findings concerning T1 relaxation times for CYP1A substrates, including earlier data for acetaminophen (Myers et al., 1994), as well as for fluoromethylanilines (Koerts et al., 1995) and caffeine (Regal and Nelson, 2000).

On the basis of our results, the mean distances of the aromatic and aliphatic protons of phenacetin to the heme center were slightly longer in CYP1A2 than in CYP1A1, but they were almost identical for acetaminophen. The symmetry of the 2,6 and 3,5 aromatic protons of phenacetin and acetaminophen led to the study of 2-hydroxy acetanilide, an analog of acetaminophen, to more precisely determine the orientation of the aromatic ring. Because all aromatic protons of 2-hydroxy acetanilide are nonequivalent and distinct, a well resolved 1H NMR spectrum was obtained (data not shown). The measured distances for these aromatic protons were similar to each other, as well as to those of phenacetin and acetaminophen, thus suggesting that the ring orientations of the substrates examined are similar. In previous studies, similar observations for other planar CYP1A substrates, such as caffeine (Regal and Nelson, 2000) and fluoromethylanilines (Koerts et al., 1995), led to the proposal of plane-parallel binding orientations. However, this does not rule out other substrate binding orientations. In fact, modeling the positions of phenacetin and acetaminophen in the active site of the enzymes led to the ligands assuming an orientation that lies somewhere in between parallel and perpendicular to the heme (Fig. 4). This orientation is, in general, similar to that observed for α-napthoflavone bound in the active site of CYP1A2 crystal (Sansen et al., 2007) and benzo[a]pyrene in a CYP1A1 model (Ericksen and Szklarz, 2005). Another explanation is that there are multiple substrate binding modes, whereas the distances calculated from NMR T1 experiments represent only the average binding orientation.

Furthermore, the distances in NMR T1 relaxation studies have been measured on the nonoxygenated P450 enzymes. Thus, NMR studies may not provide any information about binding of the substrate to the more catalytically relevant oxo-ferryl form of the enzyme. Furthermore, given the dynamic nature of the P450 enzymes and the relatively slow catalytic cycle, it is very likely that substrate binding positions in the active site may rearrange during or between catalytic cycles of the enzyme. Thus, the distances obtained from NMR and the corresponding enzyme-substrate models (Fig. 5) reflect nonproductive binding geometry representative of the initial enzyme-substrate complexes, in which the distances between the heme iron and the oxidation sites are too long to support the initial hydrogen abstraction. In the case of phenacetin, the distance between the heme iron and the oxidation site (OCH2) is shorter for CYP1A2, which may increase the probability of achieving the productive pose and/or reduce the possibility of uncoupling, which is consistent with the higher specificity of this isoform toward phenacetin.

Potentially, the binding of P450 reductase could allosterically influence active site geometry, which may, in turn, influence substrate binding position and catalysis. To test such a phenomenon, substrate-iron distances in the presence and absence of P450 reductase were measured and compared. However, on the basis of NMR results, the orientation of phenacetin did not change significantly in either CYP1A1 or CYP1A2 (data not shown).

As discussed earlier, phenacetin and acetaminophen are observed to bind in similar nonproductive orientations in the active sites of both CYP1A enzymes, with phenacetin somewhat farther from the heme prosthetic group (Fig. 5). In CYP1A2, the oxidation site of phenacetin is closer to the heme iron atom and in a position that, after a slight repositioning toward heme, would facilitate the rate-limiting hydrogen abstraction in the first step of the catalytic cycle. The reaction product, acetaminophen is found in a similar orientation slightly closer to the heme, which suggests that there is little movement of the compound in the active site except for brief approaches to the heme from the nonproductive binding orientation. Moreover, the acetamido group remains relatively close to the heme, allowing for the subsequent oxidation of acetaminophen at the amide nitrogen, producing a toxic metabolite, NAPQI. This result is in agreement with previous findings by Myers el al. (1994), who proposed a similar reaction mechanism on the basis of NMR-derived iron-nitrogen distances for acetaminophen. Furthermore, an analogous mechanism may operate during the oxidation of acetaminophen to NAPQI by CYP3A4, with the position of the amide group with respect to the heme likely to influence the reaction, as indicated by NMR T1 relaxation studies (Cameron et al., 2007).

In conclusion, CYP1A2 displayed higher catalytic efficiency of phenacetin O-deethylation than CYP1A1, consistent with stoichiometry results. The time-averaged binding orientations of phenacetin and acetaminophen within the active site of the enzymes based on NMR data showed that the distance between the OCH2 protons of the ethoxy group (site of phenacetin O-deethylation) and the heme iron was 1.5 Å shorter in CYP1A2 than in CYP1A1. The NMR results are thus in agreement with kinetic and stoichiometry studies and provide direct evidence that may help to explain more efficient metabolism of phenacetin by CYP1A2 than CYP1A1.

Acknowledgments

We thank Dr. Peter Gannett for his advice concerning NMR methods and for critical review of this manuscript. Molecular modeling studies were performed at the Computational Chemistry and Molecular Modeling Laboratory, Department of Basic Pharmaceutical Sciences, School of Pharmacy, West Virginia University, Morgantown, WV.

This work was supported by the National Institutes of Health National Center for Research Resources [Grant RR16440] (to G.D.S.) and National Institutes of Health National Institute of General Medical Sciences [GM079724] (to G.D.S.).

Article, publication date, and citation information can be found at http://dmd.aspetjournals.org.

http://dx.doi.org/10.1124/dmd.112.047308.

ABBREVIATIONS:
P450
cytochrome P450
NAPQI
N-acetyl-p-benzoquinone imine
DLPC
dilauroyl-l-3-phosphatidylcholine
Chaps
3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonic acid
MD
molecular dynamics.

Authorship Contributions

Participated in research design: Huang, Deshmukh, Ericksen, and Szklarz.

Conducted experiments: Huang, Deshmukh, Ericksen, and Tu.

Performed data analysis: Huang, Deshmukh, and Szklarz.

Wrote or contributed to the writing of the manuscript: Huang, Deshmukh, Ericksen, Tu, and Szklarz.

References

  1. Androutsopoulos VP, Tsatsakis AM, Spandidos DA. (2009) Cytochrome P450 CYP1A1: wider roles in cancer progression and prevention. BMC Cancer 9:187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Cameron MD, Wen B, Allen KE, Roberts AG, Schuman JT, Campbell AP, Kunze KL, Nelson SD. (2005) Cooperative binding of midazolam with testosterone and alpha-naphthoflavone within the CYP3A4 active site: a NMR T1 paramagnetic relaxation study. Biochemistry 44:14143–14151 [DOI] [PubMed] [Google Scholar]
  3. Cameron MD, Wen B, Roberts AG, Atkins WM, Campbell AP, Nelson SD. (2007) Cooperative binding of acetaminophen and caffeine within the P450 3A4 active site. Chem Res Toxicol 20:1434–1441 [DOI] [PubMed] [Google Scholar]
  4. Ericksen SS, Szklarz GD. (2005) Regiospecificity of human cytochrome P450 1A1-mediated oxidations: the role of steric effects. J Biomol Struct Dyn 23:243–256 [DOI] [PubMed] [Google Scholar]
  5. Gay SC, Roberts AG, Maekawa K, Talakad JC, Hong WX, Zhang Q, Stout CD, Halpert JR. (2010) Structures of cytochrome P450 2B4 complexed with the antiplatelet drugs ticlopidine and clopidogrel. Biochemistry 49:8709–8720 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Guengerich FP. (2005) Human cytochrome P450 enzymes, in Cytochrome P450: Structure, Mechanism and Biochemistry (Ortiz de Montellano PR. ed) 3rd ed, pp 377–530, Kluwer Academic/Plenum Publishers, New York [Google Scholar]
  7. Hammons GJ, Milton D, Stepps K, Guengerich FP, Tukey RH, Kadlubar FF. (1997) Metabolism of carcinogenic heterocyclic and aromatic amines by recombinant human cytochrome P450 enzymes. Carcinogenesis 18:851–854 [DOI] [PubMed] [Google Scholar]
  8. Huang Q, Szklarz GD. (2010) Significant increase in phenacetin oxidation on L382V substitution in human cytochrome P450 1A2. Drug Metab Dispos 38:1039–1045 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Hummel MA, Gannett PM, Aguilar JS, Tracy TS. (2004) Effector-mediated alteration of substrate orientation in cytochrome P450 2C9. Biochemistry 43:7207–7214 [DOI] [PubMed] [Google Scholar]
  10. Hummel MA, Gannett PM, Aguilar J, Tracy TS. (2008) Substrate proton to heme distances in CYP2C9 allelic variants and alterations by the heterotropic activator, dapsone. Arch Biochem Biophys 475:175–183 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Hummel MA, Locuson CW, Gannett PM, Rock DA, Mosher CM, Rettie AE, Tracy TS. (2005) CYP2C9 genotype-dependent effects on in vitro drug-drug interactions: switching of benzbromarone effect from inhibition to activation in the CYP2C9.3 variant. Mol Pharmacol 68:644–651 [DOI] [PubMed] [Google Scholar]
  12. Jovanovic T, Farid R, Friesner RA, McDermott AE. (2005) Thermal equilibrium of high- and low-spin forms of cytochrome P450 BM-3: repositioning of the substrate? J Am Chem Soc 127:13548–13552 [DOI] [PubMed] [Google Scholar]
  13. Kawajiri K, Fujii-Kuriyama Y. (1991) P450 and human cancer. Jpn J Cancer Res 82:1325–1335 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Koerts J, Rietjens IM, Boersma MG, Vervoort J. (1995) 1H NMR T1 relaxation rate study on substrate orientation of fluoromethylanilines in the active sites of microsomal and purified cytochromes P450 1A1 and 2B1. FEBS Lett 368:279–284 [DOI] [PubMed] [Google Scholar]
  15. Liu J, Ericksen SS, Besspiata D, Fisher CW, Szklarz GD. (2003) Characterization of substrate binding to cytochrome P450 1A1 using molecular modeling and kinetic analyses: case of residue 382. Drug Metab Dispos 31:412–420 [DOI] [PubMed] [Google Scholar]
  16. Liu J, Ericksen SS, Sivaneri M, Besspiata D, Fisher CW, Szklarz GD. (2004) The effect of reciprocal active site mutations in human cytochromes P450 1A1 and 1A2 on alkoxyresorufin metabolism. Arch Biochem Biophys 424:33–43 [DOI] [PubMed] [Google Scholar]
  17. Locuson CW, Gannett PM, Tracy TS. (2006) Heteroactivator effects on the coupling and spin state equilibrium of CYP2C9. Arch Biochem Biophys 449:115–129 [DOI] [PubMed] [Google Scholar]
  18. Mildvan AS, Gupta RK. (1978) Nuclear relaxation measurements of the geometry of enzyme-bound substrates and analogs. Methods Enzymol 49:322–359 [DOI] [PubMed] [Google Scholar]
  19. Mosher CM, Hummel MA, Tracy TS, Rettie AE. (2008) Functional analysis of phenylalanine residues in the active site of cytochrome P450 2C9. Biochemistry 47:11725–11734 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Myers TG, Thummel KE, Kalhorn TF, Nelson SD. (1994) Preferred orientations in the binding of 4′-hydroxyacetanilide (acetaminophen) to cytochrome P450 1A1 and 2B1 isoforms as determined by 13C- and 15N-NMR relaxation studies. J Med Chem 37:860–867 [DOI] [PubMed] [Google Scholar]
  21. Nerurkar PV, Park SS, Thomas PE, Nims RW, Lubet RA. (1993) Methoxyresorufin and benzyloxyresorufin: substrates preferentially metabolized by cytochromes P4501A2 and 2B, respectively, in the rat and mouse. Biochem Pharmacol 46:933–943 [DOI] [PubMed] [Google Scholar]
  22. Novak RF, Kapetanovic IM, Mieyal JJ. (1977) Nuclear magnetic resonance studies of substrate-hemoprotein complexes in solution. I. The interaction of xylidine with myoglobin, hemoglobin, and cytochrome P450. Mol Pharmacol 13:15–30 [PubMed] [Google Scholar]
  23. Paulsen MD, Ornstein RL. (1991) A 175-psec molecular dynamics simulation of camphor-bound cytochrome P-450cam. Proteins 11:184–204 [DOI] [PubMed] [Google Scholar]
  24. Paulsen MD, Ornstein RL. (1992) Predicting the product specificity and coupling of cytochrome P450cam. J Comput Aided Mol Des 6:449–460 [DOI] [PubMed] [Google Scholar]
  25. Regal KA, Nelson SD. (2000) Orientation of caffeine within the active site of human cytochrome P450 1A2 based on NMR longitudinal (T1) relaxation measurements. Arch Biochem Biophys 384:47–58 [DOI] [PubMed] [Google Scholar]
  26. Sali A, Blundell TL. (1993) Comparative protein modelling by satisfaction of spatial restraints. J Mol Biol 234:779–815 [DOI] [PubMed] [Google Scholar]
  27. Sansen S, Yano JK, Reynald RL, Schoch GA, Griffin KJ, Stout CD, Johnson EF. (2007) Adaptations for the oxidation of polycyclic aromatic hydrocarbons exhibited by the structure of human P450 1A2. J Biol Chem 282:14348–14355 [DOI] [PubMed] [Google Scholar]
  28. Shimada T, Inoue K, Suzuki Y, Kawai T, Azuma E, Nakajima T, Shindo M, Kurose K, Sugie A, Yamagishi Y, et al. (2002) Arylhydrocarbon receptor-dependent induction of liver and lung cytochromes P450 1A1, 1A2, and 1B1 by polycyclic aromatic hydrocarbons and polychlorinated biphenyls in genetically engineered C57BL/6J mice. Carcinogenesis 23:1199–1207 [DOI] [PubMed] [Google Scholar]
  29. Shimada T, Yamazaki H, Mimura M, Inui Y, Guengerich FP. (1994) Interindividual variations in human liver cytochrome P-450 enzymes involved in the oxidation of drugs, carcinogens and toxic chemicals: studies with liver microsomes of 30 Japanese and 30 Caucasians. J Pharmacol Exp Ther 270:414–423 [PubMed] [Google Scholar]
  30. Shou M, Gonzalez FJ, Gelboin HV. (1996) Stereoselective epoxidation and hydration at the K-region of polycyclic aromatic hydrocarbons by cDNA-expressed cytochromes P450 1A1, 1A2, and epoxide hydrolase. Biochemistry 35:15807–15813 [DOI] [PubMed] [Google Scholar]
  31. Tassaneeyakul W, Birkett DJ, Veronese ME, McManus ME, Tukey RH, Quattrochi LC, Gelboin HV, Miners JO. (1993) Specificity of substrate and inhibitor probes for human cytochromes P450 1A1 and 1A2. J Pharmacol Exp Ther 265:401–407 [PubMed] [Google Scholar]
  32. Tassaneeyakul W, Mohamed Z, Birkett DJ, McManus ME, Veronese ME, Tukey RH, Quattrochi LC, Gonzalez FJ, Miners JO. (1992) Caffeine as a probe for human cytochromes P450: validation using cDNA-expression, immunoinhibition and microsomal kinetic and inhibitor techniques. Pharmacogenetics 2:173–183 [DOI] [PubMed] [Google Scholar]
  33. Tu Y, Deshmukh R, Sivaneri M, Szklarz GD. (2008) Application of molecular modeling for prediction of substrate specificity in cytochrome P450 1A2 mutants. Drug Metab Dispos 36:2371–2380 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Turesky RJ, Constable A, Richoz J, Varga N, Markovic J, Martin MV, Guengerich FP. (1998) Activation of heterocyclic aromatic amines by rat and human liver microsomes and by purified rat and human cytochrome P450 1A2. Chem Res Toxicol 11:925–936 [DOI] [PubMed] [Google Scholar]
  35. Yun CH, Miller GP, Guengerich FP. (2000) Rate-determining steps in phenacetin oxidations by human cytochrome P450 1A2 and selected mutants. Biochemistry 39:11319–11329 [DOI] [PubMed] [Google Scholar]
  36. Zhou SF, Chan E, Zhou ZW, Xue CC, Lai X, Duan W. (2009) Insights into the structure, function, and regulation of human cytochrome P450 1A2. Curr Drug Metab 10:713–729 [DOI] [PubMed] [Google Scholar]

Articles from Drug Metabolism and Disposition are provided here courtesy of American Society for Pharmacology and Experimental Therapeutics

RESOURCES