throbber
Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 14, 2020
`
`0090-9556/02/3001-7–12$3.00
`DRUG METABOLISM AND DISPOSITION
`Copyright © 2002 by The American Society for Pharmacology and Experimental Therapeutics
`DMD 30:7–12, 2002
`
`Vol. 30, No. 1
`564/951813
`Printed in U.S.A.
`
`COMPUTATIONAL MODELS FOR CYTOCHROME P450: A PREDICTIVE ELECTRONIC
`MODEL FOR AROMATIC OXIDATION AND HYDROGEN ATOM ABSTRACTION
`
`JEFFREY P. JONES, MICHAEL MYSINGER, AND KENNETH RAY KORZEKWA
`
`Department of Chemistry, Washington State University, Pullman, Washington (J.P.J.); and Camitro Corporation,
`Menlo Park, California (M.M., K.R.K.)
`
`(Received August 20, 2001; accepted October 1, 2001)
`
`This paper is available online at http://dmd.aspetjournals.org
`
`ABSTRACT:
`
`Experimental observations suggest that electronic characteristics
`play a role in the rates of substrate oxidation for cytochrome P450
`enzymes. For example, the tendency for oxidation of a certain
`functional group generally follows the relative stability of the rad-
`icals that are formed (e.g., N-dealkylation > O-dealkylation > 2°
`carbon oxidation > 1° carbon oxidation). In addition, results show
`that useful correlations between the rates of product formation
`can be developed using electronic models. In this article, we at-
`tempt to determine whether a combined computational model for
`
`aromatic and aliphatic hydroxylation can be developed. Toward
`this goal, we used a combination of experimental data and
`semiempirical molecular orbital calculations to predicted activa-
`tion energies for aromatic and aliphatic hydroxylation. The result-
`ing model extends the predictive capacity of our previous aliphatic
`hydroxylation model to include the second most important group
`of oxidations, aromatic hydroxylation. The combined model can ac-
`count for about 83% of the variance in the data for the 20 compounds
`in the training set and has an error of about 0.7 kcal/mol.
`
`The P4501 enzymes are a superfamily of monooxygenases involved
`in the metabolism of both exogenous and endogenous compounds.
`Ironically, these enzymes play a central role in both the prevention
`and induction of chemical toxicities and carcinogenicity. Although
`most P450 oxidations of xenobiotics result in detoxification, occa-
`sionally a more toxic intermediate is formed. In fact, many ultimate
`toxins and carcinogens are formed by the bioactivation of less reactive
`compounds, and many bioactivation reactions are mediated by the
`P450 enzymes. Often bioactivation reactions are in competition with
`detoxification pathways for the same substrate. Since these enzymes
`play such a central role in both detoxification and bioactivation,
`predictive models for cytochrome P450 catalysis will be useful tools
`for evaluating of the potential risks of environmental exposures. One
`of the most pertinent but difficult problems in risk assessment is
`translating bench results and mechanistic information into a form that
`can be used. This article outlines steps toward the development of
`computational models from laboratory data into a form that can be
`used for risk assessments that accurately reflect experimental results.
`For example, these models now more completely describe all posi-
`tions of metabolism for nitriles and should be more complete in
`predicting the toxicity related to nitrile metabolism. These semiem-
`pirical computational models blend experimental data and computa-
`tional chemistry in such a way as to provide a consistent prediction of
`the bioactivation rates for a broad spectrum of compounds. In partic-
`ular, these models can be used to predict xenobiotic metabolism by the
`
`This work was supported by National Institute of Environmental Health Sci-
`ences Grant ES09122.
`1 Abbreviations used are: P450, cytochrome P450; AM1, Austin model-1; PNR,
`p-nitrosophenoxy radical.
`
`Address correspondence to: Jeffrey P. Jones, Department of Chemistry,
`Washington State University, Pullman, Washington, 99164. E-mail: jpj@wsu.edu
`
`P450 enzyme family, including the bioactivation of compounds to
`toxins and carcinogens.
`Models such as those presented here for P450-mediated reac-
`tions can also play a role in drug design. Tools that predict
`regioselectivity can be used to assess the pharmacokinetics of
`drugs before synthesis, saving time and money in the drug devel-
`opment process. These types of tools, either alone or in combina-
`tion with homology or pharmacophore models, can also provide for
`the design of better drugs, with higher compliance and fewer toxic
`side-effects. However, P450 enzymes are difficult to model by the
`standard methods used for most drug targets, which are based
`mainly on predicting binding affinities related to steric features,
`since accurate results depend upon the prediction of both the
`electronic and steric features of the enzymes. This is different from
`many other enzyme systems since P450 has the need to metabolize
`a vast array of xenobiotics, which makes it impractical to have one
`enzyme for each compound or even each class of compounds.
`Therefore, although most cellular functions tend to be very spe-
`cific, xenobiotic metabolism requires enzymes with diverse sub-
`strate specificities. The cytochromes P450 have apparently as-
`sumed much of this role. The required diversity is accomplished by
`families and subfamilies of enzymes with generally broad substrate
`specificities, a very reactive oxygenating species and a broad
`regioselectivity. Thus, for many reactions, the electronic features
`of the substrate are all that are required to predict regioselectivity
`(Grogan et al., 1992; Harris et al., 1992; de Groot et al., 1995,
`1999; Yin et al., 1995).
`We have reported a rapid electronic model for the prediction of
`regioselectivity in P450-mediated hydrogen atom abstraction mecha-
`nisms (Korzekwa et al., 1990a). The methods depend only on the
`calculation of the AM1 ground-state energies for the parent compound
`
`7
`
`Apotex Ex. 1016
`
`Apotex v. Auspex
`IPR2021-01507
`
`

`

`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 14, 2020
`
`8
`
`JONES ET AL.
`
`TABLE 1
`
`AM1 energies for addition of methoxy radical to aromatic compounds
`
`Compound
`
`Benzene
`Toluene (para)
`Toluene (meta)
`Toluene (ortho)
`Anisole (meta)
`Anisole (para)
`Anisole (ortho)
`Chlorobenzene (para)
`Chlorobenzene (meta)
`Aniline (para)
`Aniline (meta)
`Nitrobenzene (para)
`Nitrobezene (meta)
`Cyanobenzene (para)
`Cyanobenzene (meta)
`Cyanobenzene (ortho)
`o-Xylene (para)
`2-Methylanisole (4-position)
`Ethylbenzene (para)
`p-Xylene (ortho)
`Napthalene (ortho)
`Napthalene (para)
`Benzimidazole (ortho)
`Benzimidazole (meta)
`Benzimidazole (para)
`
`a
`Hgrd
`
`0.035
`0.023
`0.023
`0.023
`⫺0.025
`⫺0.025
`⫺0.025
`0.023
`0.023
`0.032
`0.032
`0.040
`0.040
`0.085
`0.085
`0.085
`0.012
`⫺0.036
`0.013
`0.011
`0.064
`0.064
`0.107
`0.107
`0.107
`
`b
`
`Htran
`
`0.036
`0.023
`0.023
`0.023
`⫺0.025
`⫺0.026
`⫺0.028
`0.023
`0.024
`0.030
`0.032
`0.042
`0.042
`0.086
`0.086
`0.087
`0.011
`⫺0.037
`0.013
`0.010
`0.058
`0.059
`0.104
`0.106
`0.106
`
`a The ground state energy of the parent compound in Hartrees/mol.
`b The transition state energy for methoxy addition in Hartrees/mol.
`c The heat of formation of the tetrahedral intermediate after methoxy addition in Hartrees/mol.
`d The enthalpy of activation in kcal/mol.
`e The enthalpy of reaction in kcal/mol.
`
`c
`
`Htetrs
`⫺0.012
`⫺0.025
`⫺0.024
`⫺0.024
`⫺0.071
`⫺0.075
`⫺0.075
`⫺0.025
`⫺0.023
`⫺0.019
`⫺0.014
`⫺0.006
`⫺0.005
`0.037
`0.039
`0.039
`⫺0.036
`⫺0.085
`⫺0.034
`⫺0.037
`0.006
`0.010
`0.053
`0.057
`0.059
`
`d
`Hact
`
`9.4
`9.0
`9.3
`9.1
`9.5
`8.5
`7.8
`9.0
`9.4
`7.9
`9.2
`10.2
`10.5
`9.5
`9.8
`10.3
`8.9
`8.8
`9.0
`8.9
`5.3
`5.8
`7.3
`8.3
`8.7
`
`e
`Hreac
`⫺20.3
`⫺21.2
`⫺20.3
`⫺20.5
`⫺19.8
`⫺21.8
`⫺21.7
`⫺21.4
`⫺20.4
`⫺22.9
`⫺20.0
`⫺20.1
`⫺19.1
`⫺20.9
`⫺19.8
`⫺19.6
`⫺21.1
`⫺21.4
`⫺21.0
`⫺20.6
`⫺27.6
`⫺25.0
`⫺24.6
`⫺22.1
`⫺20.8
`
`and the product radicals. The relative activation energy is predicted by
`the use of eq. 1.
`
`⌬Hact ⫽ 2.60 ⫹ 0.22共⌬H兲R ⫹ 2.38共IPrad)
`
`(1)
`
`This model has been successfully used to predict the rates of metab-
`olism of halogenated hydrocarbons (Harris et al., 1992; Yin et al.,
`1995) and the regioselectivity of nitrile metabolism (Grogan et al.,
`1992) both in vivo and in vitro. In this article, we report the devel-
`opment of a model for aromatic oxidation and the combination of this
`model with our previous model for hydrogen atom abstraction. The
`combined models allow for the prediction of the electronic component
`for the regioselectivity of most of the reactions catalyzed by the
`cytochrome P450 superfamily of enzymes. These models are general
`models for the electronic component of P450 enzymes, which we have
`shown to be rather constant across both the enzyme families and
`species (Jones et al., 1990; Karki et al., 1995; Yin et al., 1995; Higgins
`et al., 1998).
`
`Materials and Methods
`
`Data analysis and graphing were done with SigmaPlot version 5 (SPSS, Inc.
`Chicago, IL). The energy difference for regioselectivity was determined by
`taking the log of the ratio of metabolites and multiplying by 0.616, which gives
`the energetic difference at 310°K.
`Semiempirical calculations were performed with Gaussian 98 (Gaussian,
`Inc., Pittsburgh, PA.). The semiempirical activation energies for hydrogen
`atom abstraction for the reactions outlined in the text were estimated with our
`quantum mechanical model [the p-nitrosophenoxy radical (PNR) model] based
`on hydrogen abstraction reactions using PNR (Korzekwa et al., 1990b) and the
`semiempirical AM1 Hamiltonian (Dewar et al., 1985). The model uses the
`calculated (AM1) enthalpies of reaction and the ionization potentials of the
`resultant radicals to predict the AM1 activation enthalpies (Hact). All open shell
`systems were treated with an unrestricted Hartree-Fock Hamiltonian. Each of
`the reactants and the transition states for the aromatic hydroxylation reactions
`were minimized using the default procedures in Gaussian 98, and each tran-
`
`sition state was confirmed by frequency calculations and had only one negative
`eigenvalue. All transition states were found by following the most negative
`eigenvalue after the methoxy radical oxygen-aromatic carbon bond in the
`tetrahedral intermediate was lengthened to 1.95 Å.
`The aromatics compounds listed in Table 1 were selected to cover a range
`of electron-donating and electron-withdrawing groups based on Hammett ␴
`values. We emphasized single-ring compounds at the expense of multiple-ring
`compounds since all of the experimental data were for single-ring-containing
`compounds. It is possible that a different model will be required for multiple-
`ring-containing compounds, five-membered rings, and other types of aromatic
`compounds for which no experimental is available.
`
`Results
`Ideally, we would be able to use the same oxidant (PNR) to model
`P450 aromatic addition that we used for hydrogen atom abstraction.
`We have shown previously that aromatic oxidation is likely to occur
`by addition of a triplet-like oxygen to form an tetrahedral intermediate
`(Korzekwa et al., 1989). This tetrahedral intermediate can then rear-
`range to epoxides, ketones, and phenols. Although addition of the
`PNR to an aromatic carbon progresses smoothly through a transition
`state to a tetrahedral intermediate, the electronic state of the tetrahe-
`dral intermediate is not the ground state. Thus, the reactions from
`reactants to products and products to reactants occur on two different
`potential energy surfaces. In addition to complicating any transition
`state studies, the value of any Brønsted correlations will be in doubt
`since the transition states are not necessarily intermediate between
`products and reactants. Thus, the p-nitrosophenoxy radical alone is
`not likely to serve as a versatile model for other P450-mediated
`oxidations, and other small oxygen radicals were tested within the
`AM1 formalism.
`To circumvent the problems associated with PNR additions to
`aromatic systems, methoxy radical was used for aromatic addition
`reactions. Using the AM1 formalism,
`the methoxy radical adds
`smoothly to aromatic compounds and, in contrast to the p-nitrosophe-
`
`Apotex Ex. 1016
`
`

`

`COMPUTATIONAL MODELS FOR CYTOCHROME P450
`
`9
`
`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 14, 2020
`
`been shown to undergo rapid interchange between the two sites of
`interest by isotope effect experiments. The octane and hexane values
`are for C-1 versus C-2 hydroxylation, whereas ethylbenzene and
`diphenylpropane provide an estimate of the effect of a phenyl sub-
`stituent on C-1 versus C-2 hydroxylation and benzylic versus an
`internal secondary position. The remaining compounds give values for
`substituent effect on benzylic hydroxylation and aromatic O-dealky-
`lation. The computationally predicted energies are given in Table 2. A
`plot of the predicted versus the observed regioselectivity is given in
`Fig. 2. We performed a linear regression on the experimental versus
`theoretical data to correct for the differences in predicted versus
`theoretical regioselectivities, forcing the y-intercept through the ori-
`gin. The results are shown in eq. 3.
`
`⌬⌬G measured ⫽ 0.74 䡠 共⌬Hact(Habs1) ⫺ ⌬Hact(Habs2))
`
`(3)
`
`Next, we obtained from the literature different experimentally mea-
`sured regioselectivity reactions that involved hydrogen atom abstrac-
`tion and aromatic oxidation. This data are shown in Table 3. The
`methods used for determining the energies have been described pre-
`viously (Higgins et al., 2001). Ideally, the regioselectivity of reaction
`is measured in multiple enzymes known to show relatively nonspe-
`cific binding. The ratio of two metabolites then is assumed to be a
`measure of the difference in reactivity. In practice, only a very small
`number of reported regioselectivity values meet these criteria. The
`sparse dataset means that as more data becomes available the models
`are likely to change; however, the approach described here should still
`be applicable.
`The metabolism of toluene was determined to give around 70%
`benzyl alcohol and 30% aromatic hydroxylation in phenobarbital-
`induced rat microsomes (Hanzlik et al., 1984; Hanzlik and Ling,
`1990) and to give 95% benzyl alcohol and 5% aromatic hydroxylation
`in human microsomes (Tassaneeyakul et al., 1996). These results give
`differences in energy of activation for the aromatic to benzylic hy-
`droxylation of 0.7 and 1.8 kcal/mol for the two systems, respectively.
`Since the work done in rat microsomes gave a range of ortho, meta,
`and para products, we decided to take the mean of the upper and lower
`limit of the range to fit to the computational results. For ethylbenzene,
`the reported ratio of para hydroxylation to benzylic hydroxylation in
`purified CYP2B1 was around 0.0013 with an energy difference of 4.1
`kcal/mol, reflecting the faster reaction for the secondary benzylic
`position relative to the primary benzylic position of toluene (White et
`al., 1986). For o-xylene and p-xylene the ratios of benzylic to aromatic
`were measured to be 11.5 and 33, which correspond to energy differ-
`ences of 1.5 and 2.2 kcal/mol (Iyer et al., 1997). When a similar
`experiment was conducted by Lindsey-Smith for anisole metabolism,
`aromatic oxidation was found to be faster than O-dealkylation by
`about 3 times (Lindsay-Smith and Sleath, 1983), corresponding to a
`difference in energy of 0.67 kcal/mol. This is reduced to a ratio of
`around 2 for ortho hydroxylation of 2-methylanisole, when the aver-
`age value of single-expressed enzymes shown to give information
`about energetics is taken into account (Higgins et al., 2001). Of these
`experiments, only the work of Higgins et al. (2001) was conducted
`with the expressed purpose of determining the energy difference
`between two pathways, although ideally all of the data would be
`generated in the same enzyme preparation and with the goal of
`accurately determining the ratios of interest.
`With these data, the model for hydrogen atom abstraction and the
`new model for aromatic addition can be combined. Figure 3 is a plot
`of the predicted versus measured difference in energy for the oxida-
`tions in Table 3. Obviously, a linear relationship exists. The equation
`for the line is shown in eq. 4. In eq. 3, ⌬⌬Gmeasured is the measured
`
`FIG. 1.The linear correlation of aromatic activation energies and the heats of
`reaction for formation of a tetrahedral intermediate by methoxy radical.
`The R2 value is 0.92, and the 95% confidence limits are shown by the dashed
`lines.
`
`noxy radical, remains on a single potential energy surface. Methoxy
`radical was added to the aromatic compounds shown in Table 1, and
`the heats of reaction and transition state were characterized for each
`reaction. A good linear correlation was observed for the heats of
`reaction and the activation energies reported in Table 1. A plot of the
`data is shown in Fig. 1. Equation 2 is the equation that describes the
`linear correlation between activation energy and heats of reaction.
`This linear correlation has a R2 value of 0.92 and a standard error
`approximately 0.35 kcal/mol. In eq. 2, ⌬Hact is the enthalpy of
`activation, and ⌬Hreac is the enthalpy of reaction. Since this equation
`depends only on the enthalpy of reaction for predicting rate, even the
`activation energy of a large compound can be predicted extremely
`rapidly.
`
`⌬Hact ⫽ 21.91 ⫹ 0.61 䡠 ⌬Hreac
`
`(2)
`
`Thus, correlations between computational activation energies and
`computational heats of reaction can be obtained for both hydrogen
`abstraction reactions (eq. 1) and aromatic addition reactions (eq. 2)
`with separate regression models. However, the presence of Brønsted
`correlations for two different series of reactions does not necessarily
`imply that the activation energies can be directly compared. For
`example, the hydrogen atom abstractions are predicted by AM1 to
`have barriers around 20 kcal/mol, whereas aromatic oxidation has
`barriers around 10 kcal/mol. Since experimentally hydrogen atom
`abstraction is found to be favored over aromatic hydroxylation in
`many instances, these results are quantitatively erroneous. Therefore,
`a method is required to parameterize the two models to a common
`energy scale.
`The activation energies of our aliphatic model obtained from eq. 1
`has never been corrected to reflect experimental activation energies.
`Thus, we need to scale eq. 1 relative to experimentally measured
`values. Literature values for regioselectivity that appear to reflect
`intrinsic reactivity differences were found for octane (Jones et al.,
`1986, 1990), hexane (Morohashi et al., 1983), diphenylpropane
`(Hjelmeland et al., 1977a), substituted diphenylpropanes (Hjelmeland
`et al., 1977b), 2-methylanisole (Higgins et al., 2001), 4-methyanisole
`(Higgins et al., 2001), ␣-chloro-p-xylene (Higgins et al., 2001), and
`ethylbenzene (White et al., 1986). The values for octane, 2-methyl-
`anisole, 4-methyanisole, ␣-chloro-p-xylene, and ethylbenzene have
`
`Apotex Ex. 1016
`
`

`

`10
`
`JONES ET AL.
`
`TABLE 2
`
`Experimentally measured energy differences in activation energies for aliphatic P450 oxidation reactions
`
`Compound
`
`Reference
`
`Regioselectivity Positions
`
`Ethylbenzene
`␣-Chloro-p-xylene
`2-Methylanisole
`4-Methylanisole
`1,3-Diphenylpropane
`Hexane
`Octane
`1-Phenyl-3-(4-fluorophenyl)propane
`1-Phenyl-3-(4-methylphenyl)propane
`1-Phenyl-3-(4-trifluoromethylphenyl)propane
`
`White et al., 1996
`Higgins et al., 2001
`Higgins et al., 2001
`Higgins et al., 2001
`Hjelmeland et al., 1977
`Morohashi et al., 1983
`Jones et al., 1990
`Hjelmeland et al., 1977
`Hjelmeland et al., 1977
`Hjelmeland et al., 1977
`
`Primary/benzylic
`Benzylic/CI-methyl benzylic
`Benzylic/O-demethylation
`Benzylic/O-demethylation
`Benzylic/secondary
`1/2 Hexanol
`1/2 Octanol
`Benzylic/substitutedc
`Benzylic/substitutedc
`Benzylic/substitutedc
`
`Measured ⌬
`kcal/mola
`⫺4.39
`⫺0.74
`0.40
`0.72
`1.28
`⫺1.85
`⫺2.14
`0.51
`0.0
`1.81
`
`Predicted ⌬
`kcal/molb
`⫺4.58
`⫺1.31
`2.25
`2.21
`2.44
`⫺2.15
`⫺2.14
`⫺0.16
`⫺0.34
`0.82
`
`a The energy difference as determined by taking the ln of the measured ratios and multiplying by 0.616, as described under Materials and Methods.
`b The energy difference between the predicted activation energy for hydrogen atom abstractions based on Eq. 2.
`c Benzylic hydroxylation of the phenyl ring over benzylic hydroxylation of the substituted phenyl ring.
`
`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 14, 2020
`
`1998). This apparent contradiction can be resolved by an understand-
`ing of the kinetics of branched pathways, which was described by
`Jones et al. (1986). In brief, if product formation is not rate limiting,
`the reactivity of the substrate will not affect the rate of product
`formation. However, if an alternate product can be formed, differ-
`ences in reactivity will be observed. For example, if norcamphor is
`substituted with deuterium, less product is observed than if it has
`hydrogen, even though the rate-limiting step is not norcamphor hy-
`droxylation. Atkins and Sligar (1987) found that the isotope effect was
`observed because the build-up in the enzyme-substrate complex due
`to deuterium substitution was prevented by alternate product forma-
`tion. In this case, the alternate product was water. We have seen
`similar results for the C-1 hydroxylation of octane; however in this
`case, the alternate products were C-2 and C-3 oxidation products
`(Jones et al., 1986). Therefore, slowing the rate of oxidation of one
`position by deuterium substitution causes an increase in the rate of
`metabolism of another position or an increase in decoupling to water
`formation (Higgins et al., 1998). This can only occur if the rate of
`exchange of the different positions within the active site is as fast or
`faster than the rate of the oxidation step. This means that for a
`substrate that can rotate freely in the active site the regioselectivity
`will be primarily determined by the electronic reactivity of the various
`positions on the molecule (Higgins et al., 2001). For these molecules,
`product regioselectivities can be predicted by the electronic charac-
`teristics of the substrate. Conversely, regioselectivity can provide an
`important tool for testing computational models.
`We previously reported the development of an electronic model for
`hydrogen atom abstraction with can predict aliphatic hydroxylation,
`amine dealkylation, and O-dealkylation (Korzekwa et al., 1990a). To
`our knowledge, this article is the first time anyone has combined
`predictive models for P450-mediated aromatic and aliphatic oxidation
`so that a continuous prediction can be made of either aromatic,
`aliphatic, or a combination of these mechanisms (Fig. 4). However,
`qualitative models that approach these models have been described for
`CYP2D6 (de Groot et al., 1999). The model described here was
`generated by correcting the computational predictions with experi-
`mental results for microsomes and single expressed enzymes. Every
`effort was made to use only experimental data that would be expected
`to be free of specific enzyme orientation effects, as described by
`Higgins et al. (2001), and thus should be a general electronic model
`for cytochrome P450 enzymes. However, we did use a number of data
`points derived from microsomal preparations. It is possible that the
`two metabolites in these cases result from two distinct enzymes in the
`microsomal preparation, and when possible, we chose the phenobar-
`bital-induced microsomal preparation since the major isoform is
`CYP2B1, an enzyme that has shown a lack of binding preference and
`
`FIG. 2.A plot of predicted and measured energy differences for hydrogen atom
`abstraction from the values given in Table 2.
`The R2 value is 0.81, and the 95% confidence limits are shown by the dashed
`lines.
`
`difference in energies based on regioselectivity, ⌬Hact(arom) is the
`predicted activation energy for aromatic hydroxylation based from eq.
`2, and ⌬Hact(Habs) is the uncorrected AM1 activation energy for
`hydrogen atom abstraction. The fit eq. 3 gives an R2 value of 0.85,
`with a standard error approximately 0.65 kcal/mol.
`
`⌬⌬G measured ⫽ 1.22 䡠 共⌬Hact(Habs)兲 ⫺ 1.1 䡠 共⌬Hact(arom)兲 ⫺ 17.5
`
`(4)
`
`Finally, predicted regioselectivity for both eq. 3 and eq. 4 can be
`plotted together as shown in Fig. 4. The R2 value for the data is 0.83,
`with a standard error approximately 0.71 kcal/mol.
`
`Discussion
`Experimental observations suggest that electronic characteristics
`play a role in the rates of substrate oxidation for cytochrome P450
`enzymes. For example, the tendency for oxidation of a certain func-
`tional group generally follows the relative stability of the radicals that
`are formed (e.g., N-dealkylation ⬎ O-dealkylation ⬎ 2° carbon oxi-
`dation ⬎ 1° carbon oxidation). In addition, results show that useful
`correlations between the rates of product formation can be developed
`using electronic models (Grogan et al., 1992; Yin et al., 1995). These
`results are apparently contradicted by the results of kinetic experi-
`ments, which show the steps before substrate oxidation (electron
`reduction) to be rate limiting in the catalytic cycle (Higgins et al.,
`
`Apotex Ex. 1016
`
`

`

`COMPUTATIONAL MODELS FOR CYTOCHROME P450
`
`11
`
`Experimentally measured energy differences in activation energies for aromatic and aliphatic P450 oxidation reactions
`
`TABLE 3
`
`Compound
`
`Ethylbenzene
`Toluene
`p-Xylene
`o-Xylene
`Anisole
`Toluene
`Toluene
`2-Methylanisole
`Anisole
`Anisole
`
`Reference
`
`White et al., 1986
`Hanzlik et al., 1984
`Iyer et al., 1997
`Iyer et al., 1997
`Hanzlik et al., 1984
`Hanzlik et al., 1984
`Hanzlik et al., 1984
`Higgins et al., 2001
`Hanzlik et al., 1984
`Hanzlik et al., 1984
`
`Regioselectivity Positions
`
`Aromatic (para)/benzylic
`Aromatic (meta)/benzylic
`Aromatic (ortho)/benzylic
`Aromatic (para)/benzylic
`Aromatic (meta)/O-demethylation
`Aromatic (ortho)/benzylic
`Aromatic (para)/benzylic
`Aromatic (para)/O-demethylation
`Aromatic (ortho)/O-demethylation
`Aromatic (para)/O-demethylation
`
`Measureda
`⫺4.1
`⫺2.3
`⫺2.2
`⫺1.5
`⫺0.9
`⫺0.9
`⫺0.6
`0.17
`0.25
`1.02
`
`Predictedb
`⫺3.2
`⫺1.9
`⫺2.3
`⫺1.8
`⫺0.6
`⫺1.7
`⫺1.4
`0.5
`0.7
`0.8
`
`a The energy difference in kcal/mol as determined by taking the ln of the measured ratios and multiplying by 0.616, as described under Materials and Methods.
`b The energy difference between the predicted activation energy for aromatic addition based on Eq. 3 and the predicted activation energy for hydrogen atom abstractions based on Eq. 2.
`
`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 14, 2020
`
`FIG. 3.A plot of the energy predicted differences between aliphatic and
`aromatic positions versus measured difference in energy for the oxidations in
`Table 3.
`The R2 value is 0.85, and the 95% confidence limits are shown by the dashed
`lines.
`
`FIG. 4.A plot of the predicted versus actual values for the combined aromatic
`and hydrogen atom abstraction model.
`The R2 value is 0.83, and the 95% confidence limits are shown by the dashed
`lines.
`
`a dominance of electronic factors for the type of compounds use in
`this study (White et al., 1984; Higgins et al., 2001). However, to
`exclude this data would decrease an already small dataset. In general,
`we have attempted to construct our electronic models with data for
`compounds that are small and do not contain strong binding features
`(Higgins et al., 2001). This means that the best data will not have
`orienting effects associated with the enzyme tertiary structure. Ori-
`enting effects have been shown to be important for a number of P450
`enzymes and need to be added to the electronic models for accurate
`prediction of metabolism by these enzymes (Szklarz and Halpert,
`1997; de Groot et al., 1999; Lightfoot et al., 2000; Rao et al., 2000;
`Szklarz et al., 2000; Xue et al., 2001).
`Thus, although this model is general to the P450 superfamily, it is
`not going to work to any significant extent on enzymes such as those
`in the 4B family that have a large contribution from substrate-protein
`interactions determining regioselectivity (Fisher et al., 1998). One
`might expect the models to work better with enzymes and substrates
`that show little preference in orientation, such as CYP1A2 (Higgins et
`al., 2001), CYP2B1 (Higgins et al., 2001), CYP2B4 (White et al.,
`1984), CYP2E1 (Higgins et al., 2001), and CYP3A4. It should be
`noted that, although these enzymes have been shown to be free of
`excessive orientation effects for some substrates, it is possible that
`other substrates will show strong orientation effects in any one of
`these enzymes. When binding/orientation effects become large, mod-
`
`els that account for binding effects on regioselectivity can be com-
`bined with the electronic models to give a more complete description.
`It is important to note that, although for certain enzymes the electronic
`model can stand alone to predict regioselectivity, it is doubtful that
`binding models alone, without an electronic model, can predict regi-
`oselectivity. Even strongly orienting P450 enzymes can give multiple
`products, and the electronic nature of each position must be known.
`Other considerations that need to be made when using this model
`are the general steric accessibility of the positions of interest. For
`example, octane is metabolized to 1-octanol, 2-octanol, and 3-octanol
`in the ratio of 1:23:7 (Jones et al., 1990). The model has been shown
`to predict the correct regioselectivity for the C-1 and C-2 positions;
`however, the electronics of the C-2 and C-3 position are essentially
`the same so the 23:7 ratio would not be well predicted. In this case, the
`C-3 position appears to be less accessible since it is buried in the
`center of the molecule. Similarly, C-4 hydroxylation is not seen with
`octane to any appreciable extent, whereas the electronic features are
`not significantly different from those at the C-2 position. Another
`feature of this type is ortho hydroxylation. For many compounds, the
`group adjacent to the ortho position will hinder metabolism at the
`ortho position. These types of steric shielding are independent of the
`P450 enzyme and could be added to the models presented here.
`Finally, it should be noted that no attempt has been made to
`externally validate these models. Although this would obviously be
`preferred, we felt that the datasets were too small to remove data for
`
`Apotex Ex. 1016
`
`

`

`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 14, 2020
`
`12
`
`JONES ET AL.
`
`external validation. We are currently working on developing new
`datasets to validate these models. We did look at the internal consis-
`tency of the models by leave-one-out cross validation and found q2
`values of 0.79 for the combined model (data in Fig. 4). The cross-
`validated q2 for leaving out 20% of the data ranged from 0.76 to 0.81
`over 10 runs with different randomly chosen groups.
`In conclusion, computational models are presented that can predict
`the two major P450 pathways of metabolism, aliphatic and aromatic
`oxidation reactions. The models depend on experimental data to
`correct for the errors associated with the fast semiempirical methods
`used to predict each reaction mechanism. The computational models
`corrected by experimentally measured regioselectivities give the pre-
`dictive models described by eqs. 3 and 4. These models do a reason-
`able job at predicting the regioselectivity energetics and can be used
`in conjunction with models for the active site of the enzyme. Studies
`are underway in the application of these models to predicting human
`drug metabolism.
`
`References
`
`Atkins WM and Sligar SG (1987) Metabolic switching in cytochrome P-450cam: deuterium
`isotope effects on regiospecificity and the monooxygenase/oxidase ratio. J Am Chem Soc
`109:3754 –3760.
`de Groot MJ, Ackland MJ, Horne VA, Alex AA, and Jones BC (1999) A novel approach to
`predicting P450-mediated drug metabolism. CYP2D6 catalyzed N-dealkylation reactions and
`qualitative metabolite predictions using a combined protein and pharmacophore model for
`CYP2D6. J Med Chem 42:4062– 4070.
`de Groot MJ, Denkelder GMDO, Commandeur JNM, Vanlenthe JH, and Vermeulen NPE (1995)
`Metabolite predictions for parasubstituted anisoles based on Ab-Initio complete active space
`self-consistent-field calculations. Chem Res Toxicol 8:437– 443.
`Dewar MJS, Zoebisch EG, Healy EF, and Stewart JJP (1985) AM1: a new general purpose
`quantum mechanical molecular model. J Am Chem Soc 107:3902–3909.
`Fisher MB, Zheng YM, and Rettie AE (1998) Positional specificity of rabbit CYP4B1 for
`omega-hydroxylation 1 of short-medium chain fatty acids and hydrocarbons. Biochem Biophys
`Res Commun 248:352–355.
`Grogan J, DeVito SC, Pearlman RS, and Korzekwa KR (1992) Modeling cyanide release from
`nitriles: prediction of cytochrome P450-mediated acute nitrile toxicity. Chem Res Toxicol
`5:548 –552.
`Hanzlik RP, Hogberg K, and Judson CM (1984) Microsomal hydroxylation of specifically
`deuterated monosubstituted benzenes. Evidence for direct aromatic hydroxylation. Biochem-
`istry 23:3048 –3055.
`Hanzlik RP and Ling K-HJ (1990) Active site dynamics of toluene hydroxylation by cytochrome
`P450. J Org Chem 55:

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket