throbber
Controlling Deamidation Rates in a Model Peptide:
`Effects of Temperature, Peptide Concentration, and Additives
`
`LEWIS P. STRATTON,1,3 R. MICHAEL KELLY,1 JARED ROWE,1 JESSE E. SHIVELY,1 D. DAVID SMITH,2
`JOHN F. CARPENTER,1 MARK C. MANNING1
`
`1Center for Pharmaceutical Biotechnology and Department of Pharmaceutical Sciences, School of Pharmacy,
`Campus Box C238, University of Colorado Health Sciences Center, Denver, Colorado 80262
`
`2Department of Biomedical Sciences, School of Medicine, Creighton University, Omaha, Nebraska
`
`3Department of Biology, Furman University, 3300 Poinsett Highway, Greenville South Carolina 29613-0418
`
`Received 27 November 2000; revised 2 July 2001; accepted 10 July 2001
`
`ABSTRACT: The rate of deamidation of the Asn residue in Val-Tyr-Pro-Asn-Gly-Ala
`(VYPNGA), a model peptide, was determined at pH 9 (400 mM Tris buffer) as a function
`of temperature and peptide concentration. Over the temperature range 5–658C,
`deamidation followed Arrhenius behavior, with an apparent activation energy of
`13.3 kcal/mol. Furthermore, increasing the peptide concentration slows the rate of
`deamidation. Self-stabilization with respect to deamidation has not been reported
`previously. The rate of deamidation was also determined in the presence of sucrose and
`poloxamer 407 (Pluronic F127). In both cases, the rate of deamidation was retarded by
`up to 40% at 358C. In aqueous solutions containing poloxamer 407, the degree of
`stabilization is independent of formation of a reversible thermosetting gel. With sucrose,
`maximum reduction in the deamidation rate was attained with as little as 5% (w/v).
`Addition of sucrose results in a greater conformational preference for a type II b-turn
`structure, which presumably is less prone to intramolecular cyclization and subsequent
`deamidation. (cid:223) 2001 Wiley-Liss, Inc. and the American Pharmaceutical Association J Pharm Sci
`90:2141–2148, 2001
`Keywords: deamidation; poloxamer; gel; peptide stability; Arrhenius plot
`
`INTRODUCTION
`
`Deamidation of asparagine (Asn) residues is
`probably the most common pathway for chemical
`inactivation of protein pharmaceuticals1,2 The
`reaction rate for deamidation in aqueous solution
`is dependent on a number of extrinsic factors,
`such as pH,3–6 solvent dielectric,7 buffer concen-
`tration,5 and temperature,5 as well as intrinsic
`factors, such as the primary sequence8–10 and the
`presence of secondary2,11,12 and tertiary struc-
`ture.13 Despite numerous studies, little has been
`reported on the ability of additives to affect the
`
`Correspondence to: M.C. Manning (Telephone: 303-315-
`6162; Fax: 303-315-6281; E-mail: mark.manjing@uchsc.edu)
`
`Journal of Pharmaceutical Sciences, Vol. 90, 2141–2148 (2001)
`(cid:223) 2001 Wiley-Liss, Inc. and the American Pharmaceutical Association
`
`rate of deamidation in peptides and proteins. In
`the case of deamidation of human epidermal
`growth factor (hEGF), a wide variety of excipients
`were reported to have an effect, but with no
`explanation as to why some were effective and
`others were not.14
`It is now well established that the conformation
`of the peptide backbone (i.e., the secondary struc-
`ture), as well as the side chain dihedral angles,
`can affect the rate of deamidation.2,3 Therefore,
`one approach for slowing deamidation might be to
`control peptide conformation. One way this con-
`trol might be accomplished is through the use of
`preferentially excluded solutes. These solutes are
`additives that are selectively excluded from the
`hydration sphere of the protein. In other words,
`there will be a relatively higher concentration of
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`2141
`
`Novo Nordisk Ex. 2044, P. 1
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`2142
`
`STRATTON ET AL.
`
`the excluded solute in the bulk relative to near the
`surface of the protein. The mechanism by which
`these solutes stabilize proteins has been described
`in detail by Timasheff and co-workers.15–17
`It has also been demonstrated that preferen-
`tially excluded solutes can lead to compaction of
`the native structure of a globular protein.15,16 The
`same additives may be able to alter the conforma-
`tional distribution of a flexible peptide in aqueous
`solution.15 In short, one might be able to use
`any number of excluded solutes (e.g., polymers,
`sugars, salts, or amino acids) to affect the solution
`conformation of the peptide, thereby having an
`effect on the deamidation rate. If this effect can be
`demonstrated, it would provide a new strategy for
`stabilizing hydrolytically sensitive peptides in
`aqueous formulations. It is important to note that
`preferentially excluded solutes are maximally
`effective at relatively high concentrations of
`solute (as much as 1 M).
`Herein we describe the deamidation of a
`previously studied model peptide, Val-Tyr-Pro-
`Asn-Gly-Ala (VYPNGA),4–7,10 as a function of
`temperature, peptide concentration, and the addi-
`tion of solutes. We chose to examine two different
`excipients, sucrose and poloxamer 407, both of
`which are known to act as preferentially excluded
`solutes. Sucrose is well known to alter the confor-
`mation and flexibility of proteins through a prefe-
`rential exclusion mechanism.17 As has been
`shown for interleukin-1 receptor antagonist, addi-
`tion of sucrose produces a more compact protein
`structure,15,17 which is less prone to aggregation
`as well as deamidation.18 Similar effects have
`been observed for sucrose interacting with inter-
`feron-g.16 However, the effect of excluded solutes
`on the solution conformation of a peptide has not
`been assessed.
`The possible effects of poloxamer 407 (Pluronic
`F127), a triblock copolymer of poly(oxyethylene)
`and poly(oxypropylene), on deamidation of a
`peptide is less clear. Within a certain concentra-
`tion range, aqueous solutions of poloxamer 407
`will undergo a cooperative phase transition from a
`sol state to a gel with an increase in temperature.
`This transitition property has made it an attrac-
`tive vehicle for controlled drug delivery.19,20 It
`was hypothesized that poloxamer 407 might
`retard deamidation as well because it has been
`shown to salt out native proteins presumably due
`to preferential exclusion,19 but that the stabiliza-
`tion may also be related to the sol–gel transition
`rather than solely due to poloxamer 407 effects on
`peptide conformation.
`
`MATERIALS AND METHODS
`
`Materials
`
`Poloxamer 407 (Pluronic F127) was obtained from
`BASF and used as received. Sucrose was pur-
`chased from Pfannstiehl. All solvents were ob-
`tained from Aldrich.
`
`Peptide Synthesis
`
`Amino acid derivatives and resin were purchased
`from Applied Biosystems Incorporated (Foster
`City, CA). All other solvents and reagents were
`purchased from Fisher Scientific (Pittsburgh,
`PA). The peptide was assembled on Boc-Ala-
`PAM resin (0.5 meq) using an ABI 430A auto-
`mated peptide synthesizer. Boc groups were
`removed with 33% trifluoroacetic acid (TFA) in
`dichloromethane. Subsequent Boc amino acid
`derivatives were coupled to the resin in fourfold
`excess using diisopropylcarbodiimide and hydro-
`xybenzotriazole. Coupling reactions were moni-
`tored by the quantitative ninhydrin test.1 All
`yields were > 99% after a single coupling. Once
`the desired sequence was assembled, the peptide
`was simultaneously deprotected and cleaved from
`the resin using a mixture of TFA, trifluorometha-
`nesulphonic acid, ethanedithiol and thioanisole
`(80/8/4/8, v/v/v/v). The crude peptide product was
`loaded onto a Vydac 218TP101550 preparative
`C18 column (250 (cid:2) 50 mm, 300, 10–15 m) from
`The Separations Group (Hesperia, CA) that was
`previously equilibrated with triethylammonium
`phosphate buffer (100 mM, pH 2.5). The concen-
`tration of acetonitrile in the eluent was raised to
`11% over a period of 50 min. The eluent was
`continuously monitored at 230 nm and collected
`in 50 mL fractions. Fractions containing only the
`peptide were pooled, diluted twofold with water
`and desalted on the same column now equili-
`brated with 0.1% TFA. The concentration of
`acetonitrile was increased to 30% over a period
`of 30 min. Fractions containing the product were
`pooled and lyophilized to yield 150 mg of
`VYPNGA as a fluffy white powder (48% overall
`yield based on the initial resin substitution). FAB-
`MS [M (cid:135) H](cid:135) found 620.4, theoretical 620.3.
`
`HPLC Protocol
`
`The degradation of VYPNGA was monitored
`using a LKB Bromma high-performance liquid
`chromatography (HPLC) system with a Applied
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 2
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`FACTORS AFFECTING DEAMIDATION OF A MODEL PEPTIDE
`
`2143
`
`Biosystems variable wavelength detector set at
`214 nm. Separation was achieved using a Vydac
`reverse phase C-18 column (4.6 (cid:2) 25 mm length, 5
`mm bead diameter and 80 A˚ pore size). The mobile
`phase was a 24/76 mixture of water and 20%
`acetonitrile in 0.1% aqueous TFA. The flow rate
`was 1.5 mL/min. This protocol is similar to those
`reported previously for deamidation studies of
`VYPNGA.5,10
`
`Deamidation Studies
`
`Samples were prepared by dissolving a weighed
`amount of VYPNGA in 1 mL of 400 mM aqueous
`Tris buffer (pH 9), with or without poloxamer 407
`present. The sample was placed in a water bath to
`control the temperature. Periodically, samples
`were removed to assay for the extent of deamida-
`tion using the HPLC protocol already described.
`
`Circular Dichroism Studies
`
`Samples were prepared by dissolving 1 mg of
`VYPNGA in 1 mL of 400 mM Tris buffer contain-
`ing various amounts of sucrose or poloxamer 407.
`
`Samples were then placed in a 0.1 mm quartz
`cuvette. Far ultraviolet (UV) spectra (l & 250–
`180 nm) were collected using an AVIV 62-DS
`circular dichroism (CD) spectrometer with a ther-
`moelectric temperature control unit regulated to
`(cid:6)0.18C. Data were taken every 0.25 nm, using an
`averaging time of 4 s and a bandwidth of 1.5 nm.
`
`RESULTS AND DISCUSSION
`
`Deamidation of a polypeptide proceeds through a
`cyclic imide intermediate formed by intramolecu-
`lar attack of the succeeding peptide nitrogen on
`the carbonyl of the Asn side chain above pH 63,4
`(see Figure 1). This intermediate can hydrolyze,
`via nucleophilic attack of water, at either of the
`carbonyl groups; one attack will form a product
`containing normal aspartic acid (Asp) in place
`of the Asn residue, and one leads to a rear-
`ranged form of aspartic acid referred to as iso-
`aspartic acid (isoAsp). As a model of a peptide
`prone to deamidation, we chose the model hex-
`apeptide, VYPNGA, which has been used widely
`in the past.4–7,10 This peptide represents the
`
`Figure 1. General mechanism of deamidation of Asn residues under basic conditions.
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 3
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`2144
`
`STRATTON ET AL.
`
`deamidation site in the peptide hormone, ACTH.
`The chromatographic properties of this peptide
`and its degradation products are well estab-
`lished,4–7,10 making determination of the reaction
`kinetics straightforward. To have the reaction
`proceed quickly, we used conditions known to
`accelerate deamidation; namely, high pH and
`high buffer concentration. For this study, 400 mM
`Tris buffer at pH 9 was employed throughout.
`Under these conditions, a 1 mg/mL sample of
`VYPNGA will deamidate almost completely at 45
`8C within 6 h.
`The first studies were conducted to determine
`whether the mechanism of deamidation was
`altered by the use of high Tris concentrations at
`pH 9, conditions that have not been reported
`previously for this peptide. Therefore, the tem-
`perature dependence of the reaction was followed
`over a range spanning 5 to 658C. It should be
`noted that there is the possibility that Tris will
`decompose at elevated temperatures,21 producing
`reactive species, such as formaldehyde. However,
`there was no chromatographic evidence for the
`formation of species other than the deamidated
`forms of VYPNGA.
`Within experimental error, the reaction follows
`Arrhenius behavior (Figure 2). The activation
`energy was 13.3 kcal/mol, with an A-value of
`1014.9 (r2 (cid:136) 0.98), which is significantly less than
`the (cid:24)20 kcal/mol reported for this reaction.4,5
`However,
`those activation energies were for
`reactions extrapolated to zero buffer concentra-
`tion. At high buffer concentrations, buffer cata-
`lysis becomes significant, and the activation
`energy could be affected. The ratio of isoAsp to
`Asp was similar to that reported in the literature
`for VYPNGA,4,5,10 averaging (cid:24)3.0–3.5 to 1.0 at all
`temperatures tested, as determined by HPLC.
`
`Concentration Effects
`
`One aspect of peptide deamidation that has not
`been reported is whether peptide concentration
`has any effect. Presumably, increased concentra-
`tion could lead to aggregates that might be more
`stable or increased concentration may favor a
`solution conformation that is more or less reac-
`tive, depending on the relative spatial arrange-
`ment of the peptide backbone and Asn side chain.8
`On increasing the peptide concentration from 0.1
`to 100 mg/mL, an approximate threefold decrease
`in reaction rate was observed (Figure 3). A
`number of possible explanation exist for such an
`observation. First, it is possible that the peptide is
`interacting with itself, assembling to form a
`structure that is less reactive than the free
`monomer in solution. Second, small changes in
`pH (< 1 pH unit) could cause a modest decrease in
`rate, and high concentrations of the peptide may
`be able to shift the pH to this degree.22 However,
`no significant change in overall pH was measured
`for the concentrated solutions. Third, a peptide
`concentration of 100 mg/mL will cause increases
`in the viscosity of the solution. It has been
`reported that viscosity increase of <5 cP can slow
`the reaction to this extent.23 Whereas this most
`plausible explanation may be self-assembly, these
`other mechanisms cannot be excluded.
`
`Additive Effects-Sucrose
`
`Sucrose is well known to alter both the conforma-
`tion and flexibility of proteins through a prefer-
`ential exclusion mechanism.15–17 Briefly, an
`exclusion of solute, such as sucrose, means that
`the solute would rather exist in the bulk solution
`rather than at the surface of the protein. Such
`
`Figure 2. Arrhenius
`deamidation
`for
`plot
`VYPNGA (1 mg/mL) at pH 9 (400 mM Tris buffer).
`
`of
`
`Figure 3. Concentration effects on the rate of dea-
`midation of VYPNGA (pH 9, 400 mM Tris buffer, 358C).
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 4
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`FACTORS AFFECTING DEAMIDATION OF A MODEL PEPTIDE
`
`2145
`
`‘negative’ binding is thermodynamically destabi-
`lizing. The same exclusion exists for the unfolded
`state, as well as the native state. Because the
`unfolded state is typically larger in surface area
`than the folded state, the effect is larger in
`magnitude, meaning that
`the denatured or
`unfolded state is destabilized more than the
`native state. Together, these effects result in
`an increase in the free energy required to go
`from the folded to unfolded state; that is, there is a
`net stabilization. Moreover, for most proteins,
`excluded solutes also induce the protein to adopt a
`more compact native state.15,16 For interleukin-1
`receptor antagonist, the more compact protein
`structure is less prone to aggregation15 as well
`as deamidation.18 However, the effects of sucrose
`on peptide stability have not been well charac-
`terized. On addition of increasing amounts of
`sucrose,
`the reaction rate for deamidation
`decreased, even with as little as 2% sucrose
`present. The maximum effect could be achieved
`with as little as 5% sucrose (Figure 4). With 5%
`sucrose present, the relative deamidation rate
`was 77%. With 10% sucrose present, the relative
`rate was 78% of comparable solutions with no
`sucrose present.
`At these concentrations, lower rates due solely
`to increased viscosity is unlikely, especially
`because the rate-limiting step is intramolecular
`cyclization.6 At a sucrose concentration of 2%,
`there should be little change in the viscosity of an
`aqueous solution (< 5 cP). Although the spectro-
`scopic evidence given later suggests that rate
`retardation is due to a conformational change, the
`effects of the increased viscosity, although mod-
`est, cannot be excluded as having some modulat-
`
`ing effect on deamidation rates, especially in light
`of the work of Li et al.23
`As just hypothesized, retardation of deamida-
`tion rates by an excluded solute suggests that the
`peptide has changed conformation, making intra-
`molecular cyclization to the succinimide inter-
`mediate more difficult. The effect of secondary
`structure on deamidation has been attributed to
`this type of effect, where the peptide and the
`asparagine side chain are in less favorable
`positions for intramolecular attack when in an
`ordered secondary structure.2,3,8 In this case, the
`deamidation site is not within a fixed secondary
`structure, but the distribution within the native
`state ensemble is being shifted.
`CD spectra were obtained for VYPNGA both in
`the absence and presence of sucrose at pH 9. The
`spectra were taken at 258C and within 10 min of
`sample preparation to minimize the extent of
`deamidation. A difference spectrum for a sample
`in 5% sucrose versus no sucrose reveals that a
`certain secondary structure is induced by the
`addition of sucrose (Figure 5). The induced
`structure is easier to detect using difference
`spectroscopy because the original CD spectra are
`dominated by contributions from the Tyr side
`chain. The difference CD spectrum shown a
`negative band near 225 nm and a positive band
`indicative of a type II b-turn
`near 200 nm,
`structure,24,25 presumably with Pro at position
`i (cid:135) 1. This structure would place the Asn residue
`at the corner of a type II turn. It is known that
`there is a subset of conformations that favors
`deamidation by positioning the side chain for
`attack by the succeeding nitrogen atom in the
`polypeptide chain.3 Such a structure would be
`
`Figure 4. Effect of added sucrose on the deamidation
`rate of VYPNGA (1 mg/mL) at pH 9, 400 mM Tris
`buffer) at 358C.
`
`Figure 5. Difference CD spectrum of VYPNGA in the
`(A) presence and (B) absence of 5% (w/v) sucrose (pH 9,
`400 mM Tris buffer, 258C).
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 5
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`2146
`
`STRATTON ET AL.
`
`predicted to slow deamidation for two reasons.
`First, the orientation of the carbonyl group of Asn
`relative to the succeeding amide nitrogen of the
`Gly residue is not favorable for nucleophilic
`attack. Second, the amide NH is presumably
`involved in stabilizing the b-turn, making its
`deprotonation more difficult. The deprotonation
`must occur to generate a species sufficiently
`nucleophilic to form the cyclic imide intermediate.
`Other researchers have observed that residues
`located in b-turns do deamidate more slowly than
`residues in unordered or random coil struc-
`ture.2,26,27 Analysis of the sequence preferences
`for certain amino acids in various positions of
`b-turns indicates that the combination of Pro-Asn
`is highly favored for occupying positions i (cid:135) 1 and
`i (cid:135) 2 in a type II b-turn.28,29
`Although these data do not constitute definitive
`proof for rate retardation arising from an altered
`solution conformation of a peptide, the magnitude
`of the rate changes and the substantial increase in
`the b-turn population present a compelling obser-
`vation. The ability of additives to modulate
`peptide conformation and thereby affect chemical
`degradation rates appears to be one possible
`formulation strategy. Recently, a similar finding
`was observed for active site oxidation retardation
`by the addition of sucrose.30
`
`Additive Effects- Poloxamer 407
`
`Few data have been reported on the ability of
`excipients to slow chemical degradation in pro-
`teins, especially for deamidation. Son and Kwon
`reported that numerous polymers were effective
`at retarding deamidation in hEGF.14 However, no
`mechanistic explanation was given and the
`additives were very different in chemical and
`physical properties (various polymers, surfac-
`tants, etc.), making generalization difficult.
`Poloxamer 407, when dissolved in water at
`high concentrations, has the ability to form either
`a fluid solution (at low temperatures) or a stiff
`gel (at higher temperatures). It was of interest
`to see if formation of an aqueous gel would
`retard deamidation in a peptide, particularly as
`a function of physical state. Relatively little is
`known about the detailed interaction of polox-
`amers and proteins. Recently, we demonstrated
`that poloxamer 407 is a preferentially excluded
`solute that is able to salt-out proteins when
`dissolved in aqueous buffer at high concentra-
`tions.19 Therefore, as an excluded solute, it should
`be possible for poloxamer 407 to affect the solution
`
`conformation of VYPNGA in the same fashion as
`sucrose, and thereby alter the rate of cyclic imide-
`mediated deamidation.
`At concentrations > 17% (w/w), solutions of
`poloxamer 407 can form stiff gels as the tempera-
`ture is increased. The process is highly coopera-
`tive, occurring within a narrow temperature
`range. This process can be monitored by observing
`the physical state of the solution or by spectro-
`scopic methods, such as Fourier transform infra-
`red spectroscopy.19 Using a 22% gel solution,
`which gels at (cid:24)15 8C,
`the deamidation of
`VYPNGA was investigated under conditions
`analogous to those in solution (pH 9, 400 mM
`Tris). Whether the poloxamer solution was in the
`gel or sol state, there was a decrease in the rate of
`deamidation by > 40% versus the peptide in
`aqueous solution. Above the gel transition tem-
`perature (358C),
`the deamidation rate was
`58.6 (cid:6) 2.9% (n (cid:136) 3) relative to solutions without
`poloxamer present. Below the gel
`transition
`temperature (58C), the relative deamidation rate
`was 45 (cid:6) 3% (n (cid:136) 2).
`In the presence of 22% poloxamer, the tem-
`perature dependence of the reaction was investi-
`gated for temperatures above the transition
`temperature of the gel, meaning all measure-
`ments were made on poloxamer solutions existing
`in the gel state. Under those conditions, the
`system displays Arrhenius behavior, and the acti-
`vation energy was determined to be 12.1 kcal/mol,
`with an A-value of 1014.2 (Figure 6, r2 (cid:136) 0.92),
`similar to that determined for VYPNGA in buffer
`alone. Considering the variance in the data, the
`
`Figure 6. Arrhenius
`of
`deamidation
`for
`plot
`VYPNGA (1 mg/mL) at pH 9 (400 mM Tris buffer) in
`a 22 % (w/w) aqueous solution of poloxamer 407.
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 6
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`FACTORS AFFECTING DEAMIDATION OF A MODEL PEPTIDE
`
`2147
`
`activation energies are probably not significantly
`different.
`Given that sucrose appears to slow deamida-
`tion by altering the conformational preference of
`the peptide, similar CD studies were conducted
`looking for evidence of a structural change in the
`presence of poloxamer 407. Difference spectra
`showed a similar far UV CD pattern as with the
`addition of sucrose, but much weaker. Given that
`the solute concentration was much greater (22
`versus 5%), it appears that although poloxamer
`can act as an excluded solute, it is much less
`effective than sucrose on a weight-to-weight basis.
`
`SUMMARY
`
`These results demonstrate that the addition of
`excluded solutes can alter the conformational
`preference of a flexible peptide while also altering
`its propensity to deamidate. The effect is analo-
`gous to the role of secondary structure on
`modulating deamidation by constraining back-
`bone dihedral angles. In the case of VYPNGA, the
`addition of sucrose drives the conformational
`ensemble toward a greater preference for a type-
`II b-turn structure, most likely around the Pro-
`Asn residues.28,29
`Furthermore, it is becoming clear that polox-
`amers, much like PEGs and other polymers, act as
`excluded solutes at high concentrations (as
`demonstrated herein and previously19). Possibly,
`the mechanism of stabilization is the same as with
`sucrose, although viscosity effects and small
`changes in pH cannot be ruled out. Certainly,
`one must consider the possibility for retardation
`of deamidation in such concentrated solutions is
`that it is simply a viscosity effect, even for
`poloxamer in the sol state.23 However, the results
`with sucrose suggest that the effect is conforma-
`tional, not a function of microviscosity, especially
`because the rate-limiting step is an intramolecu-
`lar event. Furthermore, the fact that the extent of
`stabilization of poloxamer solutions was similar in
`both the sol and gel states, where the viscosity is
`dramatically different, also argues against visc-
`osity playing a major role in slowing deamidation
`in flexible peptides. Together, these findings
`suggest that control of the conformational state
`of a peptide may not only retard physical pro-
`cesses, such as aggregation, but can also slow
`degradation due to hydrolytic reactions, such as
`deamidation.
`
`REFERENCES
`
`1. Manning MC, Patel K, Borchardt RT. 1998.
`Stability of protein pharmaceuticals. Pharm Res
`6:903–918.
`2. Xie M, Schowen RL. 1999. Secondary structure and
`protein deamidation. J Pharm Sci 88:8–13.
`3. Clarke S. 1987. Propensity for spontaneous succi-
`nimide formation from aspartyl and asparaginyl
`residues in cellular proteins. Int J Peptide Protein
`Res 30:808–829.
`4. Geiger T, Clarke S. 1987. Deamidation, isomeriza-
`tion, and racemization at asparaginyl and aspartyl
`residues in peptides. Succinimide-linked reactions
`that contribute to protein degradation, J Biol Chem
`262:785–794.
`5. Patel K, Borchardt RT. 1990. Chemical pathways of
`peptide degradation. II. Kinetics of deamidation of
`an asparaginyl residue in a model hexapeptide.
`Pharm Res 7:703–711.
`6. Xie M, Morton M, Vander Velde D, Borchardt RT,
`Schowen RL. 1996. pH-induced change in rate-
`determining step for the hydrolysis of the Asn/Asp
`derived cyclic imide intermediate in protein degra-
`dation. J Am Chem Soc 118:8955–8956.
`7. Brennan TV, Clarke S. 1993. Spontaneous degra-
`dation of polypeptides at aspartyl and asparaginyl
`residues: Effects of solvent dielectric., Protein Sci
`2:331–338.
`8. Wright TE. 1991. Sequence and structure determi-
`nants of the nonenzymatic deamidation of aspar-
`agine and glutamine residues in proteins. Protein
`Eng 4:283–294.
`9. Bischoff R, Lepage P, Jaquinod M, Cauet G, Acker-
`Klein M, Clesse D, Laporte M, Bayol A, Van
`Dorsselaer A, Roitsch C. 1993. Sequence specific
`deamidation: isolation and biochemical character-
`ization of succinimide intermediates of recombi-
`nant hirudin. Biochemistry 32:725–734.
`10. Patel K, Borchardt RT. 1990. Chemical pathways of
`peptide degradation.
`III. Effect
`of primary
`sequence on the pathways of deamidation of
`asparaginyl residues in hexapeptides. Pharm Res
`7:787–793.
`11. Lura R, Schirch V. 1988. Role of peptide conforma-
`tion in the rate and mechanism of deamidation of
`asparaginyl residues. Biochemistry 27:7671–7677.
`12. Stevenson CL, Friedman AR, Kubiak TM, Donlan
`ME, Borchardt RT. 1993. Effect of secondary
`structure on the rate of deamidation of several
`growth hormone releasing factor analogs. Int J
`Pept Protein Res 42:497–503.
`13. Kossiakoff AA. 1988. Tertiary structure is a
`principal determinant
`to protein deamidation.
`Science 240:191–194.
`14. Son K, Kwon C. 1995. Stabilization of human
`epidermal growth factor (hEGF) in aqeuous for-
`mulation. Pharm Res 12:451–454.
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 7
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

`

`2148
`
`STRATTON ET AL.
`
`15. Kendrick BS, Chang BS, Arakawa T, Randolph T,
`Manning MC, Carpenter JF. 1997. Preferential
`exclusion of sucrose from recombinant human
`interleukin-1 receptor antagonist: Role of re-
`stricted conformational mobility and compaction
`of native state. Proc Natl Acad Sci USA 94:11917–
`11922.
`16. Kendrick BS, Carpenter JF, Cleland JL, Randolph
`TW. 1998. A transient expansion of the native
`state precedes aggregation of recombinant human
`interferon-g. Proc Natl Acad Sci, USA 95:14142–
`14146.
`17. Lee JC, Timasheff SN. 1981. Stabilization of
`proteins by sucrose. J Biol Chem 256:7193–7201.
`18. Chang BS, Reeder G, Carpenter JF. 1996. Devel-
`opment of a stable freeze-dried formulation of
`recombinant
`interleukin-1 receptor antagonist.
`Pharm Res 13:243–249.
`19. Stratton LP, Dong A, Manning MC, Carpenter JF.
`1997. Drug delivery matrix containing native
`protein precipitates suspended in a poloxamer gel.
`J Pharm Sci 86:1006–1010.
`20. Suh H, Jun HW. 1996. Physicochemical and release
`studies of naproxen in poloxamer gels. Int J Pharm
`129:13–20.
`21. Song Y, Schowen RL, Borchardt RT, Topp EM.
`2001. Effect of ‘pH’ on the rate of asparagine
`deamidation in polymeric formulations: ‘pH’-rate
`profile. J Pharm Sci 90:141–156.
`22. Brennan TV, Clarke S. 1995. Deamidation and
`isoaspartate formation in model synthetic peptides:
`The effects of sequence and solution environment.
`In: Aswad D, editor. Deamidation and isoaspartate
`formation in peptides and proteins. Boca Raton, FL:
`CRC Press.
`
`23. Li R D’Souza AJ, Laird BB, Schowen RL, Borchardt
`RT, Topp EM. 2000. Effects of solution polarity and
`viscosity on peptide deamidation. J Peptide Res
`56:326–334.
`24. Perczel A, Hollo´si M, Foxman BM, Fasman GD.
`1991. Conformational analysis of pseudo cyclic
`hexapeptides based on quantitative circular dichro-
`ism (CD), NOE and X-ray data. The pure CD
`spectra of type I and type II beta-turns. J Am Chem
`Soc 112:9772–9784.
`25. Perczel A, Fasman GD. 1992. Quantitative analysis
`of cyclic b-turn models. Protein Sci 1:378–
`395.
`26. Sharma S, Hammen PK, Leung A, Georges G,
`Hestenberg W, Klevit RE, Waygood EB. 1993.
`Deamidation of HPr, a phosphocarrier protein of
`the phosphoenolpyruvate: Sugar phosphotransfer-
`ase system, involves asparagine 38 (HPr-1) and
`asparagine 12 (HPr-2) in isoaspartyl acid forma-
`tion. J Biol Chem 268:17695–17704.
`27. Brennan TV, Anderson JW, Jia Z, Waygood EB,
`Clarke S. 1993. Repair of spontaneously deami-
`dated HPr phosphocarrier protein catalyzed by the
`L-isoaspartate (D-aspartate) O-methyltransferase.
`J Biol Chem 269:25486–25595.
`28. Sibanda BL, Thornton JM. 1993. Accommodating
`sequence changes in beta-hairpins in proteins. J
`Mol Biol 229:428–447.
`29. Wilmot CM, Thornton JM. 1988. Analysis and
`prediction of the different types of beta-turn in
`proteins. J Mol Biol 203:221–232.
`30. De Paz RA, Barnett CC, Dale DA, Carpenter JF,
`Gaertner AL, Randolph TW. 2000. The excluding
`effects of sucrose on a protein chemical degradation
`pathway: Methionine oxidation in subtilisin. Arch
`Biochem Biophys 384:123–132.
`
`JOURNAL OF PHARMACEUTICAL SCIENCES, VOL. 90, NO. 12, DECEMBER 2001
`
`Novo Nordisk Ex. 2044, P. 8
`Mylan Institutional v. Novo Nordisk
`IPR2020-00324
`
`

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