throbber
Pharmaceutical Development and Technology, 7(1), 1–32 (2002)
`
`REVIEW ARTICLE
`
`Stabilization of Pharmaceuticals to
`Oxidative Degradation
`
`Kenneth C. Waterman,* Roger C. Adami,
`Karen M. Alsante, Jinyang Hong, Margaret S. Landis,
`Franco Lombardo, and Christopher J. Roberts
`
`Pfizer Global Research and Development, Eastern Point Road,
`Groton, CT 06340
`
`ABSTRACT
`
`A guide for stabilization of pharmaceuticals to oxidation is presented. Literature is
`presented with an attempt to be a ready source for data and recommendations for
`formulators. Liquid and solid dosage forms are discussed with options including
`formulation changes, additives, and packaging documented. In particular, selection
`of and methods for use of antioxidants are discussed including recommended levels.
`
`INTRODUCTION
`
`Scope
`
`This review article sets the stage for a pharmaceutical
`formulator to deal with the problem of drug-product
`chemical instability. In particular, this article focuses on
`one of the more common modes of degradation in drug
`products, namely oxidation. Methods are suggested for
`establishing that oxidation is indeed the problem, and
`what the particular oxidative pathway is for degradation.
`Although each new drug presents unique challenges,
`guidance is provided for resolving this problem based on
`the best information currently available.
`This review is organized to provide information on
`recognizing and predicting drug oxidation in dosage
`
`forms. Relevant oxidation mechanisms and various
`remedies are discussed for the different dosage forms
`including traditional
`liquid and solid dosage forms
`containing small molecules, as well as formulations
`containing protein and DNA-based pharmaceuticals.
`This review also provides a decision tree for addressing
`oxidative degradation, along with a detailed table of
`antioxidant additives and their commercial precedence.
`Other approaches discussed include packaging, counter-
`ions and pH, and mitigation of impurities.
`
`Formal Oxidation (Recognition of Oxidation)
`
`Oxidation is defined broadly as the loss of electrons
`from a molecule (increase in oxidation number). For
`organic molecules, this can be restated as an increase in
`
`*Corresponding author. Fax: (860) 441-3972; E-mail: ken_waterman@groton.pfizer.com
`
`Copyright q 2002 by Marcel Dekker, Inc.
`
`www.dekker.com
`
`1
`
`Eton Ex. 1027
`1 of 32
`
`

`

`rm §1
`
`H+H H
`
`1
`
`2
`
`3
`
`4
`
`alkene
`I
`
`alkyne
`2
`
`Oxidation
`Increase in oxygen content
`Decrease in hydrogen content
`
`l
`
`RCH,OH
`
`[OJ
`
`[HJ
`
`RCH,
`
`t
`
`Reduction
`Decrease in oxygen content
`Increase in hydrogen content
`
`-
`
`1--C'-'-
`, ,,
`--y
`I alkane
`I O
`- i-CH20R
`
`ether (alcohol for R=H)
`
`1
`
`/C=CH2 - -C=CH
`"
`-i-~0
`- -i-(; - i-}-o,
`
`ketone
`(aldehyde for R=H)
`2
`
`ester
`(carboxylic acid
`for R=H)
`3
`
`R
`
`O
`carbonate ester
`(carbonic acid for R=H)
`4
`
`~
`
`1
`i ·c:
`~
`li
`M
`
`I @
`J
`MARCEL DEKKER, INC. n
`3
`
`270 Madison Avenue, New York, New York 10016
`
`•
`
`ORDER REPRINTS
`
`2
`
`Waterman et al.
`
`oxygen or decrease in hydrogen content (1). Alterna-
`tively, oxidation can be defined as a reaction that
`increases the content of more electronegative atoms in a
`molecule (2). With organic systems, these electronega-
`tive heteroatoms are generally oxygen or halogens.
`
`structures shown below:
`
`Listed below are the general structures and names of
`some common pharmaceutically relevant species in the
`order of increasing oxidation states with the oxidation
`numbers listed below each species:
`
`When a compound is oxidized, another compound
`must be reduced. Hydration and dehydration are not
`oxidation/reduction reactions, though they add oxygen,
`since the reaction is essentially an internal oxidation and
`reduction: one carbon atom is oxidized while another is
`reduced. The net change to the oxidation state of the
`molecule is therefore zero. In the example below,
`ethylene is hydrated to ethanol. While the carbon on the
`left gains a hydrogen atom and is therefore effectively
`reduced, the carbon on the right gains an oxygen atom
`and is effectively oxidized. The net change to the
`molecule is zero.
`
`H2C ¼ CH2
`
`H2Oÿ! CH3CH2OH
`
`If there is any doubt as to the oxidation status of a
`molecule, the following procedure helps to identify the
`oxidation state of a given compound (3):
`
`1.
`
`Imagine a water molecule being added onto all
`unsaturated bonds in that molecule. If there is a
`ring in the molecule, open the ring with a water
`molecule.
`2. Count the number of heteroatoms in the water
`addition product. This is the oxidation number of
`that molecule.
`3. By comparing the oxidation numbers for reactants
`and products, one can determine if the reaction
`represents an oxidation.
`
`The italicized numbers in the following scheme
`represent the corresponding oxidation numbers for the
`
`is
`it
`To study an oxidation reaction mechanism,
`important to understand the electron transfer between the
`species involved by tracking the fate of electrons (using
`Lewis structures and arrows). For single-electron
`transfers, single-headed arrows are used, while for two-
`electron processes, double-headed arrows are used. As an
`example of tracking an electron transfer, disproportiona-
`tion of two radicals is shown with arrows indicating the
`“flow” of electrons:
`
`Preliminary Screening for Oxidative Instability
`
`Solid Dosage Forms
`
`There is currently no best practice purposeful
`degradation protocol to study oxidative degradation in
`solid-state drug substances or drug products. As a result,
`solid-state reactivity of the drug substances towards
`oxidation in drug products is explored most readily using
`excipient compatibility screening under thermal/humid-
`ity challenge conditions. Although there has been some
`discussion on using ternary mixtures of drug and
`excipients (with statistical design) (4), currently, binary
`mixtures of drug compound and tableting excipients
`(binders, fillers, disintegrants, etc.) are used commonly in
`preliminary screening experiments. Samples are pre-
`pared using standard mortar and pestle geometric
`dilution techniques (i.e., over a wide range of drug-to-
`
`Eton Ex. 1027
`2 of 32
`
`

`

`rm
`~
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`ORDER REPRINTS
`
`Oxidation Stabilization
`
`3
`
`excipient ratios). The mechanical forces involved in using
`the mortar and pestle are hoped to mimic some of the
`forces involved in milling and tableting, which often lead
`to amorphous or disordered drug regions (see “Oxidation
`in Solid Dosage Forms”). These blends are stored under
`different temperature and humidity conditions (indicated
`typically 58C,
`by the relative humidity, RH),
`room
`temperature/ambient humidity, 408C/closed bottle,
`408C/75% RH open bottle, 508C/closed bottle and
`508C/20% RH open bottle. These studies provide an
`indication of the oxidative instability of a drug in a binary
`mixture where the pure drug substance may be quite
`stable. Such studies help to identify particular excipients
`that may need to be avoided. Unfortunately, these blend
`stability studies do not, in general, predict rates for the
`system after granulation, milling, and tableting, and can
`therefore only be used qualitatively.
`If oxidative instability is suspected, studies can be
`performed to determine if molecular oxygen is involved in
`the oxidation process. Stability challenges that involve
`filling the headspace of vials with nitrogen (negative
`control) and pure O2 (positive control) are often useful to
`determine if molecular oxygen is involved in the
`degradation. If the headspace environments have little or
`no effect on the oxidative reactivity of the drug substance
`in these stability screenings, molecular oxygen may be
`involved in the reaction but not
`in the rate-limiting
`oxidation step, or the level of oxygen may still be
`sufficiently high that the reaction readily occurs. More
`thorough removal of oxygen in the headspace can be
`accomplished by a freeze – pump – thaw cycle. Alterna-
`tively, oxidation may be related to highly reactive
`impurities (peroxides, superoxides, hypochlorites, formic
`acid) present in the excipients as manufacturing-related
`impurities (see “Detecting and Controlling Impurities”).
`Such drug oxidative instability will often show itself in the
`form of greater decomposition rates for more dilute drug
`mixtures. Pre-treatment of the excipients with heat and
`radical scavengers such as nitric oxide or benzoquinone
`may also be helpful in implicating these impurities as the
`source of oxidative instability. While these techniques
`may be helpful in determining the causes and mechanisms
`of a drug oxidation, they have not been adapted to use in
`the actual solid dosage form stabilization.
`
`Liquid Dosage Forms
`
`Liquid dosage form oxidative stability screening
`generally involves examining the drug stability under a
`
`number of conditions. These conditions will vary
`depending on the type of dosage form (oral, parenteral,
`etc.) and limitations due to solubility and decomposition
`pathways competitive with oxidation. Among the factors
`to evaluate for a new drug candidate are the following:
`
`1. Acidity. The pH can impact the oxidative stability
`of ionizable drugs (see “Acidity and pH Effects”).
`2. Concentration. Excipient
`impurities will have
`more relative impact on dilute solutions than on
`concentrated solutions.
`3. Temperature. Although Arrhenius behavior
`is
`seldom observed, an indication of oxidative
`stability is found by examining the reactivity over
`the range 5 – 708C.
`4. Oxygen in the headspace. Both oxygen enriched
`and inert atmosphere samples can provide an
`indication of the tendency for a drug to oxidize in a
`particular formulation.
`5. Photo-oxidation. The stability of the drug in the
`presence of light and oxygen can be important. In
`addition, the use of added sensitizers (such as rose
`bengal) in examining the photostability of the drug
`can help determine whether
`the mechanism
`involves singlet oxygen.
`6. Metals. Addition of metal ions to solutions can
`indicate whether there is a tendency for the drug to
`be catalytically oxidized (see “Catalysis”). Typi-
`cally, FeCl3 or CuCl2 are added at levels less than
`100 ppm.
`7. Packaging. For parenteral dosage forms, a range of
`stoppers should be examined. For oral dosage
`forms, both plastic and glass bottles should be
`evaluated (see “Packaging/Liquid Dosage Forms”).
`
`General Issues
`
`It is difficult to use the Arrhenius equation to describe
`oxidative instability, or accelerated screening methods in
`general to predict room temperature shelf-life. This can be
`due to the following factors:
`
`1.
`
`Instability may be related to the amount of a
`peroxide impurity present in the particular lot of
`the excipient
`(see “Detecting and Controlling
`Impurities”). This could be correlated with the
`manufacturing processes involved, age of
`the
`excipient, and the conditions (temperature, humid-
`ity, sunlight) under which the material was stored.
`Oxidation rates can be very rapid at early time
`
`Eton Ex. 1027
`3 of 32
`
`

`

`roil
`~
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`ORDER REPRINTS
`
`4
`
`Waterman et al.
`
`2.
`
`points when peroxide impurities are plentiful and
`plateau or fall off after the impurity is consumed
`by the degradation reaction.
`If the generation of radicals is rate limiting, the
`kinetics can show autocatalysis; i.e., the rate of
`drug degradation increases as the radical concen-
`tration increases.
`3. There can be more than one mechanism involved
`in the degradation with the differences in
`activation parameters similar enough that tem-
`perature changes can essentially lead to different
`mechanisms dominating.
`In the solid state, temperature-sensitive properties
`of the drug product such as percent amorphous
`content, degree of hydration of the components,
`and molecular mobility may affect the oxidative
`degradation.
`5. The permeability of oxygen through packaging is
`temperature dependent.
`6. The solubility of oxygen in excipients (solvents)
`is inversely temperature dependent.
`
`4.
`
`Detection and identification of oxidative degradants is
`aided by the use of mass spectrometry. Table 1 gives
`some characteristic mass peaks that may be associated
`with a drug substance that has undergone some type of
`oxidative reaction. Often,
`tandem liquid chromato-
`graphy-mass spectrometry/mass (LC-MS/MS) techniques
`can be used in identifying specific sites of oxidation
`within complex molecules.
`
`Predicting Oxidation
`
`Purposeful Degradation (5,6)
`
`Oxidative studies executed to force drug substance
`degradation are useful
`to predict primary oxidative
`
`degradants in drug products. Published stability guide-
`lines (7) suggest
`the use of a concentrated oxygen
`atmosphere to generate oxidative degradation products
`for chromatographic identification. Since degradants can
`arise from reaction of the drug product with molecular
`oxygen (see “Autoxidation—Chain Processes” and
`“Electron Transfer”) or with oxidizing agents present
`in the formulation (usually peroxides, see “Peroxides and
`Other Oxidizing Agents”), it is important to conduct
`purposeful degradation studies with oxygen as well as
`hydrogen peroxide. Hydrogen peroxide is often non-
`predictive of molecular oxygen reactions (8) because it
`does not involve the radical chain process common with
`the oxygen-based reactions (see “Autoxidation—Chain
`Processes”). Hydrogen peroxide stress testing is useful in
`drug-product studies where hydrogen peroxide itself is
`an expected impurity in an excipient (see “Formation and
`Presence of Oxidants in Excipients” and “Detecting and
`Controlling Impurities”).
`For an oxygen-atmosphere purposeful degradation
`study, a solvent must be chosen that solubilizes the
`drug sufficiently (1 – 10 mg/mL) and ideally mimics
`the proposed formulation. Although ideally one would
`use protic
`solvents
`to mimic
`common protic
`excipients,
`this is complicated by the tendency for
`alcohols to slow oxidation reactions by competing
`with the drug for initiator radicals (9). For this reason,
`the polar aprotic solvent acetonitrile is often used
`in place of alcohols in model studies. A co-solvent
`may be necessary to achieve a sufficiently high
`concentration.
`Radical
`initiators can be effective in accelerating
`autoxidation (see “Initiation”), thereby allowing for easier
`characterization (10 – 12). Although electron-transfer (see
`“Peroxides and Other Oxidizing Agents”) rather than free-
`
`Possible Products of Drug Oxidation Based on Mass Spectral Data
`
`Table 1
`
`Mass Spectral Analyses
`
`Possible Products
`
`Parent 2 2
`Parent þ 14
`Parent þ 15
`Parent þ 16
`Parent þ 32
`
`Oxidation of – CH(OH) – to – CO – , primary or secondary amine to imine
`Oxidation of – CH to – C(O) –
`Oxidation of secondary amine to N-oxide
`Hydroxylation or conversion of – C – H to – C – OH; epoxidation of a double
`bond; sulfide to sulfoxide conversion; tertiary amine to N-oxide
`Hydro- or endoperoxide formation; sulfide to sulfone conversion
`
`Eton Ex. 1027
`4 of 32
`
`

`

`roil
`~
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`ORDER REPRINTS
`
`Oxidation Stabilization
`
`5
`
`radical propagation may dominate the mechanism in the
`dosage form, in most cases the products (though not their
`relative distributions) are the same in either case. Because
`the addition of initiators allows for generation of 10 – 20%
`degradation product within 10 days (using 1 – 10 mol%
`radical initiator in the presence of pressurized oxygen), it is
`generally advantageous to use this approach along with
`appropriate controls.
`To perform an autoxidative degradation study, the
`drug substance is dissolved in an appropriate solvent and
`transferred to a reaction vessel pressurized at 50 – 300 psi
`O2 to increase the oxygen concentration in solution, and
`heated to form radicals from the initiator (see Table 2 for
`a list of typical initiators).
`In carrying out a purposeful oxidative degradation
`study, it is critical to run the appropriate controls in order
`to get a better mechanistic understanding of whether the
`degradation results from a thermal,
`free-radical, or
`nonfree-radical process. Controls for these experiments
`include the following:
`
`1.
`2.
`
`3.
`
`4.
`
`drug with oxygen without initiator;
`drug with initiator purged of oxygen (with
`nitrogen or argon);
`drug without initiator purged of oxygen (at the
`reaction temperature); and
`initiator at appropriate level without drug
`substance to determine if any observed high
`pressure liquid chromatography (HPLC) peaks
`result from oxidation products that are not drug
`substance related.
`
`Redox Potential
`
`As discussed in “Formal Oxidation (Recognition of
`Oxidation)”, oxidation involves the (formal) loss of
`
`electrons. A compound’s oxidation potential (or redox
`potential) gives its thermodynamic tendency to lose an
`electron. A classical way to determine the redox potential
`and provide some kinetics for the decomposition of the
`oxidized species, involves the use of cyclic voltammetry
`(CV). Sweep rates of several thousand volts per second
`are possible with adequate instrumentation, which offers
`the opportunity of clocking the lifetime of radicals
`generated by oxidation, even for very fast processes (13).
`Because of the preparation and analysis time involved,
`CV techniques are more appropriate for detailed
`mechanistic investigations rather than for fast prelimi-
`nary screening of drug candidates.
`A comparison based on the use of known standards
`(antioxidants, drugs, and general organic compounds)
`can be used to assess the relative ease with which a
`compound undergoes electron transfer. This allows a
`prediction of stability based on a compound’s redox
`potential and the relative stability of standards with
`similar potentials. A compound, which by comparison to
`known oxidatively labile compounds yields a low
`oxidative potential (more easily oxidized), is likely to
`be prone to oxidative degradation. In Table 3 are
`tabulated some redox potentials of reference compounds
`and rough potentials of some oxidatively labile groups
`(the lower the potential, the more readily the species will
`be oxidized). Molecules that are generally stable to
`electron-transfer oxidation could still be unstable to
`oxidation by hydrogen-atom abstraction (see “Propa-
`gation”). These data allow for general rules for predicting
`electron-transfer based oxidative stability:
`
`1.
`
`2.
`
`oxidation potential $ 1300 mV: stable to electron-
`transfer oxidation;
`oxidation potential between 850 and 1300 mV:
`depends on specific drug and formulation; and
`
`Table 2
`
`Free-Radical Initiators Useful in Purposeful Oxidative Degradation Studies (from Waco Pure Chemical Industries)
`
`Initiator
`
`Chemical Abstract Service
`(CAS) #
`
`Temperature for
`10 hr T1/2 (8C)
`
`Solubility
`(mg/mL)
`
`0
`0
`0
`0
`0
`0
`
`-Azobis(N,N0
`-dimethyleneisobutyramidine)dihydrochloride
`-Azobis(4-cyanopentanoic acid)
`-Azobis(2-amidinopropane)dihydrochloride
`-Azobis(2-methyl-N-hydroxymethyl) propionamide))
`-Azobisisobutyronitrile (AIBN)
`-Azobis(2,4-dimethylvaleronitrile)
`
`2,2
`4,4
`2,2
`2,2
`2,2
`2,2
`
`27776-21-2
`61630-29-3
`2997-92-4
`61551-69-7
`927-83-3
`52406-55-0
`
`44 (H2O)
`69 (H2O)
`56 (H2O)
`86 (H2O)
`65 (toluene)
`51 (toluene)
`
`35.2 (H2O)
`1 (H2O)
`23.2 (H2O)
`2.4 (H2O)
`7.5 (MeOH)
`22 (MeOH)
`
`Eton Ex. 1027
`5 of 32
`
`

`

`roil
`~
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`ORDER REPRINTS
`
`6
`
`Waterman et al.
`
`Table 3
`
`Selected Potentials for Antioxidants and Drugs
`
`Compound
`
`Oxidizable Functional Group
`
`Oxidative Peak Potential
`
`Observed Stability Towards Oxidation
`
`Dextromethorphan
`Amitriptyline
`Fluoxetine
`Flurazepam
`Captopril
`Chlorpromazine
`Morphine
`L-Ascorbic acid
`Vitamin E
`Phenylbutazone
`Tetracyline(s)
`
`Tertiary amine
`Tertiary amine
`Secondary amine, benzylic ether
`Tertiary amine, imine
`Thiol
`Thioether
`Allylic alcohol, phenol
`Allylic alcohol
`Phenol
`3,5-Dioxopirazolidine
`Phenol, enols, tertiary amine
`
`1400 (14)
`1300 (15)
`1300 (15)
`900 (15)
`, 900 (19)
`840 (15)
`820 (21)
`700 (21)
`700 (21)
`660 (22)
`550 (23)
`
`Stable (9)
`Stable (16)
`Stable (17)
`Stable (18)
`Unstable (20)
`Unstable (20)
`Unstable (20)
`Unstable (20)
`Unstable (20)
`Unstable (9)
`Unstable (20,24)
`

`The potentials were obtained under different conditions, and are intended only as illustrative examples. All potentials are in mV vs. Ag/Ag
`electrode.
`
`reference
`
`3.
`
`oxidation potential # 850 mV: easily oxidized.
`
`Identifying Oxidative Reactivity Based on Structure
`
`Autoxidation
`
`Assuming autoxidation of a drug molecule occurs by
`an initial abstraction of a labile hydrogen atom followed
`by reaction with molecular oxygen (see “Autoxidation—
`Chain Processes”), oxidative degradation should be
`linked to the lability of hydrogen atoms within the
`molecular framework. The lability of hydrogen atoms
`within a molecular structure is described by the
`corresponding bond-dissociation energies;
`i.e.,
`the
`lower the bond energy, the more likely a hydrogen
`atom will be abstracted by a radical. Table 4 shows the
`representative bond-dissociation energies of some
`common functional moieties.
`A qualitative way to estimate hydrogen atom lability
`is to evaluate the stability of the radical that is formed.
`In general, the more stable the resulting radical, the
`easier for the hydrogen atom to be abstracted from the
`corresponding position on the molecule. The rules of
`radical stability are similar to the rules that govern
`reactive center stability (i.e., cations and anions). In
`general, for sp3 hybridized centers,
`tertiary centered
`radicals are more stable than secondary centered
`radicals which are more stable than primary radicals
`so that the corresponding hydrogen atom abstraction
`will occur most readily from tertiary hydrogens, and
`
`least readily from primary hydrogens. This stability
`order is supported by the bond-dissociation energies in
`Table 4, where the C – H bond energy for ethane is less
`than the C – H bond energy for methane. With sp2 and
`
`Table 4
`
`Representative Hydrogen Bond-Dissociation
`Energies for Some Common Structural
`Moieties (25,26)
`
`Species
`
`CH3 – H
`CH3CH2 – H
`vCH – H
`CH2
`CH2vCHCH2 – H
`PhCH2 – H
`H2N – H
`CH3NH – H
`CH3O – H
`PhNH – H
`HO – H
`HOCH2 – H
`HOCH(CH3) – H
`CNCH2 – H
`HO – OH
`CH3O – OCH3
`HOO – H
`
`Bond-Dissociation
`Energy (kcal/mol)
`
`104
`98
`103
`85
`85
`103
`92
`102
`80
`119
`93
`90
`86
`51
`36
`90
`
`Eton Ex. 1027
`6 of 32
`
`

`

`rm]
`~
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`
`
`HO lw /
`
`nitrogen can undergo
`electron transfer oxidation
`I
`
`adjacent to
`heteroatom
`
`-
`+ H CH3
`H zN~OC
`HO ~ /;
`::::.....
`N
`I
`HHH HHH
`\
`\
`' b 1-
`phenolic position can
`enzy 1c
`J
`d
`benzylic and tertiary and
`un ergo e ectron
`adjacent to
`adjacent to
`transfer
`oxidation and_ hydrogen heteroatom
`heteroatom
`atom abstractJon
`
`ORDER REPRINTS
`
`Oxidation Stabilization
`
`7
`
`is
`radicals
`stabilization of
`sp hybridized centers,
`generally poor due to the orthogonal arrangement of
`the p-orbitals with the unpaired electron. Referring to
`Table 4,
`the hydrogen bond energy for ethylene is
`similar to that of methane.
`Particularly stable radicals are formed from hydrogen
`atom abstraction of allylic or benzylic positions. These
`methylene/methine positions adjacent
`to aromatic or
`vinylic systems are very susceptible to hydrogen-atom
`abstraction because the resulting radical centers are
`stabilized through delocalization (see Table 4). Simi-
`larly, hydrogen atom abstraction from carbons adjacent
`to heteroatoms (nitrogen or oxygen) is facile due to
`electronic stabilization of the product.
`
`Other Oxidation Mechanisms
`
`Oxidation can also occur via an electron-transfer
`reaction (see “Electron Transfer”) to form reactive
`radical anions or cations. Nitrogen (amines), sulfur
`heteroatoms (sulfides, disulfides, and sulfoxides), and
`oxygen-based anions (phenol anions) are common sites
`for electron-transfer induced oxidation, producing final
`products such as N-oxides, sulfoxides, sulfones, and
`ketones. These sites can also act as nucleophiles to react
`with peroxide impurities in the formulation (see
`“Peroxide Reactions with Drugs”) giving similar
`products.
`
`Example
`
`As an example of using these rules to predict reactive
`centers, Labetalol is shown below with the potentially
`labile hydrogen atoms and other oxidation sites
`indicated:
`
`Computational Modeling
`
`Computational prediction of degradation can be
`extremely helpful in the understanding of degradation
`
`mechanisms and the characterization of degradation
`products found by purposeful degradation studies (see
`“Purposeful Degradation”). Several commercial compu-
`ter programs designed to predict reaction products are
`available including elaboration of reactions for the
`organic syntheses (EROS)
`(27), workbench for
`the
`organization of data for chemical applications (WODCA)
`(28), and logic and heuristics applied to synthetic analysis
`(LHASA) (29). Since these programs are parameterized
`around the solution chemistry reactions, they may be less
`predictive for reactivity of drug molecules in the solid
`state. The EPIWIN (estimation program interface for
`Windowse) (30,31) is designed to predict the reactivity
`of organic molecules towards hydroxyl radicals (HOz).
`This approach can be useful in estimating the reactivity of
`a given molecule towards hydrogen abstraction; how-
`ever, its use in predicting actual degradation products and
`reaction rates has not been reported. One program found
`to be particularly useful is computer assisted mechanistic
`evaluation of organic reactions (CAMEO)
`(32). The
`CAMEO program computationally predicts the products of
`organic reactions given the starting materials, reagents,
`and conditions including the oxidative conditions.
`Solvent variation (protic/aprotic) and temperature
`changes can be considered, as well as more than one
`equivalent of reagents. The CAMEO degradation predic-
`tion results can be used as an initial guide to possible
`decay products. The CAMEO predicted decay products
`may not in reality be observed during the stability studies,
`and some degradants can be formed,
`that are not
`predicted in the calculations. In general, however, the
`CAMEO algorithms have been biased to predict more
`products than are likely to be observed. It is also likely
`that certain products predicted can undergo further
`decomposition. For example, primary and secondary
`hydroperoxides predicted in many oxidation processes
`typically undergo further reactions in actual degradation
`studies to produce ketones and alcohols. Although not
`reported in the literature, CAMEO allows for a standard
`oxidative set of reagents and conditions to be created and
`used routinely to predict oxidation products. This
`approach would allow for consistency for all compounds
`being evaluated. The reagents and conditions used do not
`necessarily need to mimic the real stability conditions,
`but rather could be predictive of those conditions. In spite
`of CAMEO’s limitations (only available for Macintoshes,
`limited to 64 total atoms), future adaptations promise to
`make this and other computational methods more
`routinely useful for pharmaceutical scientists.
`
`Eton Ex. 1027
`7 of 32
`
`

`

`roil
`~
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`ORDER REPRINTS
`
`8
`
`Waterman et al.
`
`MECHANISMS OF OXIDATION
`
`Autoxidation—Chain Processes
`
`The free-radical process of autoxidation consists of a
`chain sequence involving three distinct
`types of
`reactions: initiation, propagation, and termination.
`
`Initiation
`
`Initiation produces a free radical to begin a chain
`reaction (11,33,34). Initiation can occur due to homolytic
`cleavage of a weak bond such as with a polymeric
`excipient, by electron-transfer processes (see “Electron
`Transfer”), by light-induced reactions, or by metal
`catalysis (see “Catalysis”).
`In the solid state, special issues exist with regard to
`initiation. In particular, homolytic bond cleavage to
`give a pair of radicals (caged radicals) will generally
`not lead to free radicals since cage escape is strongly
`dependent on the diffusional mobility within a solid
`matrix. Because of this, initiation is slow to nonexistent
`in the solid state, though any radicals generated tend to
`survive for long periods since termination is slow. At
`this point,
`it
`remains unknown how much drug
`oxidation in the solid state is a product of chain
`processes vs. reaction of drug with oxidant impurities
`(see “Peroxides and Other Oxidizing Agents”) and
`mobile active oxygen species (hydroxyl, peroxyl, and
`superoxide) generated by metal catalysis.
`
`Propagation (11,34)
`
`Once a free radical is produced, its fate can include
`propagation by hydrogen-atom abstraction from the drug
`or excipient (solvent), propagation by a radical addition
`reaction to the drug or excipient, propagation by reaction
`with molecular oxygen to form a peroxyl
`radical,
`rearrangement, cyclization, or termination (see “Termin-
`ation”). The peroxyl radical itself can propagate through
`the same types of reactions.
`Propagation:
`Inz þ RH ! InH þ Rz
`Rz þ O2 ! ROOz
`Rz þ RH ! HR – Rz
`ROOz þ RH ! ROOH þ Rz
`
`Though hydrogen-atom abstraction from the drug is
`formally an oxidation process, it is not an autoxidation
`process since there is no oxygen in the reaction path.
`When oxygen is absent, the drug radical will reform the
`drug by abstraction of a hydrogen atom from other
`molecules in the system. When oxygen is involved,
`however,
`the hydroperoxide product rarely loses an
`oxygen molecule, such that irreversible drug decompo-
`sition occurs. In solution, reaction of a radical with
`molecular oxygen typically occurs with no activation
`barrier;
`i.e.,
`the reaction occurs at
`the diffusion-
`controlled rate of approximately 109 M 21sec21 depend-
`ing on the solvent viscosity (35,36).
`In solution,
`under normal atmospheric conditions, the rate-limiting
`step is hydrogen-atom abstraction,
`therefore the
`resulting rate expression is determined by the rate of
`radical
`formation (initiation), propagation (typically
`104 – 1027 M 21 sec21) (37,38) and termination, but is
`not dependent on the concentration of oxygen in solution.
`This is shown in the following rate expression for chain
`oxidation (37,38):
`kiÿ!Inz þ Inz
`In – In
`Inz þ R – Hfastÿ!Rz þ InH
`Rz þ Rz fastÿ!R – R typically 109 – 1010 M 21 sec21
`fastÿ!R – OOz
`Rz þ O2
`typically 108 – 109 M 21 sec21
`
`R – OOz þ H – R kpÿ!R – OOH þ Rz
`
`typically 1 – 102 M 21 sec21
`
`R – OOz þ R – OOz ktÿ!R – OOH þ R
`ðRef:ð38ÞÞ
`
`typically 106 – 109 M 21 sec21
`
` 
`1=2½RHŠ
`
`d½RHŠ
`dt
`
`2
`
`¼ kp
`
`ki
`kt
`
`One of the by-products of this chain process is a
`hydroperoxide of either the drug or an excipient, which
`itself can act as an oxidant for the drug (see “Peroxides
`and Other Oxidizing Agents”). With oxygen present,
`the radical propagation (with oxygen consumption) can
`lead to large turnovers where every step creates a
`degradation product of the drug. Eventually, the build-
`
`Eton Ex. 1027
`8 of 32
`
`

`

`rm]
`~
`
`OH I
`
`H2C:?"' -
`
`#CH
`
`+
`
`0
`
`II
`
`/CH
`H3C
`
`R + R----- R-R
`Termination: R + ROD· ------ R-OOR
`/
`ROD· + ROD·-
`ROO-OOR
`R + R _ . , . disproportionatio~
`
`l + o, + Ho-b--
`
`H
`
`/
`
`\
`'--.
`(R= pri mary or secondary)
`ROOR + 0 2
`(R= teni ary)
`
`-
`
`H" ::=/H
`,..,..-N-.-"-: +HOO'
`R
`Ar
`
`j
`I (Q)
`i ·c t
`MARCEL DEKKER, INC. ~
`270 Madison Avenue, New York, New York I 0016 ~
`
`ORDER REPRINTS
`
`Oxidation Stabilization
`
`9
`
`up of hydroperoxides can lead to changes in the
`reaction mechanism, where the hydroperoxide com-
`petes with the drug in the hydrogen-atom abstraction
`step.
`In the solid state, propagation is hindered by low
`mobility. Reactions therefore tend to occur over
`extremely short distances with fewer chain propagation
`steps. Alternatively, solid-state oxidation can be
`propagated through diffusable species such as peroxide
`and superoxide. Although the products from these
`processes are often the same as for the chain process
`described above for solutions, the mechanism and rates
`in the solid state can be significantly different (see
`“Oxidation in Solid Dosage Forms”). Another impli-
`cation of having fewer propagation steps in the solid state
`compared to the solution state is that antioxidants
`designed to intercept the chain (see “Chain Termin-
`ators”) may be considerably less effective in the solid
`state. The low mobility of the solid state requires that an
`antioxidant be dispersed into each small domain to
`quench the reaction effectively.
`
`Termination (11,34)
`
`Termination of propagating radicals occurs on
`combination of two radicals to form nonradical products.
`This can manifest itself in radical combination reactions
`where two radicals form a new bond joining them
`together or by disproportionation where one radical is
`reduced while the other is oxidized. This latter process
`typically involves

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