throbber
Kinetics and Mechanism of the Reaction of Cysteine and
`Hydrogen Peroxide in Aqueous Solution
`
`DAYONG LUO,1 SCOTT W. SMITH,2 BRADLEY D. ANDERSON1
`
`1Department of Pharmaceutical Sciences, University of Kentucky, Lexington, Kentucky 40506
`
`2Pfizer Global Research and Development, San Diego, California 92121
`
`Received 10 August 2004; revised 22 September 2004; accepted 22 September 2004
`
`Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/jps.20253
`
`ABSTRACT: The oxidation of thiol-containing small molecules, peptides, and proteins in
`the presence of peroxides is of increasing biological and pharmaceutical interest.
`Although such reactions have been widely studied there does not appear to be a consensus
`in the literature as to the reaction products formed under various conditions, the reaction
`stoichiometry, and the reaction mechanisms that may be involved. This study examines
`the reaction kinetics of cysteine (CSH) with hydrogen peroxide (H2O2) in aqueous buffers
`(in the absence of metal ions) over a wide range of pH (pH 4–13) and at varying ratios of
`initial reactant concentrations to explore the range of conditions in which a two-step
`nucleophilic model describes the kinetics. The disappearance of CSH and H2O2 and
`appearance of cystine (CSSC) versus time were monitored by reverse-phase high-
`performance liquid chromatography (HPLC). The effects of oxygen, metal ions (Cu2þ), pH
`(4–13), ionic strength, buffer concentration, and temperature were evaluated. Data
`obtained at [H2O2]0/[CSH]0 ratios from 0.01–2.3 demonstrate that the reaction of CSH
`with H2O2 in the absence of metal ions is quantitatively consistent with a two-step
`nucleophilic reaction mechanism involving rate-determining nucleophilic attack of
`thiolate anion on the unionized H2O2 to generate cysteine sulfenic acid (CSOH) as an
`intermediate. Second-order rate constants for both reaction steps were generated
`through model fitting. At [H2O2]0/[CSH]0 > 10, the % CSSC formed as a product of the
`reaction declines due to the increased importance of alternative competing pathways for
`consumption of CSOH. A thorough understanding of the mechanism in aqueous solution
`will provide valuable background information for current studies aimed at elucidating
`the influence of such factors on thiol oxidation in solid-state formulations.
`ß 2004 Wiley-Liss, Inc. and the American Pharmacists Association J Pharm Sci 94:304–316, 2005
`Keywords:
`cysteine; hydrogen peroxide; thiol oxidation; peptide stability; protein
`stability; nucleophilic substitution; disulfide formation; sulfenic acid; sulfinic acid;
`sulfonic acid
`
`INTRODUCTION
`
`The oxidation of peptide and protein pharma-
`ceutical products has become an increasingly
`important problem in drug development as more
`biotechnology derived products progress toward
`clinical studies and commercialization.1–3 Pro-
`
`Correspondence to: Bradley D. Anderson (Telephone: 859-
`257-2300, ext. 235; Fax: 859-257-2489;
`E-mail: bande2@email.uky.edu)
`
`Journal of Pharmaceutical Sciences, Vol. 94, 304–316 (2005)
`ß 2004 Wiley-Liss, Inc. and the American Pharmacists Association
`
`teins containing one or more free cysteine resi-
`dues (e.g., human serum albumin,4 recombinant
`human a1-antitrypsin,5 the superfamily of pro-
`tein tyrosine phosphatases,6,7 and a variety of
`others8), as well as thiol-containing peptides (e.g.,
`glutathione9,10 the angiotensin-converting enzyme
`inhibitor captopril11 and the prostate-specific
`antigen peptides12) are particularly susceptible
`to thiol-oxidation through a variety of mechan-
`isms. In vitro, these mechanisms may involve
`free-radical scavenging of various reactive oxy-
`gen species such as superoxide9,10 or hydroxyl
`
`304
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`Eton Ex. 1056
`1 of 13
`
`

`

`REACTION OF CYSTEINE AND HYDROGEN PEROXIDE IN AQUEOUS SOLUTION
`
`305
`
`radical,13 resulting in thiol loss and the consump-
`tion of oxygen. Alternatively, direct reaction of
`thiols via nucleophilic substitution with certain
`reactive oxygen species such as hydrogen per-
`oxide may occur.4,5,7,14,15
`Because of their reactivity as scavengers of
`reactive oxygen species, thiol-containing com-
`pounds can be highly effective as antioxidants for
`stabilizing pharmaceuticals.13,16–20 However, this
`propensity toward oxidation makes formulation of
`proteins and peptides containing free cysteine
`residues more difficult. In pharmaceutical formu-
`lations, thiol autoxidation is generally thought to
`involve a series of complex reaction processes
`catalyzed by traces of transitional metal
`ions
`(e.g., CuII and FeIII).11,21–23 For this reason, metal
`ion chelators such as EDTA have proven to be
`effective stabilizers of thiol containing peptides
`and proteins.24–26 Considering, however, that
`peroxides are also found as impurities in some
`common formulation excipients (e.g., polyvinyl-
`pyrrolidone, polyethylene glycol, polysorbate 80,
`etc.),27–34 and that H2O2 is produced as a bypro-
`duct of metal-catalyzed oxidations,35–37 an under-
`standing of the mechanism of the thiol–H2O2
`reaction is a prerequisite for understanding the
`general case of thiol oxidation in formulations.
`Our interest in stabilizing thiol-containing
`peptides stems from recent attempts to prepare
`lyophilized prototype formulations of several pros-
`tate-specific antigen (PSA) peptides under con-
`sideration by the National Cancer Institute for
`vaccine therapy for human prostate cancer,12,38
`including the cys ras peptide (KLVVVGAC-
`GVGKS),25 PSA-3 (VISNDVCAQV),23,25 and
`PSA-OP (FLTPKKLQCVDLHVISNDVCAQVH-
`PQKVTK).26 Although EDTA was found to be
`highly effective in preventing metal-catalyzed
`disulfide formation in these lyophiles, the poten-
`tial for the reaction of cysteine residues with trace
`peroxide impurities in excipients remains even in
`the presence of EDTA.
`Several studies of the reaction between various
`low molecular weight thiols (e.g., cysteine, N-
`acetylcysteine, glutathione, and others) and hy-
`drogen peroxide (H2O2) or other peroxides (e.g.,
`benzoyl peroxide, lipid hydroperoxides, etc.) have
`been reported, but there does not appear to be a
`consensus as to the reaction products formed
`under various conditions,10,14,15,39 –42 the mechan-
`ism(s) by which these products form,10,14,15,41 –44 or
`the rate constants for various reaction steps.
`This study examines the reaction kinetics of
`cysteine with H2O2 in aqueous buffers (in the
`
`absence of metal ions) over a wide range of pH
`(pH 4–13) and at varying ratios of initial reactant
`concentrations to explore the range of conditions
`over which a two-step nucleophilic model as
`depicted in Scheme I accounts for the reaction
`kinetics and stoichiometry. In aqueous solution
`the first step in Scheme I is rate-determining, but
`it may be possible to estimate rate constants for
`each process, depending on the extent to which the
`two reaction steps differ.
`
`Scheme I.
`
`A thorough understanding of the mechanism of
`reaction between low molecular weight thiols and
`H2O2 in aqueous solution should be useful in
`rationalizing the fate of thiol-containing proteins
`undergoing reaction with hydrogen peroxide,
`where the buildup of the sulfenic acid intermediate
`(RSOH) suggests that the second reaction step
`may become rate-limiting in some circumstances.
`This may reflect the increased importance of
`reactant mobility in the second reaction step.
`Similarly, the reaction of the RSOH intermediate
`with a thiol residue in a second molecule of a thiol-
`containing peptide or protein may also become rate
`limiting in amorphous (e.g., lyophilized) pharma-
`ceutical formulations that contain peroxide impu-
`rities. Evidence that the disulfide may not be the
`sole reaction product under certain conditions led
`to an exploration of the possible factors influencing
`the fate of the sulfenic acid intermediate (RSOH)
`and the reaction products ultimately formed when
`RSOH reacts with H2O2 or another molecule of
`RSOH.
`
`EXPERIMENTAL
`
`Reagents
`
`Cysteine (>98% purity by TLC), cystine (Sigma-
`Ultra, >99% purity by TLC), cysteine sulfinic acid
`(water content 1 mol/mol), cysteic acid (mono-
`hydrate), and ethylenediaminetetraacetic acid
`(EDTA) were purchased from Sigma Chemical
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`Eton Ex. 1056
`2 of 13
`
`

`

`306
`
`LUO, SMITH, AND ANDERSON
`
`Co. (St. Louis, MO). H2O2 (30% aqueous solution)
`and cupric sulfate (CuSO4 5H2O) were obtained
`from Mallinckrodt Baker Inc. (Phillipsburg, NJ).
`All reagents were analytical grade and used as
`supplied. Deionized ultra-filtered (DIUF1) water
`used to prepare CSH and H2O2 solutions and
`HPLC grade acetonitrile were purchased from
`Fisher Scientific Co. (Pittsburgh, PA). Water used
`in mobile phase was deionized and further
`purified through a Milli-Q1 UV Plus Ultrapure
`Water System, Millipore Ltd. (Billerica, MA).
`
`High-Performance Liquid Chromatography (HPLC)
`
`The analyses of CSH, CSSC, and H2O2 were
`carried out by reverse-phase high-performance
`liquid chromatography (HPLC). The system con-
`sisted of a Beckman 110A Solvent Delivery
`Module (Beckman Coulter, Fullerton, CA), a
`WatersTM 717 plus Autosampler (Waters Corp.,
`Milford, MA), and a Hewlett-Packard 1040M
`Series II HPLC detector (Hewlett-Packard Com-
`pany, Palo Alto, CA) operating at 214 nm, which
`was connected to a PC for data acquisition and
`analysis using HP LC/MSD ChemStation soft-
`ware. A Supelcosil ABZþPlus, 5-mm (250
`4.6 mm)
`(Supelco, Bellefonte, PA) analytical
`column and 5 mm (2 cm 2.1 mm) guard column
`were employed. Elutions were performed isocra-
`tically using a mobile phase consisting of 50%
`(v/v) of an aqueous solution of 50 mM phosphoric
`acid (solution A) and 50% (v/v) of a solution
`containing 5 mM sodium 1-nonanesulphonate
`(99%, Lancaster Synthesis, Inc., Windham, NH),
`50 mM phosphoric acid, and 5% (v/v) acetonitrile
`in water (solution B) at a flow rate 1.5 mL/min.
`Solution B was also used as the quench solution.
`The pH of the mobile phase was adjusted to 2.5
`with NaOH. The retention times for H2O2, CSH,
`and CSSC were 2 min, 4.8 min, and 6.5 min,
`respectively. Cysteine sulfinic acid (CSO2H) and
`cysteic acid (CSO3H) eluted prior to H2O2 at
`1.8 min, but could not be adequately resolved
`for quantitation purposes under these conditions.
`Separations of CSO2H from CSO3H required two
`Supelcosil ABZþPlus columns connected in series
`and a flow rate of 0.5 mL/min. Under these condi-
`tions, retention times for CSO2H and CSO3H were
`10.4 and 11 min, respectively.
`
`General Procedure for Kinetic Studies
`
`Fresh stock solutions of CSH were prepared by
`weighing 20 mg CSH into a 1.5-mL plastic
`microcentrifuge tube (VWR Scientific Products
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`Co., Buffalo Grove, IL) and dissolving in 1 mL
`phosphate buffer prepared from phosphoric acid
`solution (50 mM) adjusted to the desired pH with
`sodium hydroxide. Stock solutions of H2O2 were
`prepared by combining 0.1 mL 30% aqueous H2O2
`solution and 0.9 mL of the same phosphate buffer
`(50 mM) in a 1.5-mL plastic microcentrifuge tube.
`Reactions were initiated after first diluting the
`above stock solutions with the same buffer to a
`concentration twice that in the final reaction,
`equilibrating to the desired reaction temperature
`in a water bath set at 258C. Equal volumes of each
`diluted solution were then combined by adding the
`diluted H2O2 solution to the CSH solution and
`rapidly mixing in a 15 mL plastic centrifuge tube
`(Falcon1, Becton Dickinson Co., Franklin Lakes,
`NJ). Reaction solutions were stored at 258C and
`aliquots (0.1 mL) were taken at predetermined
`time intervals and transferred into 1-mL HPLC
`vials. An aliquot of 0.1-mL solution B was added to
`quench the reaction by lowering the pH and the
`total volume was adjusted to 1 mL with the mobile
`phase. Samples were analyzed immediately by
`HPLC.
`Kinetic studies to obtain a preliminary assess-
`ment of reaction order, metal ion effects, and the
`influence of oxygen utilized the initial rate method
`whereby the formation of cystine was monitored at
`several time points in the very early stages of the
`reaction (first minute) at varying initial concen-
`trations of CSH and H2O2. More comprehensive
`studies for use in fitting mathematical models to
`the data (see Data Analyses), generating pH-rate
`profiles, and examining the effects of buffer
`concentration and ionic strength involved the
`determination of complete reactant (i.e., CSH
`and H2O2) and product concentration versus time
`profiles.
`
`Metal Ion and Oxygen Effects
`
`Solutions containing equal (5 mM) concentrations
`of CSH and H2O2 and varying in EDTA concen-
`tration (0, 25, and 50 mM) were prepared by
`combining stock solutions in pH 6.0 (50 mM)
`sodium phosphate buffer either with or without
`addition of 50 mM cupric sulfate (CuSO4). Buffers
`employed in the absence of added cupric ion were
`prepared using either Milli-Q treated deionized
`water or DIUF1 water (see Reagents).
`The influence of oxygen was explored using
`initial rate studies both in air-equilibrated sam-
`ples and samples prepared and stored under a
`nitrogen atmosphere (in a glove box). Sample
`
`Eton Ex. 1056
`3 of 13
`
`

`

`REACTION OF CYSTEINE AND HYDROGEN PEROXIDE IN AQUEOUS SOLUTION
`
`307
`
`equation to obtain an estimate for the energy of
`activation.
`
`Effect of [CSH]0/[H2O2]0 Ratio on Kinetics and
`Reaction Products
`
`To further explore the range of applicability of
`the simple nucleophilic mechanism depicted in
`Scheme I, additional experiments were carried
`out in which the ratio of starting concentrations of
`CSH and H2O2 ([CSH]0/[H2O2]0) was varied from
`100 to 0.001. A ratio of [CSH]0:[H2O2]0¼ 100
`using 50 mM CSH and 500 mM H2O2 at pH 6.0
`was employed to assess the effect of a large excess
`of CSH on the rate of disappearance of H2O2. The
`effects of [CSH]0:[H2O2]0 ratios 1.0 (i.e., 1 to
`0.001) were evaluated at pH 5.0, 6.0, and 7.0 to
`explore the possible generation of new reaction
`products in the presence of excess H2O2. The
`concentration of [CSH]0 in these experiments was
`3 mM at [CSH]0:[H2O2]0 ratios of 1 and 0.5 and
`1 mM at ratios between 0.2 and 0.001. Solution
`aliquots were taken after the reactions were
`complete (72 h at pH 7.0, 96 h at pH 6.0, and
`120 h at pH 5.0) and analyzed by HPLC for the
`CSSC product formed during the reaction. Ana-
`lyses of some of the samples at later time points
`verified that the CSSC concentrations did not
`change after completion of the reaction.
`
`DATA ANALYSIS
`
`temperatures were controlled by placing samples
`in a water jacketed container at 258C. An OM-4
`oxygen meter (Microelectrodes, Inc., Bedford, NH)
`was used to verify the removal of oxygen from
`diluted stock solutions that were bubbled with
`nitrogen prior to being combined to start a given
`reaction. The oxygen meter was calibrated by
`adjusting the reading to 0% in buffer solutions
`treated by bubbling with a N2 stream for more than
`30 min (only 20 min was necessary to achieve a
`constant reading) and to 100% in oxygen saturated
`buffer solutions. Experiments were at pH 6.0 and
`at a fixed starting concentration of CSH (25 mM)
`while H2O2 starting concentrations were varied
`(i.e., 5, 10, 25, and 50 mM) or a fixed starting
`concentration of H2O2 (10 mM) while CSH
`starting concentrations were varied (i.e., 2, 5, 10,
`and 25 mM).
`
`Buffer, Ionic Strength, pH, and
`Temperature Effects
`
`Phosphate buffers (pH 7) were prepared at dif-
`ferent buffer concentrations (10, 30, and 50 mM)
`and at a constant ionic strength of 0.5 M (adjusted
`with sodium chloride (NaCl)). The kinetics of
`reaction in solutions containing 4 mM CSH and
`4 mM H2O2 were monitored as a function of
`buffer concentration and second-order reaction
`rate constants generated through model fitting
`were compared.
`The effect of ionic strength was explored by
`comparing the kinetics of reaction between 4 mM
`CSH and 4 mM H2O2 in 50 mM sodium phosphate
`buffers at pH 5.0, 7.0, 10.0, and 13.0 either in the
`absence of added NaCl or in buffers adjusted to an
`ionic strength of 0.5 M with NaCl. Second-order
`rate constants generated through model fitting of
`the experimental data were then compared.
`The influence of pH on the reaction between
`4 mM CSH and 4 mM H2O2 was monitored in
`buffers ranging from pH 4–13 at 258C. Buffers
`were prepared with 50 mM phosphoric acid
`solution and sodium hydroxide (NaOH). Ionic
`strength was not adjusted. The observed reaction
`rate constants (kobs) obtained through the model
`fitting of the experimental data were employed to
`construct the pH-rate profile.
`Reactions between CSH and H2O2 were con-
`ducted at different temperatures (0, 25, and 508C)
`in pH 6.0 buffer solution using an ice bath for 08C, a
`water bath at 258C and an oven at 508C. Rate
`constants (k1) generated from model fitting (see
`Data Analysis) were then fit to the Arrhenius
`
`Concentration versus time curves for the disap-
`pearance of CSH and H2O2 and appearance of
`CSSC were fit to a mathematical model derived
`from Scheme I using nonlinear least-squares
`regression analysis (SCIENTIST1, Micromath
`Inc., Salt Lake City, UT) to obtain estimates for
`the two second-order rate constants, k1 and k2.
`The following differential equations can be gen-
`erated from Scheme I:
`d½CSŠ
`dt
`
`¼ k1 CS

`
`
`
`Š H2O2½
`
`Š þ k2 CS

`
`
`
`Š CSOH½ Š
`
`
`
`ð1Þ
`

`

`d CSOH
`dt
`
`d HOOH½
`
`
`dt
`

`
`¼ k1 CS

`
`
`
`Š H2O2½
`

`
`ð2Þ
`
`¼ k1 CS

`
`
`
`Š H2O2½
`
`Š k2 CS

`
`
`
`Š CSOH½
`
`Š ð3Þ
`


`d CSSC
`dt
`
`¼ k2 CS

`
`
`
`Š CSOH½
`

`
`ð4Þ
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`Eton Ex. 1056
`4 of 13
`
`

`

`308
`
`LUO, SMITH, AND ANDERSON
`
`where [CS], [H2O2], [CSOH], and [CSSC] are the
`concentrations of cysteine thiolate anion, hydro-
`gen peroxide, the cysteine sulfenic acid inter-
`mediate, and cystine, respectively, at time t.
`
`RESULTS AND DISCUSSION
`
`Simultaneous analysis of CSH, H2O2,
`and CSSC by HPLC
`
`The reaction between CSH and H2O2 has been
`previously studied by several other groups.10,14,43,45
`Normally in these studies, the concentrations of
`thiol and H2O2 were measured by colorimetric
`methods and CSSC concentration was not ana-
`lyzed. Although HPLC methods to separate CSH
`and CSSC46 and to separately determine H2O2
`concentration in various media34,47 have been
`reported, we are not aware of attempts to analyze
`all three reaction components in a single HPLC
`method. Figure 1 displays typical HPLC chroma-
`tograms obtained during the reaction of CSH with
`H2O2 leading to the formation of CSSC. Response
`factors were verified to be linear for all three
`compounds over concentration ranges of 0.025–
`2.5 mM for CSH, 0.1–10 mM for H2O2, and 3–
`400 mM for CSSC (solubility limited the range for
`CSSC) with intraday precision of <1% for all
`analytes. Interday precision in response factors
`was 1.6, 1.3, and 6.6% for CSH, H2O2, and CSSC,
`respectively. Detection limits at 214 nm estimated
`from three times the standard deviation for the
`lowest concentrations analyzed were 1.4 mM
`(0.14 nmol), 6.5 mM (.65 nmol), and 0.3 mM
`(0.03 nmol), respectively, for CSH, H2O2, and
`CSSC. These detection limits indicate that HPLC
`detection at 214 nm is not as sensitive for H2O2 as
`methods using electrochemical detection34 nor as
`sensitive for CSH and CSSC as HPLC with cou-
`lometric detection.46 However, Vignaud et al.46
`have observed that UV detection is more stable
`and easier to handle.
`
`Elimination of Metal Ions and Oxygen
`as Variables in the Kinetic Studies
`
`Because autoxidation of CSH (i.e., the direct
`reaction between CSH and molecular oxygen) is
`a spin-forbidden process, this reaction was not
`expected to contribute to the kinetics in this
`study.48 However, the presence of trace of transi-
`tion metal ions will significantly accelerate thiol
`autoxidation. Transition metal ions such as Cu2þ
`and Fe3þ not only catalyze autoxidation, but also
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`Figure 1. HPLC chromatograms (214 nm) at various
`times during the reaction of CSH with H2O2. Conditions:
`[CSH]0¼ 4 mM, [H2O2]0¼ 2 mM in pH 6.0 (50 mM)
`phosphate buffer at 258C. Times: (A) 15 s; (B) 30 min; (C)
`90 min; (D) 900 min. [Color figure can be seen in the
`online version of this article, available on the website,
`www.interscience.wiley.com.]
`
`act as direct oxidants in thiol oxidation.22,36,37,49 –52
`Given the above, it is clear that transition metal
`ions had to be avoided to minimize the contribu-
`tion of alternate routes of thiol oxidation.
`The ability of the metal ion chelator EDTA to
`eliminate the effect of metal ions on the reaction
`between CSH and H2O2 was investigated in pH 6.0
`solutions containing equal (5 mM) concentrations
`of CSH and H2O2 and varying in EDTA concentra-
`tion (0, 25, and 50 mM) either with or without the
`addition of 50 mM cupric sulfate (CuSO4). Initial
`rates of formation of CSSC, shown in Figure 2,
`
`Eton Ex. 1056
`5 of 13
`
`

`

`REACTION OF CYSTEINE AND HYDROGEN PEROXIDE IN AQUEOUS SOLUTION
`
`309
`
`quantitative modeling of the reaction, a prelimin-
`ary study designed to ascertain the reaction order
`with respect to CSH and H2O2 was conducted.
`Initial rates of formation of CSSC during the first
`minute after mixing at 258C were obtained for
`solutions containing a fixed starting concentra-
`tion of CSH (25 mM) and varying H2O2 concen-
`tration (i.e., 5, 10, 25, and 50 mM) or fixed H2O2
`concentration (10 mM) and varying CSH concen-
`tration (i.e., 2, 5, 10, and 25 mM). Plots of CSSC
`concentration versus time (data not shown) were
`linear for every combination, but in some cases,
`particularly at high reactant concentrations, the
`extent of the reaction as determined by either
`CSSC formation or loss of CSH exceeded 20%.
`Therefore, the slopes of the CSSC concentration
`versus time plots (d[CSSC]/dt) were normalized
`by dividing by the average concentration of the
`fixed reactant to adjust for the fact that mean
`reactant concentrations were less than 100% even
`during the first minute of the reaction. The slopes
`obtained appeared to be proportional to the total
`concentrations of each reactant in all ionization
`states, [CSH]t and [H2O2]t, indicating that the
`following rate law may be applicable:
`Št
`Š=dt ¼ kobs CSH½
`
`Št H2O2½

`d CSSC
`Slopes of log–log plots of the normalized rates of
`CSSC formation versus the variable reactant con-
`centration were 0.96 (0.01) versus 0.94 (0.03) for
`varying H2O2 concentrations under nitrogen ver-
`sus air, respectively, and 0.92 (0.01) versus 0.92
`(0.02), for varying CSH concentrations under
`nitrogen versus air, respectively. None of these
`slopes were significantly different from 1.0 as
`determined by the overlap of their 95% confidence
`intervals with 1.0, indicating that the reaction
`appears to be first-order in terms of each reactant.
`A more comprehensive set of experiments was
`then performed in which the concentrations of CSH,
`H2O2, and CSSC were monitored versus time and fit
`to the model represented by eqs 1–4 and depicted in
`Scheme I. Initially, a set of five experiments at a fixed
`pH (6.0) and CSH concentration (4 mM) and varying
`concentrations of H2O2 (2, 4, 6, 8, and 9.2 mM) were
`conducted at 258C so that pH effects could be
`neglected. Panels representing three of these experi-
`ments are shown in Figure 3.
`Evident from all three plots, exactly 2 mol of
`CSH are consumed per mol of H2O2, consistent
`with Scheme I and the overall stoichiometry for the
`reaction shown below:
`2CSH þ H2O2 ! CSSC þ H2O
`
`ð5Þ
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`Figure 2.
`Initial rates of CSSC formation in solutions
`containing equal (5 mM) concentrations of CSH and
`H2O2 versus EDTA concentration (0, 25, and 50 mM) in
`pH 6.0 (50 mM) sodium phosphate buffer either with (&)
`or without (~, }) addition of 50 mM cupric sulfate
`(CuSO4). Buffers employed in the absence of added cu-
`pric ion were prepared using either Milli-Q treated
`deionized water (~) or commercially obtained deionized
`ultrafiltered (DIUF1) water (}).
`
`indicate that the addition of cupric ion (Cu2þ)
`dramatically increases the reaction rate in the
`absence of EDTA, but that addition of EDTA
`suppresses this effect. In the presence of 25 mM
`EDTA, the rate of CSSC formation with the
`addition of Cu2þ is identical to the rate in deionized
`water containing no added cupric ion. EDTA addi-
`tion had no apparent effect itself on the reaction
`rate in the absence of added metal ion but did
`interfere with the HPLC analysis of hydrogen
`peroxide. Because the reaction rate was the same
`with or without EDTA when a high purity com-
`mercially available deionized ultrafiltered water
`(DIUF1, see Reagents) was employed, subsequent
`kinetic studies in which H2O2 was monitored were
`conducted without EDTA addition.
`Initial rates of CSSC formation were monitored
`in solutions at pH 6.0 at a fixed starting concen-
`tration of CSH (25 mM), while H2O2 starting
`concentrations were varied or at a fixed start-
`ing concentration of H2O2 (10 mM) while CSH
`starting concentrations were varied (see rate law
`studies). Replicate experiments were performed in
`air-equilibrated samples and in samples prepared
`and stored under a nitrogen atmosphere. No signi-
`ficant differences were observed in any of the re-
`actant solutions, indicating that air oxygen is not
`involved in the reaction between CSH and H2O2.
`
`Rate Law, Stoichiometry, and Reaction Mechanism
`
`Prior to generating complete reactant and product
`concentration versus time profiles for use in
`
`Eton Ex. 1056
`6 of 13
`
`

`

`310
`
`LUO, SMITH, AND ANDERSON
`
`Figure 3. Concentration versus time plots of cysteine
`(&) cystine (~), and hydrogen peroxide (*) during the
`reaction of CSH ([CSH]0¼ 4 mM) with varying concen-
`trations of H2O2 in 50 mM pH 6.0 phosphate buffer at
`258C. (A) [H2O2]0¼ 2 mM; (B) [H2O2]0¼ 4 mM; (C).
`[H2O2]0¼ 9.2 mM. The curves are nonlinear least-
`squares best fits using the kinetic model depicted in
`Scheme I (eqs. 1–4).
`
`The solid curves in Figure 3A–C represent
`simultaneous nonlinear least-squares fits of the
`data to the model depicted in Scheme I. Assuming
`a pKa for the cysteine sulfhydryl ionization of 8.44
`(see later discussion), and correcting the pKa for
`ionic strength, values for the two second-order
`reaction rate constants at pH 6.0 and 258C
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 94, NO. 2, FEBRUARY 2005
`
`generated through model fitting are k1¼ 15.2
`0.1 M1s1, k2¼ 720 70 M1s1. The excellent fit
`of the data strongly supports the nucleophilic
`mechanism that Scheme I describes.
`The apparent first-order dependence on CSH
`and H2O2 concentrations determined in the initial
`rate studies can only be consistent with Scheme I if
`the first step in the reaction, the nucleophilic
`attack of thiol (or thiolate anion) on H2O2 to
`generate a highly reactive sulfenic acid intermedi-
`ate is the rate-determining step. The relative
`values obtained for k1 and k2 (k1 k2) are consis-
`tent with the first step being rate-determining. To
`our knowledge, this is the first attempt to obtain a
`quantitative estimate for the bimolecular rate con-
`stant for thiolate anion attack on cysteine sulfenic
`acid to generate cystine.
`Among those studies in the literature that have
`concluded that hydrogen peroxide reacts with
`thiols to produce exclusively the corresponding
`disulfides there is disagreement as to the reaction
`stoichiometry and mechanism. Although Darkwa
`et al.43 suggested that the oxidation of cysteine by
`H2O2 is predominantly free radical-mediated,
`most groups have generated results consistent
`with a two-step nucleophilic substitution mechan-
`ism as described in Scheme I.10,14,15,44 As demon-
`strated in Figure 3A–C,
`the overall molar
`stoichiometry (H2O2:RSH) of the nucleophilic
`mechanism is 1:2. However, Abedinzadeh et al.
`have argued that this stoichiometry depends on
`the reactant ratios, such that, at initial reactant
`concentration ratios ([RSH0]/[H2O2]) > 2.5 the
`expected 1:2 stoichiometry is observed while at
`ratios <2.5 (i.e., excess peroxide) the stoichiometry
`is 1:1.41,42 The study described in Figure 3C
`contained excess peroxide yet the overall stoichio-
`metry remained 1:2.
`Abedinzadeh et al.41,42,53 also reported that the
`concentration of H2O2 undergoes abrupt decreases
`in the initial stage of reactions between glu-
`tathione or N-acetylcysteine, which they attribu-
`ted to the formation of complexes between RSH
`and H2O2. It is evident in Figure 3A–C that there
`is no abrupt decrease in hydrogen peroxide con-
`centration in the early stage of this reaction.
`Moreover, the excellent fit of the data to the model
`represented in Scheme I, which does not consider
`complex formation, suggests that a more compli-
`cated mechanism is unnecessary. However, apart
`from the fact that different thiols were the subject
`of their investigations, Abedinzadeh et al. mon-
`itored their reactions using absorbance measure-
`ments, whereas these studies employed HPLC.
`
`Eton Ex. 1056
`7 of 13
`
`

`

`REACTION OF CYSTEINE AND HYDROGEN PEROXIDE IN AQUEOUS SOLUTION
`
`311
`
`Further studies are necessary to rationalize these
`apparently disparate results.
`
`Buffer, Ionic Strength, pH, and
`Temperature Effects
`
`Parallel experiments were conducted in 50 mM
`phosphate buffers at different pH values (pH 5, 7,
`10, and 13) with or without control of
`ionic
`strength. In the absence of ionic strength adjust-
`ment, 50 mM phosphate buffers have an ionic
`strength ranging from 0.05 at pH 4 to 0.43 and
`pH 13. In the ionic strength-controlled buffer
`solutions, ionic strength was adjusted to 0.5 M
`with addition of NaCl. The experimental results
`in both groups of reactions were similar after
`correcting for ionic strength effects on pKa values
`(i.e., no statistically significant difference was
`observed). The absence of an ionic strength effect
`on the reaction between CSH and H2O2 is
`consistent with a previous study of the reaction
`of 2-mercaptoethanol with H2O2 as a function of
`ionic strength over the range of 0.05 M <I <0.4 M
`reported by Leung et al.15
`The kinetics of reaction in solutions containing
`4 mM CSH and 4 mM H2O2 were monitored as a
`function of buffer concentration in phosphate
`buffers (pH 7) varying in concentration (10, 30,
`and 50 mM) and at a constant ionic strength of
`0.5 M [adjusted with sodium chloride (NaCl)].
`Fitted values of k1 were 14.0, 17.4, and 20.5 M1s1
`in 10, 30, and 50 mM buffer, respectively. Although
`the 95% confidence intervals at 10 and 50 mM did
`not overlap, indicating that the differences were
`significant, the dominant contribution to k1 is the
`nonbuffer-catalyzed reaction.
`The influence of pH on the reaction between
`4 mM CSH and 4 mM H2O2 was monitored at 258C
`in buffers ranging in pH from 4–13. Concentration
`versus time curves such as those shown in Figure 3
`were generated at each pH and fit simultaneously
`to obtain estimates for the thermodynamic pKa
`values of cysteine and hydrogen peroxide, and
`values for k1 and k2 reflecting the entire pH range.
`The pKa value found for cysteine, 8.44 0.02, is in
`reasonable agreement with literature estimates of
`8.3354 and 8.53,55 while the pKa obtained for H2O2,
`11.51 0.02, also agrees favorably reported values
`of 11.4515 and 11.6.56 The values found for k1
`(¼14.7 0.35 M1s1) and k2 (¼570 170 M1s1)
`are not significantly different from those obtained
`at pH 6.0 in which five groups of kinetics experi-
`ments varying in CSH and H2O2 concentration
`were analyzed. Shown in Figure 4 is the pH
`
`Figure 4. pH dependence of the rate constant for the
`nucleophilic reaction between cysteine and hydrogen
`peroxide, k1, at 258C.
`
`rate profile generated from the pH 4–13 data,
`where kobs¼ k1*fCS *fHOOH and fCS and fHOOH
`are the fractions of thiolate anion and the neutral
`species of hydrogen peroxide, determined from
`their pKa values. The solid line (which was not
`fitted to the points) in Figure 4 represents the k1
`(¼14.9 M1s1) obtained from averaging the fits of
`the pH 4–13 and pH 6 concentration versus time
`data and the pKa values found from those previous
`fits (i.e., 8.44 for cysteine and 11.51 for H2O2).
`Our estimate of k1 (¼14.9 M1s1) at 258C
`appears to be in reasonable accord with literature
`estimates at other temperatures. Barton et al.14
`obtained a value of 12.4 M1s1 at room tempera-
`ture, while Winterbourn and Metodiewa10 esti-
`mated k1 to be 26 M1s1 at 378C. Radi et al.57
`estimated a k1 of 17.1 M1s1 at 378C.
`Values of k1 generated in this study through
`model fitting at different temperatures from 0–
`508C are well described by the Arrhenius equation
`(Fig. 5) with an activation energy (Ea) of 70.96 kJ/
`mol (16.96 kcal/mol). Alternatively, use of the
`Eyring equation yielded an enthalpy of activation,
`DH{, of 16.4 0.3 kcal/mol and entropy of acti-
`vation, DS{, of 1.7 1.1 cal K1mol1. These
`activation parameters result in an estimated k1
`of 42.6 M1s1 at 378C.
`The pH profile in Figure 4 covering a range of pH
`from 4–13 and the kinetic model used to fit these
`data are consistent with the results of Leung
`et al.,15 who examined the r

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