throbber
Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 30, 2021
`
`1521-009X/12/4003-625–634$25.00
`DRUG METABOLISM AND DISPOSITION
`Copyright © 2012 by The American Society for Pharmacology and Experimental Therapeutics
`DMD 40:625–634, 2012
`
`Vol. 40, No. 3
`42770/3752997
`
`Deuterium Isotope Effects on Drug Pharmacokinetics. I. System-
`Dependent Effects of Specific Deuteration with Aldehyde Oxidase
`Cleared Drugs
`
`Raman Sharma, Timothy J. Strelevitz, Hongying Gao, Alan J. Clark, Klaas Schildknegt,
`R. Scott Obach, Sharon L. Ripp, Douglas K. Spracklin, Larry M. Tremaine, and Alfin D. N. Vaz
`Departments of Pharmacokinetics Dynamics and Metabolism (R.S., T.J.S., H.G., A.J.C., R.S.O., S.L.R., D.K.S., L.M.T.,
`A.D.N.V.) and Pharmaceutical Sciences (K.S.), Pfizer Global Research and Development, Groton, Connecticut
`
`Received September 14, 2011; accepted December 15, 2011
`
`ABSTRACT:
`
`The pharmacokinetic properties of drugs may be altered by kinetic
`deuterium isotope effects. With specifically deuterated model sub-
`strates and drugs metabolized by aldehyde oxidase, we demon-
`strate how knowledge of the enzyme’s reaction mechanism, spe-
`cies differences in the role played by other enzymes in a drug’s
`metabolic clearance, and differences in systemic clearance mech-
`anisms are critically important for the pharmacokinetic application
`of deuterium isotope effects. Ex vivo methods to project the in vivo
`outcome using deuterated carbazeran and zoniporide with hepatic
`systems demonstrate the importance of establishing the extent to
`which other metabolic enzymes contribute to the metabolic clear-
`
`ance mechanism. Differences in pharmacokinetic outcomes in
`guinea pig and rat, with the same metabolic clearance mechanism,
`show how species differences in the systemic clearance mecha-
`nism can affect the in vivo outcome. Overall, to gain from the
`application of deuteration as a strategy to alter drug pharmacoki-
`netics, these studies demonstrate the importance of understand-
`ing the systemic clearance mechanism and knowing the identity of
`the metabolic enzymes involved, the extent to which they contrib-
`ute to metabolic clearance, and the extent to which metabolism
`contributes to the systemic clearance.
`
`Introduction
`Deuteration of drugs to enhance their pharmacokinetic, pharmaco-
`dynamic, or toxicological properties has gained momentum as judged
`by a search of the SciFinder database with the search term “deuterated
`drugs.” Of 179 registries retrieved, 151 are since 2005 with an
`exponential growth since 2006. These include deuterated versions of
`patented and off-patent drugs with claims of increased efficacy, de-
`creased toxicity, reduced interpatient variability, and decreased drug
`dose or dosing frequency. Belleau et al. (1961) were among the first
`to demonstrate the pharmacodynamic effect of deuteration with ␣␣-
`dideuterated p-tyramine. The effect was attributed to decreased me-
`tabolism of p-tyramine by monoamine oxidases. Several reports that
`have examined the effect of deuteration on the pharmacokinetic and
`pharmacodynamic properties of drugs reveal results that include little
`to no effect (Tanabe et al., 1970; Farmer et al., 1979; Taylor et al.,
`1983; Burm et al., 1988; Dunsaed et al., 1995); increased systemic
`exposure, a pharmacodynamic effect, and receptor selectivity (Dyck
`et al., 1988; Schneider et al., 2006, 2007); and decreased toxicity
`(Najjar et al., 1978). However, in these studies the mechanisms
`underlying the observed effects or lack thereof were not examined.
`With the use of formyl-deuterated N-methylformamide, the hepato-
`
`Article, publication date, and citation information can be found at
`http://dmd.aspetjournals.org.
`http://dx.doi.org/10.1124/dmd.111.042770.
`
`toxicity was shown to be due to oxidative metabolism at the formyl
`carbon (Threadgill et al., 1989). Pohl and Gillette (1984 –1985) out-
`lined the kinetic basis for use of deuterated compounds to determine
`toxic metabolic pathways, and, in addition, Nelson and Trager (2003)
`have reviewed distinctions between “intrinsic KDIEs” and “observed
`KDIEs” in enzyme reaction mechanisms with particular emphasis on
`cytochrome P450 reactions. Foster (1984) and Kushner et al. (1999)
`have also discussed the application of deuterated drugs to drug phar-
`macokinetics, pharmacodynamics, and toxicity.
`A KDIE on the intrinsic metabolic clearance (CLint or Vmax/Km) is
`fundamental to the application of a deuteration strategy to alter drug
`pharmacokinetics. Multiple factors mute the magnitude of this isotope
`effect. These include substantial contribution to the metabolic clear-
`ance by conjugating enzymes (UDP-glucuronosyltransferases, sulfo-
`transferases, and glutathione transferases) and heteroatom oxidizing
`enzymes (flavin monooxygenases), in which carbon-hydrogen bonds
`are not broken; aspects of enzyme reaction mechanisms such as
`“metabolic switching” due to deuterium substitution, particularly im-
`portant with cytochrome P450 cleared molecules (Miwa and Lu,
`1987; Nelson and Trager, 2003); rate-limiting product release from
`enzymes, which mask intrinsic KDIEs (Ling and Hanzlik, 1989; Hall
`and Hanzlik, 1990; Bell-Parikh and Guengerich, 1999); and other
`biological processes such as organ blood flow-limited clearance, renal
`and/or biliary clearance by passive or active transport
`involving
`uptake or efflux pumps and enterohepatic recycling. Consequently, a
`
`ABBREVIATIONS: KDIE, kinetic deuterium isotope effect; LC, liquid chromatography; MS, mass spectrometry; QC, quality control; MRM, multiple
`reaction monitoring; AO, aldehyde oxidase; AUC, area under the curve.
`
`625
`
`Auspex Exhibit 2004
`Apotex v. Auspex
`IPR2021-01507
`Page 1
`
`

`

`626
`
`SHARMA ET AL.
`
`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 30, 2021
`
`SCHEME 1. Sequence of synthetic steps used to deuterate compounds used in this
`study.
`
`5.79 mmol, 3 Eq), and KI (20 mg, 0.12 mmol, 0.06 Eq). The resulting
`suspension was heated at 110°C for 4 h and then cooled to room temperature.
`The reaction was diluted with 20 ml of H2O and extracted with 20 ml of
`EtOAc. The EtOAc layer was washed two times with 10 ml of brine, dried over
`anhydrous Na2SO4, and concentrated under reduced pressure to afford 680 mg
`(89% yield) of b3 as a white solid. A suspension of b3 (500 mg, 1.27 mmol)
`in 35 ml of EtOD and 2 ml of tetraethylammonium was purged with N2 gas.
`To the suspension was added 30 mg of 10% Pd/C, and after purging with D2
`gas the reaction was stirred under a balloon of D2 gas for 3 h at room
`temperature. The reaction was then purged with N2 gas and filtered, and the
`filtrate was concentrated under reduced pressure. The residue was dissolved in
`30 ml of EtOAc, washed with 30 ml of H2O followed by 10 ml of brine, and
`dried over Na2SO4. The solvent was evaporated under reduced pressure, and
`the resulting residue was purified by flash column chromatography (ISCO
`silica gel cartridge eluting with 100% CH2Cl2 to 90% CH2Cl2/10% MeOH) to
`afford 360 mg (79% yield) of 1-[2H]carbazeran as a tan solid. LC-MS: m/z 362
`(MH⫹); monodeuterium isotopic content ⬎99%.
`2-[2H]Zoniporide. 2-[2H]zoniporide (CAS Registry Number 241800-98-6)
`was synthesized as shown in Scheme 1c. A suspension of quinoline carboxylic
`acid (c1, 4.00 g, 14.32 mmol) in 70 ml of EtOH and 1 ml of H2SO4 was heated
`at reflux for 18 h. An additional 1.5 ml of H2SO4 was added to the reaction,
`and reflux was continued for 20 h. The resulting reaction solution was cooled
`to room temperature and concentrated to approximately one-third of its volume
`by rotary evaporation. The residual solution was diluted with 150 ml of Et2O
`and washed two times with 50 ml of saturated aqueous NaHCO3 followed by
`50 ml of brine. The organic layer was dried over Na2SO4 and concentrated
`under reduced pressure to afford 3.85 g (88% yield) of the ethyl ester as a
`brown oil. The ester (2.35 g, 7.65 mmol) and 3-chloroperoxybenzoic acid (2.32
`g; 13.44 mmol, 1.8 Eq) in 70 ml of CHCl3 were stirred at room temperature for
`16 h. The reaction solution was diluted with 30 ml of CHCl3 and washed with
`50 ml of saturated aqueous NaHCO3, 50 ml of saturated aqueous NaHSO3, and
`finally 50 ml of brine. The organic layer was dried over Na2SO4 and concen-
`trated under reduced pressure to afford 2.74 g of crude c3 as a light brown oil.
`A suspension of c3 (3.30 g, 10.21 mmol) in 70 ml of D2O (99.9% D) was
`treated with 2 ml of 50% NaOD in D2O, and the resulting solution was heated
`at 100°C for 3 h (Kawazoe and Onishi, 1967). The reaction was cooled to room
`temperature, and the pH of the solution was adjusted to 4.0 by dropwise
`addition of D2SO4. The resulting suspension was extracted three times with
`100 ml of CHCl3. The combined organic layers were dried over Na2SO4 and
`concentrated under reduced pressure. The crude product residue was purified
`by flash column chromatography (ISCO silica gel cartridge eluting with 97.4%
`CH2Cl2/2.5% MeOH/0.1% AcOH to 94.9% CH2Cl2/5% MeOH/0.1% AcOH)
`to afford 2.20 g (73% yield) of c4 as a light yellow solid. LC-MS: m/z 297
`(MH⫹); monodeuterium isotopic content ⬎98%.
`
`KDIE on the intrinsic metabolic clearance alone may not translate into
`an alteration of the overall pharmacokinetics of a drug.
`Drug design strategies have successfully decreased the impact of
`cytochrome P450 enzymes in metabolic clearance, partly by increased
`use of nitrogen heteroaromatics in drug substructures. As a conse-
`quence, aldehyde oxidase is increasingly observed as an alternate
`metabolic pathway for clearance because of its ability to oxidize
`nitrogen heteroaromatic ring systems. The broad differential tissue
`distribution of this enzyme results in an inability to correlate in vitro
`intrinsic clearance to in vivo clearance, and failure of allometric
`scaling of clearance due to interspecies differences in this enzyme
`have resulted in design strategies to avoid its role in metabolic
`clearance (Pryde et al., 2010). An alternate approach to decrease
`clearance when this enzyme is involved may be the use of KDIEs with
`specifically deuterated substrates.
`In this study, we focused on aldehyde oxidase as the metabolic
`clearance enzyme using specifically deuterated substrates to establish
`mechanistic consistency across species and two Pfizer drugs, carba-
`zeran and zoniporide, for which some preclinical and clinical infor-
`mation was available, to demonstrate the importance of knowing the
`enzyme(s) involved in the metabolic clearance, their reaction mech-
`anisms, the species differences in metabolic pathways, and the use of
`in vitro methods to assess the probability for in vivo success. We have
`determined 1) in vitro intra- and intermolecular KDIEs for several
`aldehyde oxidase substrates, 2) in vitro KDIE on the intrinsic clear-
`ances and metabolic profiles in hepatocytes and hepatic subcellular
`fractions, and 3) the in vivo KDIEs on pharmacokinetic parameters
`for carbazeran and zoniporide administered orally and intrave-
`nously. Carbazeran was a phosphodiesterase-2 inhibitor for the
`treatment of chronic heart failure, discontinued because of low oral
`bioavailability (⬍5%) and short half-life in humans (Kaye et al.,
`1984), and zoniporide was a Na⫹/H⫹ exchanger-1 inhibitor for
`treatment of perioperative myocardial ischemic injury after sur-
`gery, for which high clearance was observed in humans and rats
`(Dalvie et al., 2010).
`
`Materials and Methods
`Unless otherwise stated, all reagents used in chemical syntheses and bio-
`chemical and biological studies were of reagent grade and used as such without
`further purification.
`Deuterated Substrates. Scheme 1 shows the synthetic approaches used to
`deuterate the substrates used in this study. The palladium-catalyzed reductive
`deuteration of ␣-chloro-heterocycles and the base-catalyzed deuterium ex-
`change of the ␣-hydrogen in heterocycle-N-oxides are well established meth-
`ods for the specific introduction of deuterium into selected sites of nitrogen
`heterocycles (Kawazoe and Onishi, 1967; Rylander, 1985).
`2-[2H]Quinoxaline, 1-[2H]phthalazine, and 2-[2H]quinoline. These were
`synthesized from their corresponding chloro derivatives by palladium-cata-
`lyzed reduction (Scheme 1a) (Rylander, 1985) and described for 2-[2H]qui-
`noxaline. A solution of 2-chloroquinoxaline (0.30 g, 1.82 mmol) in 10 ml of
`EtOD and 0.3 ml of tetraethylammonium was purged with N2 gas, and 20 mg
`of 20% Pd(OH)2/C was added. The suspension was then purged with D2
`(⬎99% isotopic purity) gas followed by stirring at room temperature for 4 h
`under a balloon of D2 gas. Before filtering, the reaction was purged with N2
`gas, and the filtrate was concentrated under reduced pressure. The residue was
`dissolved in 30 ml of EtOAc, washed with 30 ml of H2O, and concentrated
`under reduced pressure. Flash column chromatography of the residue on an
`ISCO silica gel cartridge eluting with 20% EtOAc/80% hexanes afforded 23
`mg (10% yield) of quinoxaline-d as an off-white solid. LC-MS: m/z 132
`(MH⫹); monodeuterium isotopic content ⬎99%.
`1-[2H]Carbazeran. 1-[2H]carbazeran (CAS Registry Number 70724-25-3)
`was synthesized as shown in Scheme 1b. A solution of 1,4-dichloro-6,7-
`dimethoxyphthalazine b1 (500 mg, 1.93 mmol) in 5 ml of dimethylformamide
`was treated with piperidine b2 (432 mg, 1.93 mmol, 1.3 Eq), K2CO3 (800 mg,
`
`Auspex Exhibit 2004
`Apotex v. Auspex
`IPR2021-01507
`Page 2
`
`

`

`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 30, 2021
`
`DEUTERIUM ISOTOPE EFFECTS ON DRUG PHARMACOKINETICS
`
`627
`
`A solution of c4 (0.72 g, 2.42 mmol) in 18 ml of MeOD was purged with
`N2 gas and treated with 126 mg of 10% Pd/C followed by ammonium formate
`(0.72 g, 11.42 mmol, 4.7 Eq). The resulting suspension was heated at 45°C for
`1 h and then cooled to room temperature and filtered. The filtrate was
`concentrated, and the resulting residue was diluted with 50 ml of H2O and
`extracted into 50 ml of CHCl3 containing 0.5 ml of AcOH. The organic layer
`was dried over Na2SO4 and concentrated under reduced pressure to afford 570
`mg (84% yield) of c5 as an off-white solid.
`A solution of c5 (0.35 g, 1.25 mmol) in 8 ml of SOCl2 was heated at reflux
`for 1 h. The solution was then cooled to room temperature and concentrated to
`a yellow solid by rotary evaporation. The solid was treated with 5 ml of toluene
`and again evaporated under reduced pressure to a solid. The solid was then
`suspended in 9 ml of tetrahydrofuran, and this suspension was added to a
`solution of guanidine-HCl (0.74 g, 7.75 mmol, 6.2 Eq) in 14 ml of 1 M aqueous
`NaOH and 7 ml of tetrahydrofuran. The reaction was heated at 45°C for 1 h
`and then cooled to room temperature to afford a biphasic solution. The organic
`layer was partially concentrated, and the resulting liquid was extracted with 25
`ml of 5:1 CHCl3-isopropyl alcohol. The organic layer was dried over Na2SO4
`and dried under vacuum. The resulting crude product was slurried in 5 ml of
`ice-cold EtOAc and filtered to afford 150 mg (38% yield) of 2-[2H]zoniporide
`as an off-white solid. LC-MS: m/z 322.2 (MH⫹); monodeuterium isotopic
`content ⬎98%.
`Biological Reagents. Human and rat liver cytosols were purchased from
`BD Gentest (Woburn, MA). Guinea pig liver S-9 and cytosol were purchased
`from XenoTech, LLC (Lenexa, KS). Pooled and cryopreserved hepatocytes
`from either human or rat livers were obtained from Celsis In Vitro Technol-
`ogies (Baltimore, MD).
`Human aldehyde oxidase was partially purified from pooled liver cytosol by
`ammonium sulfate precipitation as follows. To a liter of stirred human liver
`cytosol preparation (obtained as a recovery fraction from the preparation of
`liver microsomes) maintained in an ice-water bath, buffered with 100 mM
`potassium phosphate buffer, pH 7.4, and constantly monitored for pH, was
`added solid ammonium sulfate (in small amounts at a time) to four weight to
`volume cuts of 5, 15, 20, and 25% ammonium sulfate. The pH was constantly
`adjusted with a 1 Msolution of potassium mono-hydrogen phosphate
`(K2HPO4) to maintain it between 7.0 and 7.4. After each weight to volume
`addition of ammonium sulfate, the suspension was stirred for 30 min to
`equilibrate, then centrifuged at 9000g for 20 min to pellet precipitated protein.
`The protein pellets were redissolved in 100 ml of 10 mM potassium phosphate
`buffer, pH 7.4, and assayed for aldehyde oxidase activity, as measured by the
`oxidation of phenanthridine to phenanthridone. The aldehyde oxidase activity
`was recovered in the 20 and 25% w/v ammonium sulfate protein pellets. The
`oxidase activity of this preparation of aldehyde oxidase was stable at ⫺40°C
`for more than 1 year.
`Bioanalytical Procedures. A 1 atomic mass unit difference between the
`proto and deutero forms of the substrates used in these studies required an
`accurate correction of the mass spectral contribution from the natural abun-
`dance from 13C in the proto forms of the compounds to the base mass of the
`deutero forms. The difference in the contribution determined empirically from
`standard curves to that calculated from molecular formulas is less than 1%.
`Sample Preparation. All standards, QC samples, and samples were pre-
`pared using the Hamilton MicroLab STAR (Reno, NV) robotic sample prep-
`aration station. A working solution of 4 ␮g/ml zoniporide or deuterated
`zoniporide was prepared separately by diluting 1 mg/ml stock solution in 1:1
`dimethyl sulfoxide-acetonitrile. Sequential dilution of the working solution in
`blank Sprague-Dawley plasma yielded standard solutions of 0.1, 0.2, 0.5, 1, 5,
`10, 50, 100, and 200 ng/ml and QC samples of 0.4, 8, and 80 ng/ml in plasma.
`A working solution of 10 ␮g/ml carbazeran or deuterated carbazeran was
`prepared separately by diluting 1 mg/ml stock solution in 1:1 dimethyl sul-
`foxide-acetonitrile. Sequential dilution of the working solution in blank guinea
`pig plasma yielded standard solutions of 0.1, 0.2, 0.5, 1, 2, 5, 10, 50, 100, 200,
`and 500 ng/ml and QC samples of 0.8, 8, 80, and 400 ng/ml in plasma. In the
`cassette-dosed studies of zoniporide the early time points between 0.16 and
`0.75 h were diluted 10- and 5-fold in blank Sprague-Dawley plasma. For the
`carbazeran cassette-dosed study, samples at time points between 10 and 30 min
`were diluted 10-fold in blank guinea pig plasma with the exception of the
`10-min time point in the intravenous study that was diluted 20-fold. To each
`50-␮l standard or sample solution from the zoniporide study was added 200 ␮l
`
`of acetonitrile containing 10 ng/ml internal standard for protein precipitation.
`The solutions were mixed and centrifuged at 3000g for 20 min. Then 120 ␮l
`of the supernatant from standard and sample mixtures was transferred to a
`96-deep well plate and diluted 1:1 with 0.1% formic acid in water. To 120 ␮l
`of standard and sample solutions in the carbazeran study was added 480 ␮l of
`0.1% formic acid in 1:1 acetonitrile-water. After mixing, the solutions were
`analyzed by LC-tandem mass spectrometry as follows. An API 4000 mass
`spectrometer (Applied Biosystems/MDS Sciex, Foster City, CA) equipped
`with Turbo V sources and a TurboIonSpray interface integrated with a Prom-
`inence LC-AD20 binary pump (Shimadzu, Columbia, MD) and an autosampler
`(PAL; CTC Analytics AG, Zwingen, Switzerland) with a cool stack temper-
`ature controlled at 4 – 8°C was used for analysis. All instruments were con-
`trolled and synchronized by Analyst software from Applied Biosystems/MDS
`Sciex.
`Ten-microliter aliquots of the zoniporide samples were injected onto a C-18
`reversed phase column (Luna C18-2, 5-␮m, 2.0 ⫻ 30 mm; Phenomenex,
`Torrance, CA) equilibrated with 5% solvent B (0.1% formic acid in acetoni-
`trile) in solvent A (0.1% formic acid in water) and maintained for 0.6 min after
`injection followed by a linear gradient to 98% solvent B over 0.65 min and
`held for 0.55 min and then returned to the original conditions in 0.4 min for a
`cycle time of 3.0 min. The flow rate for the zoniporide analysis was 0.5
`ml/min. The column was equilibrated at 5% solvent B for 0.8 min before
`reinjection. For the analysis of carbazeran, the flow rate was 0.6 ml/min. The
`gradient was maintained at 5% solvent B for 0.6 min, followed by a linear
`increase to 98% solvent B in 0.65 min, and kept at 98% solvent B for 0.95 min,
`followed by a linear decrease to 5% in 0.2 min. The column was equilibrated
`at 5% B for 0.6 min before reinjection.
`The ionization parameters for the compounds were optimized by direct
`infusion of undiluted standards in 50% aqueous acetonitrile containing 0.1%
`formic acid. The multiple reaction monitoring (MRM) transitions for carbaz-
`eran and 1-[2H]carbazeran were m/z 361.2 to m/z 272.2 and m/z 362.2 to m/z
`273.2, respectively. For zoniporide and 2-[2H]zoniporide, the MRM transitions
`used were m/z 321.2 to m/z 262.1 and m/z 322.1 to m/z 263.1, respectively. A
`proprietary Pfizer compound was used as internal standard, transitions for
`which were m/z 364.3 to m/z 228.2. The dwell time of each MRM transition is
`50 ms.
`Response Correction and Data Processing for Pharmacokinetic Param-
`eters. Data were processed using Applied Biosystems/MDS SCIEX Analyst
`software, Excel, and Watson LIMS (version 7.2; Thermo Fisher Scientific,
`Waltham, MA). Analyte peaks were integrated using Analyst 1.4.2 for quan-
`titation and then exported to Excel for response correction. Because the
`difference between the proto and deutero isotopomers is 1 atomic mass unit, a
`correction for the deutero compound was necessary because of the contribution
`from the 13C natural abundance in the proto compound. The response correc-
`tion factor was calculated from the proto standards in the linear range of the
`instrument response; the sample response of the deuterated compound was
`then corrected by subtracting the interference from the proto compound. The
`response correction factor of deuteron zoniporide was determined from
`the proto zoniporide standards in the linear range of the instrument response.
`The responses for deutero zoniporide were corrected by subtracting the con-
`tribution from proto zoniporide. The corrected responses were uploaded to
`Watson LIMS for linear regression and calculations for sample concentrations
`and pharmacokinetic parameters.
`In Vitro KDIEs. The intramolecular deuterium isotope effects for
`1-[2H]phthalazine and 2-[2H]quinoxaline were determined from the ratio of the
`m/z 148 and m/z 147 mass spectrometric responses in their 1-phthalazone and
`2-quinoxalone products. Intermolecular isotope effects on the rate constants for
`substrate depletion were determined in cytosol, hepatocytes, S-9, or partially
`purified human aldehyde oxidase using a 1:1 mixture of the deutero and proto
`forms of quinoline, carbazeran, and zoniporide at 1.0 ␮M.
`Michaelis-Menten kinetic parameters for quinoline and 2-[2H]quinoline
`were determined with guinea pig liver cytosolic aldehyde oxidase with eight
`substrate concentrations of quinoline (spanning ⫾5 ⫻ Km) after the linear
`dependence on time and protein concentration were established. 2-Quinolone
`was quantitated by UV absorbance at 250 nm for the peak matched with m/z
`146. Kinetic parameters were determined from v versus [S] plots using XL-Fit
`version 4.0.
`
`Auspex Exhibit 2004
`Apotex v. Auspex
`IPR2021-01507
`Page 3
`
`

`

`SHARMA ET AL.
`
`MH+ = 148 amu
`
`MH+ = 147 amu
`
`H
`
`NN
`
`N
`
`N
`
`NHNH
`
`O
`
`D
`H
`
`kH
`
`kD
`
`NN
`
`H
`
`D
`
`628
`
`a
`
`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 30, 2021
`
`substrate to internal standard versus time. Half-lives were calculated from the
`equation t1/2 ⫽ 0.693/first-order rate constant. Metabolites were identified by
`LC-MS from reactions used to determine half-lives in hepatocytes or S-9
`supplemented with cofactors or from reactions conducted at 10 ␮M concen-
`trations of the appropriate substrates.
`Kinetic Deuterium Isotope Effects on the Pharmacokinetics of Carba-
`zeran and Zoniporide in Guinea Pig and Rat. All procedures, including
`dosing methods, are within the guidelines approved by the Pfizer Institutional
`Animal Care and Use Committee. Male Hartley guinea pigs (325–350 g) and
`male Sprague-Dawley rats (250 –300 g) were used in all pharmacokinetic
`studies. Carbazeran and zoniporide were dosed as 1:1 mixtures of deutero and
`proto forms orally (in water) and intravenously as a bolus dose (in saline) via
`the jugular vein. Carbazeran was dosed at approximately 10 mg/kg b.wt. (5
`mg/kg each isotopic form) for both routes of administration, and zoniporide
`was administered orally at approximately 5 mg/kg b.wt. (2.5 mg/kg each
`isotopomer) and intravenously at approximately 2 mg/kg b.wt. (1 mg/kg each
`isotopomer) in saline. Blood samples (0.5 ml) were taken via the carotid artery
`at appropriate time intervals (between 10 min and 6 h). All samples were kept
`frozen at ⫺20°C until analysis. Pharmacokinetic parameters were determined
`only from the experimentally acquired data sets using Watson LIMS.
`
`Results
`Synthesis. NMR and mass spectrometric analysis of the deuterated
`compounds were consistent with their assigned specifically monodeu-
`terated structures as shown in Scheme 1.
`In Vitro Kinetic Deuterium Isotope Effects. To establish that
`interspecies differences observed for metabolism by aldehyde oxidase
`is not due to species-specific reaction mechanisms, the KDIE was
`determined for the metabolism of several substrates with liver cyto-
`solic aldehyde oxidase from human, rat, and guinea pig. The isotope
`effects determined were 1) intramolecular KDIE for 1-[2H]phthala-
`zine and 2-[2H]quinoxaline, 2) intermolecular KDIEs on the first-
`order rate constants for the oxidations of quinoline/2-[2H]quinoline,
`carbazeran/1-[2H]carbazeran, and zoniporide/2-[2H]zoniporide; and
`3) KDIE on the steady-state kinetic parameters for quinoline and
`2-[2H]quinoline with guinea pig liver cytosolic aldehyde oxidase. The
`aldehyde oxidase-susceptible carbon– hydrogen bonds adjacent to the
`aromatic nitrogens of phthalazine and quinoxaline are equivalent due
`to symmetry (Scheme 2). Replacement of either carbon– hydrogen
`bond by a carbon– deuterium bond provides a direct measure of the
`intrinsic KDIE from the ratio of the m/z 148 and m/z 147 ion current
`intensities in their respective lactam products (Nelson and Trager,
`2003). Table 1 shows the results for intra- and intermolecular KDIEs
`for oxidation of 1-[2H]phthalazine, 2-[2H]and quinoxaline, quinoline/
`2-[2H]quinoline,
`carbazeran/1-[2H]carbazeran,
`and zoniporide/2-
`[2H]zoniporide by aldehyde oxidase from human, rat, and guinea pig
`liver. Across species, the intramolecular KDIE was found to be
`between 4.7 and 5.1 for 1-[2H]phthalazine and 2-[2H]quinoxaline,
`
`kH/kD = m/z148 / m/z147
`
`O
`
`O
`MH+ = 148 amu
`D
`
`H
`MH+ = 147 amu
`O
`
`HN
`
`NN
`
`N
`
`NH
`
`kH
`
`kDD
`
`
`
`DH
`
`NN
`
`b
`
`kH/kD = m/z148 / m/z147
`SCHEME 2. Intramolecular deuterium isotope effect determined from the mass
`spectra of the 2-[2H]quinoxaline and 1-[2H]phthalazine metabolites formed by
`aldehyde oxidase.
`
`In general, reactions with hepatocytes and S-9 were conducted in a 5-ml
`volume (0.75 ⫻ 106 cells/ml for hepatocytes and 1 mg/ml protein for S-9) in
`Williams’ E medium (hepatocytes) or 50 mM potassium phosphate buffer, pH
`7.4 (liver S-9), whereas reactions with liver cytosol or partially purified human
`aldehyde oxidase were conducted in a 1.0-ml reaction volume in 50 mM
`potassium phosphate buffer, pH 7.4. Reactions were initiated by addition of
`substrate to the reaction mixtures that were preincubated at 37°C for approx-
`imately 5 min. Over a period of 60 min for cytosol and S-9 reactions and 90
`min for hepatocyte reactions, eight 100-␮l aliquots were removed and
`quenched in 100 ␮l of acetonitrile containing 1% formic acid. A 100-␮l aliquot
`of an internal standard solution (0.5 ␮M) was added; after mixing, the samples
`were filtered through a protein-binding filter membrane in a 96-well format. A
`25- to 50-␮l aliquot was analyzed by MRM of the proto and deutero substrates
`using substrate-specific transitions. Correction for the peak area of the 2H
`substrate was done by subtracting the appropriate percentage contribution due
`to the natural abundance 13C contribution from the 1H substrate. First-order
`rate constants were determined from the semilogarithmic plots of the ratio of
`
`TABLE 1
`KDIEs for the oxidation of quinoxaline, phthalazine, quinoline, carbazeran, and zoniporide by liver cytosolic aldehyde oxidase from human, rat, and guinea pig,
`hepatocytes from human and rat, and guinea pig S-9 supplemented with cofactors
`
`Substrate
`
`Human
`
`Hk/Dk
`
`Rat
`
`Guinea Pig
`
`Cytosol
`
`Hepatocytes
`
`Cytosol
`
`Hepatocytes
`
`Cytosol
`
`S-9 Supplemented
`
`2-[2H]Quinoxalinea
`1-[2H]Phthalazinea
`Quinolineb
`Carbazeranb
`Zoniporideb
`
`5.0
`5.1
`5.5
`4.8
`5.8
`
`N.D.c
`N.D.
`N.D.
`1.5
`1.9
`
`5.1
`5.0
`6.1
`5.0
`3.6
`
`N.D.
`N.D.
`N.D.
`4.6
`2.7
`
`4.7
`4.9
`6.0
`6.0
`4.8
`
`N.D.
`N.D.
`N.D.
`5.0
`1.5
`
`N.D., not determined.
`a Determined from the ratio of the peak area for m/z 148 (Hk, corrected for the M ⫹ 1 contribution from the m/z 147 ion) to the m/z 147 ion (Dk).
`b Determined from the ratio of the rate constants for disappearance of 2-[1H]quinoline (Hk) and 2-[2H]quinoline (Dk), 1-[1H]carbazeran (Hk) and 1-[2H]carbazeran (Dk), and 2-[1H]zoniporide (Hk)
`and 2-[2H]zoniporide (Dk).
`
`Auspex Exhibit 2004
`Apotex v. Auspex
`IPR2021-01507
`Page 4
`
`

`

`Downloaded from
`
`dmd.aspetjournals.org
`
` at ASPET Journals on April 30, 2021
`
`DEUTERIUM ISOTOPE EFFECTS ON DRUG PHARMACOKINETICS
`
`629
`
`respectively. The intermolecular KDIEs on the first-order rate con-
`stants for metabolism of quinoline/2-[2H]quinoline, carbazeran/1-
`[2H]carbazeran, and zoniporide/2-[2H]zoniporide are between 3.6 and
`6.1 across species. The KDIEs on the steady-state kinetic constants for
`quinoline and 2-[2H]quinoline with guinea pig liver cytosolic alde-
`hyde oxidase show that the KDIE is primarily on Vmax (5.2) with a
`small effect on Km (1.1) resulting in a KDIE of 6.0 on the intrinsic
`clearance (Table 2). These results are consistent with a common
`reaction mechanism for aldehyde oxidases from human, rat, and
`guinea pig liver, where C–H bond cleavage occurs in the rate-limiting
`step and the KDIE is expressed on the intrinsic clearance in all
`species.
`Metabolically active hepatocytes have the complete complement of
`drug-metabolizing enzymes and consequently serve as the closest in
`vitro surrogate for in vivo hepatic metabolism (Fabre et al., 1990). The
`extent to which aldehyde oxidase contributes to the overall hepatic
`metabolic transformation of a drug may then be established by exam-
`ining the KDIE on the intrinsic clearance of the drug in hepatocytes.
`Guinea pig hepatocytes are not commercially available; therefore, the
`guinea pig liver S-9 fraction supplemented with NADPH and UDP-
`glucuronic acid cofactors and alamethicin (to permeate the micro-
`somal membrane) (Fisher et al., 2000) was used to mimic the hepa-
`tocyte system as closely as possible (Dalvie et al., 2009). The KDIEs
`for carbazeran in rat hepatocytes and guinea pig S-9 are 4.6 and 5.0,
`respectively (Table 1). These values are comparable to the KDIEs
`with cytosolic aldehyde oxidase of the three species examined (Table
`1) and suggest that in the guinea pig and rat aldehyde oxidase is
`probably the primary route of drug metabolic clearance. In contrast, in
`human hepatocytes the KDIE is significantly decreased (1.5) (Table
`1), suggesting that in humans other metabolic pathways contribute a
`greater extent of carbazeran’s hepatic metabolic clearance. Consistent
`with this interpretation, the glucuronide of carbazeran was identified
`as the major metabolite in human hepatocyte reactions with the
`aldehyde oxidase product secondary (Fig. 1A). Although the carbaz-
`eran glucuronide metabolite was also detected in the guinea pig S-9
`and rat hepatocyte reactions (Fig. 1, B and C, respectively), the levels
`were negligible in comparison with that for the aldehyde oxidase
`metabolite.
`The KDIEs for zoniporide with cytosolic aldehyde oxidase from
`human and guinea pig are similar (5.8 and 4.8) (Table 1). However,
`with human hepatocytes and guinea pig S-9 supplemented with co-
`factors, the KDIE is substantially reduced (1.9 and 1.5, respectively)
`(Table 1). This result suggests that the aldehyde oxidase pathway is
`not a major metabolic clearance mechanism for zoniporide in the
`human and guinea pig liver. With rat cytosol the KDIE is somewhat
`smaller (3.5) (Table 1) than would be expected if aldehyde oxidase
`were the only enzyme responsible for metabolism and is further
`decreased in rat hepatocytes (2.7) (Table 1). The decreased KDIE in
`rat cytosol and hepatocytes is accounted for by other metabolic
`pathways that are evident from the metabolic profile shown in Fig. 1D
`for rat hepatocytes. Hydrolysis of the acyl guanidine function to the
`carboxylic acid (M1) contributes approximately 10%, and other me-
`tabolites (M2, M3, M5, M7, M8, and M10) derived from oxidations
`by cytochromes P450 co

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