throbber
International Journal of Pharmaceutics 203 (2000) 115–125
`
`www.elsevier.com:locate:ijpharm
`
`The degradation pathways of glucagon in acidic solutions
`
`Anjali B. Joshia, Elena Rusb, Lee E. Kirscha,*
`a Di6ision of Pharmaceutics, College of Pharmacy, The Uni6ersity of Iowa, Iowa City, IA 52242, USA
`b Molecular Analysis Facility, The Uni6ersity of Iowa, Iowa City, IA 52242, USA
`
`Received 2 March 2000; received in revised form 28 April 2000; accepted 1 May 2000
`
`Abstract
`
`Objecti6e: Glucagon is a 29 amino acid peptide hormone that exhibits degradation via both chemical and physical
`pathways. The objective of the studies reported herein was to identify the degradation products and scheme for
`glucagon hydrolysis in acidic solutions. Methods: Solutions of glucagon in 0.01 N HCl (pH 2.5) were degraded at
`60°C for 70 h. One isocratic and two gradient RP-HPLC methods were developed to separate the degradation
`products. Structure elucidation of the separated peaks was achieved using amino acid sequencing, amino acid
`analysis, and mass spectrometry. Degradation was carried out in the pH range 1.5–5 to check for changes in
`degradation scheme with pH. Authentic samples of degradation products were degraded under similar acidic
`conditions to confirm precursor successor relationships in the degradation scheme. Results: Sixteen major degradation
`products were isolated and identified. The major pathways of degradation were found to be aspartic acid cleavage at
`positions 9, 15, and 21 and glutaminyl deamidation at positions 3, 20, and 24. Cleavage occurred on both sides of
`Asp-15 but only on the C-terminal side of Asp-9 and Asp-21. Deamidation of the Asn residue at position 28 was not
`detected. © 2000 Elsevier Science B.V. All rights reserved.
`
`Keywords: Degradation pathways; Deamidation; Glucagon; Glutaminyl; Hydrolysis; Peptide cleavage
`
`1. Introduction
`
`Glucagon is a polypeptide hormone that is used
`for the emergency treatment of insulin induced,
`sulphonylurea induced, and spontaneous hypo-
`glycemia. It has the ability to raise blood glucose
`concentrations by increasing hepatic glycogenoly-
`sis through activation of liver phosphorylase and
`to stimulate insulin secretion by direct action on
`pancreatic b-cells. Recovery of consciousness is
`
`* Corresponding author. Fax: (cid:27)1-319-3359349.
`E-mail address: lee-kirsch@uiowa.edu (L.E. Kirsch).
`
`usually achieved within 15–30 min of the injec-
`tion whereupon oral glucose may be used to
`continue treatment (Marks, 1983). Glucagon also
`has the ability to reduce smooth muscle tone and
`motility. It is used as a diagnostic aid in radio-
`logic procedures of the gastrointestinal tract that
`require a diminished tone and motility of the
`organ under study (Diamant and Picazo, 1983).
`Glucagon for injection is a mixture of glucagon
`hydrochloride with one or more suitable, dry dilu-
`ents supplied in single-dose or multiple-dose con-
`tainers. The pH of the reconstituted solution is
`between 2.5 and 3.0 (US Pharmacopeia 24, 1999).
`
`0378-5173:00:$ - see front matter © 2000 Elsevier Science B.V. All rights reserved.
`PII: S 0 3 7 8 - 5 1 7 3 ( 0 0 ) 0 0 4 3 8 - 5
`
`Page 1
`
`NPS EX. 2048
`CFAD v. NPS
`IPR2015-00990
`
`

`
`116
`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`Thus the stability of glucagon in acidic aqueous
`solutions is relevant to drug product quality.
`Structurally, glucagon contains 29 amino acids
`(Fig. 1). In the crystalline state, a completely
`helical conformation was proposed for glucagon
`(Sasaki et al., 1975). However, optical rotatory
`dispersion (Gratzer et al., 1968) and circular
`dichroism studies (Srere and Brooks, 1969) of
`glucagon in dilute aqueous solutions revealed a
`predominantly random coil conformation with at
`most 15% a-helix at the C-terminal end. The
`molecule has an isoelectric point of (cid:2) 7 and
`exhibits very low solubility (B0.1 mg:ml) in the
`approximate pH range of 4–8. It is readily soluble
`(\10 mg:ml) at pH values less than 3 or greater
`than 9 (Bromer, 1983). It is known to self-associ-
`ate at high concentrations and forms aggregates
`and gels at mild temperatures in acidic and basic
`
`Fig. 1. Sequences of glucagon and its major degradation
`products in acidic aqueous solutions. The sequence of each
`degradation fragment is denoted using the one-letter abbrevia-
`tions for the amino acids and by the numbers of the first and
`last amino acid residues. Deamidated fragments are denoted
`by parentheses followed by ‘dn’ where ‘n’ is the number of the
`glutamine residue which deamidates.
`
`solutions. The aggregation is promoted by salt,
`increased pH (within the acid range of solubility),
`agitation, and increased temperature up to 30°C
`(Beaven et al., 1969). Aggregation is accompanied
`by an increase in secondary structure in the form
`of anti-parallel b-sheets (Gratzer et al., 1968).
`The degradation pathways of glucagon have
`not been reported. In one study, glucagon degra-
`dation in 0.03 N HCl at 105°C was found to
`release three aspartic acid residues
`in 13 h
`(Schultz, 1967). The possibility of iso-aspartyl for-
`mation at the position of Asp-9 at neutral pH has
`also been suggested (Ota et al., 1987). The objec-
`tive of the studies presented herein was to identify
`the degradation products
`and scheme
`for
`glucagon hydrolysis in acidic solutions.
`
`2. Materials and methods
`
`2.1. Materials
`
`Purified porcine glucagon was obtained from
`Lilly Research Laboratories (Indianapolis, IN).
`Sodium dihydrogen phosphate,
`o-phosphoric
`acid, acetic acid, formic acid, hydrochloric acid,
`trifluoroacetic acid, potassium chloride, sodium
`formate, and sodium acetate were from Fisher
`(Springfield, NJ). All chemicals were of reagent
`grade and used as received. Solvents used for
`chromatography were HPLC grade. Cellulose es-
`ter dialysis membranes were from Spectrum
`(Houston, TX). Degradation product fragments
`were synthesized manually by SynPep Corpora-
`tion (Dublin, CA).
`
`2.2. Degradation of glucagon and glucagon
`fragments in acidic pH range
`
`(0.1–1 mg:ml)
`Aqueous glucagon solutions
`were degraded at 60°C in the pH range 1.5–5
`using dilute HCl, phosphate, formate, or acetate
`buffers. Degradation was allowed to proceed to
`two to three half-lives. Aliquots were removed
`from reaction mixtures at various time intervals
`and stored at 4°C until analyzed. Authentic
`degradation products were subject to hydrolysis at
`pH 1.5 and 2.5 under conditions identical to those
`used for glucagon.
`
`Page 2
`
`

`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`117
`
`2.3. Separation of degradation products
`
`chromatography
`liquid
`performance
`High
`(HPLC) analyses were performed using a Shi-
`madzu RP-HPLC system consisting of an SCL-
`10AVP system controller, LC-10ATVP pumps,
`SIL-10ADVP auto-injector, SPD-10AVP UV-VIS
`detector, and a CTO-10ASVP column oven.
`Chromatograms were integrated and data stored
`using Class VP Chromatography Data System
`software (Version 4.2).
`products was
`Separation
`of
`degradation
`achieved on a Lichrospher RP-18, 4.6(cid:29)250 mm,
`5m-column using isocratic elution (Method A).
`The mobile phase contained 29:71 acetonitrile:
`buffer
`(buffer(cid:30)0.1 M sodium phosphate
`monobasic and 0.002 M cysteine free base, ad-
`justed to pH 2.6 with 85% phosphoric acid). Flow
`rate was 1.0 ml:min, detection was at 214 nm, and
`column temperature was 35°C. Degradation prod-
`ucts
`that were unresolved with the isocratic
`method were separated by gradient elution using a
`Zorbax 300SB C8, 4.6(cid:29)250 mm, 5m column. The
`mobile phase contained methanol and a 0.024%
`solution of trifluoroacetic acid in water. The gra-
`dient used was 5–35% methanol
`in 30 min
`(Method B). Flow rate was 1.0 ml:min, detection
`was at 214 nm, and column temperature was
`25°C. Another gradient method (Method C) was
`developed in an attempt to separate additional
`degradation peaks using the same column and
`mobile phase as the previous gradient method.
`The gradient was 5–50% methanol
`in 90 min.
`Flow rate was 1.0 ml:min, detection was at 214
`nm, and column temperature was 25°C.
`Degradation product peaks were collected from
`repeated 200 ml reaction mixture samples using a
`FRC 10A fraction collector (Shimadzu). Seven
`peaks were collected with the isocratic method
`(Method A), five more peaks with the first gradi-
`ent method (Method B), and an additional three
`peaks with the second gradient method (Method
`C). The collection procedure was validated by
`re-injecting peak fractions back on the column to
`check peak integrity. Individual peaks were col-
`lected in flat top 1.5 ml polypropylene microcen-
`trifuge tubes (Fisher Scientific) and stored in the
`refrigerator at 4°C until analysis. Since analysis of
`
`the fractions was typically done 4–7 days after
`fraction collection, the stability of the fractions at
`4°C was evaluated. No significant degradation of
`peptide fractions was observed up to 18 days at
`4°C.
`
`2.4. Degradation product identification
`
`Structure elucidation of degradation products
`was done using either amino acid sequencing,
`amino acid analysis, or mass spectrometry. Most
`of the fractions were identified using more than
`one method.
`Amino acid analysis was performed by hy-
`drolyzing an aliquot of each fraction using 100 ml
`of 6 N HCl at 110°C for 24 h. When hydrolysis
`was completed, the vials were placed in a Speed-
`Vac® (Model SC210A) where all the residual liq-
`uid was
`removed by
`evaporation. Sample
`hydrolysates were then re-dissolved in sodium cit-
`rate buffer diluent. A Beckman 6300 high-perfor-
`mance ion-exchange analyzer was used to analyze
`each sample. Separation was done using a 12 cm
`hydrolysate column and a three-step temperature
`program in combination with three sodium citrate
`buffers to separate amino acids in various charged
`states. Standards consisting of a mixture of all
`amino acids plus an internal standard were also
`hydrolyzed. The system used ninhydrin to react
`with the amino acid giving a color reaction. The
`intensity of the color was proportional to the
`concentration of amino acid.
`For sequencing, fraction aliquots were concen-
`trated in a SpeedVac® and passed through a
`Prosorb sample preparation cartridge (Perkin
`Elmer). The adsorbent filter in these cartridges
`draws sample solution through a membrane by
`capillary action. The membrane
`immobilizes
`proteins and peptides while buffer components
`that could potentially interfere with sequencing
`pass through. The process not only desalts but
`also concentrates the sample. The membrane
`holding the peptide was directly loaded on the
`sequencer. Sequencing was done using an Applied
`Biosystems 492 automatic sequencer with an on-
`line PTH analyzer. A standard consisting of a
`mixture of all amino acids was first injected to
`obtain reference retention times for each amino
`
`Page 3
`
`

`
`118
`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`acid. One fraction was assigned more than one
`sequence (one major and one minor) due to the
`presence of a mixture of peptides.
`Degradation product identity was also verified
`using Fast Atom Bombardment Mass Spectrome-
`try (FABMS). The instrument used was a Hewlett
`Packard 1100 LCMSD. Organic solvents present
`in peak fractions collected with the isocratic
`method (Method A) were removed by evapora-
`tion under nitrogen. Sample peak fractions were
`then desalted using cellulose ester dialysis mem-
`branes (MWCO(cid:30)500). De-ionized water was
`used as the dialysis medium and was changed
`twice daily. The extent of dialysis was monitored
`by measuring the conductivity of the dialysate on
`an Orion Model 160 conductivity meter. Dialysis
`was stopped when the conductivity measurement
`was close to that of pure de-ionized water. The
`dialyzed fractions were then freeze dried (Virtis
`Advantage lyophilizer) and analyzed by FABMS.
`Peak fractions collected using the gradient meth-
`ods (Methods B and C) were simply concentrated
`in a SpeedVac® before FABMS analysis since
`these fractions contained only water and volatile
`solvents.
`To verify the sequencing results, two peaks
`obtained by isocratic elution (Method A) were
`analyzed by FABMS. The molecular ion masses
`obtained corresponded to the fragments obtained
`by sequencing. Hence, additional FABMS analy-
`ses were not performed on fractions that were
`sequenced. However, FABMS was used to
`confirm every sequence obtained with amino acid
`analysis alone.
`
`3. Results
`
`3.1. Chromatographic separation of degradation
`products
`
`degraded
`of
`separation
`Chromatographic
`glucagon using isocratic elution (Method A) gave
`seven major degradation peaks (Peaks I-VII, Fig.
`2). Sequencing of these products revealed that
`they were peptide fragments containing the C-ter-
`minus of glucagon. A gradient elution method
`(Method B) was developed to separate the corre-
`
`sponding fragments containing the more polar
`amino acids of the N-terminus end. An additional
`five peaks were obtained (Peaks VIII–XII, Fig.
`3). A second gradient method (Method C) was
`developed to check for additional degradation
`products. This method involved a slow gradient
`(90 min) from 5 to 50% methanol and yielded
`several peaks. The peak at retention time of 5.5
`min was found to be the same as peak XII
`obtained with Method B (Fig. 3). The peaks
`eluting at 35.3, 59.3 min, and 69.3 min were found
`to be the same as peak I, peak III, and peak V
`respectively, obtained by Method A (Fig. 2).
`These peak assignments were made by co-inject-
`ing peak fractions or authentic samples of peaks
`XII, I, III, and V and comparing retention times.
`The peak eluting at 72.1 min was glucagon. Thus,
`three new peaks, at 15.3, 20.2, and 26.6 min were
`obtained by Method C (peaks XIII, XIV, and XV
`in Fig. 4).
`An overlay of sample chromatograms obtained
`from degradation reactions conducted over the
`pH range 2.5–5 (Fig. 5) demonstrated that the
`retention times of the major degradation product
`peaks were similar at pH 2.5 and 3.5. At higher
`pH values (4.5 and 5), the chromatograms were
`somewhat similar but separation quality was com-
`promised by glucagon aggregation.
`
`3.2. Degradation product identities
`
`Of the seven peaks obtained by Method A (Fig.
`2), peaks I, III, and V were composed of peptide
`fragments 22–29, 16–29, and 10–29 respectively,
`indicating that peptide cleavage had occurred on
`the C-terminal side of Asp-21, Asp-15, and Asp-9.
`Peaks II, IV, and VI eluted immediately after
`peaks I, III, and V (Fig. 2) and were found to be
`the deamidated forms of fragments 22–29 (peak
`I), 16–29 (peak III), and 10–29 (peak V) wherein
`Gln-24 converted to Glu-24. Peak VII was the
`deamidated form of glucagon with Glu at position
`24. A small peptide fragment starting at Tyr-10
`co-eluted with deamidated 16–29 in peak IV.
`Both these peptides could be detected through
`separate PTH-AA signals. However, both frag-
`ments have Asp as their sixth residue, hence the
`chromatogram for the sixth cycle of sequencing
`
`Page 4
`
`

`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`119
`
`Fig. 2. Separation of degradation products of glucagon by isocratic elution using RP-HPLC Method A (column was a Lichrospher
`RP-18, mobile phase was 29:71 acetonitrile:buffer, buffer(cid:30)0.1 sodium phosphate monobasic and 0.002 M cysteine free base,
`adjusted to the pH 2.6 with 85% phosphoric acid).
`
`Page 5
`
`

`
`120
`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`Fig. 3. Separation of degradation products containing the more polar amino acid residues of the N-terminal end using RP-HPLC
`Method B (column was a Zorbax SB300 C8, mobile phase was methanol and 0.024% TFA in water, gradient was 5–35% methanol
`in 30 min).
`
`Page 6
`
`

`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`121
`
`Fig. 4. Separation of additional degradation products containing the more polar amino acid residues of the N-terminal end using
`RP-HPLC Method C (column was a Zorbax SB300 C8, mobile phase was methanol and 0.024% TFA in water, gradient was 5–50%
`methanol in 90 min).
`
`Page 7
`
`

`
`122
`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`Fig. 5. Overlaid chromatograms of glucagon degraded in the pH range 2.5–5.0 using Method C.
`
`Page 8
`
`

`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`123
`
`gave only one peak for Asp. This made it difficult
`to determine whether the signal came just from
`the major fragment (deamidated 16–29) or from
`both fragments. A comparison of the combined
`picomole yields for the two fragments (as the
`sequencing progressed) indicated that the Asp pi-
`comole yield was not large enough to include two
`signals. Hence the sequence for the co-eluting
`peak was assigned as 10–14. This suggested the
`possibility that the Asp-15 residue cleaved on
`both its C-terminal and N-terminal sides.
`Peaks VIII through XII (Fig. 3) were obtained
`by gradient elution (Method B) and contained the
`peptide
`fragments
`containing the N-terminal
`residues that were not recovered by isocratic elu-
`tion. Peaks VIII and X were fragments 1–9 and
`10–21, respectively. This confirmed the result that
`Asp-9 and Asp-21 did not cleave on the N-termi-
`nal side. Peak XII was fragment 10–15, which
`further confirmed cleavage on both sides of the
`Asp-15 residue. Peaks IX and XI were the deami-
`dated products of peaks VIII and X respectively,
`which indicated that the glutamine residues in
`positions 3 and 21 were susceptible to deamida-
`tion as well as Gln-24 (peaks II, IV, VI, and VII).
`The three peaks obtained with Method C (Fig.
`4) were analyzed using amino acid analysis and
`confirmed by FABMS. Peak XIII was found to be
`fragment 1–21, peak XIV was fragment 1–15,
`and peak XV was fragment 1–14. This further
`confirmed that peptide cleavage occurred on both
`sides of Asp-15 and only on the C-terminal side of
`Asp-21. The peaks that
`immediately followed
`peaks XIII, XIV, and XV in Fig. 4 were not
`analyzed. These are expected to be the deami-
`dated forms of the fragments 1–21 (peak XIII),
`1–15 (peak XIV), and 1–14 (peak XV), respec-
`tively. The results of structure elucidation of the
`degradation products are summarized in Fig. 1.
`
`3.3. Verification of precursor-successor
`relationships
`
`The synthesized peptide fragments 22–29 and
`16–29 corresponding to peaks I and III in Fig. 2
`were subjected to degradation under acidic condi-
`tions. The resultant degradation chromatograms
`indicated that these fragments were subject to the
`
`expected peptide cleavage and deamidation path-
`ways. Degradation of fragment 22–29 resulted in
`the appearance of a peak with retention time
`corresponding to deamidated 22–29. Degradation
`of the peptide fragment 16–29 gave rise to HPLC
`peaks corresponding to 22–29 and the deami-
`dated forms of 16–29 and 22–29. The peaks
`corresponding to fragment 10–29 (peak V, Fig. 2)
`and fragment 1–15 (peak XIV, Fig. 3) were col-
`lected and also degraded under acidic conditions.
`Fragment 10–29 gave peaks corresponding to
`22–29 and 16–29, and the deamidated forms of
`10–29. Fragment 1–15 gave peaks corresponding
`to 1–14 and 1–9.
`
`4. Discussion
`
`Sixteen degradation products from thermally
`stressed, acidic solutions of glucagon were iso-
`lated and identified. The major pathways of
`glucagon degradation were glutaminyl deamida-
`tion and aspartyl peptide cleavage. The former
`was demonstrated to occur at residues 3, 20, and
`24 and the latter at residues 9, 15, and 21. Cleav-
`age at Asp-9 and Asp-21 occurred on the C-termi-
`nal side of the amino acid whereas cleavage at
`Asp-15 occurred on both N- and C-terminal sides.
`The proposed degradation scheme is illustrated in
`Fig. 6.
`Deamidation is a common protein degradation
`pathway that involves the loss of ammonia from
`the side chain amides of asparagine or glutamine
`to form the corresponding side chain carboxylic
`acid residues: aspartic or glutamic acid (Manning
`et al., 1989). Numerous examples of asparaginyl
`and, to a lesser extent, glutaminyl deamidation
`have been reported (Robinson and Rudd, 1974;
`Geiger and Clarke, 1987; Patel and Borchardt,
`1990a; Windisch et al., 1997). Some of the effects
`of primary, secondary, tertiary, and quaternary
`structure on deamidation kinetics have been de-
`scribed (Kossiakoff, 1988; Wearne and Creighton,
`1989; Patel and Borchardt, 1990b; Wright, 1991;
`Darrington and Anderson,
`1994; Xie
`and
`Schowen, 1999). In general, the presence of neigh-
`boring glycine (specifically at the N(cid:27)1 residue),
`serine
`or
`threonine
`facilitates
`deamidation
`
`Page 9
`
`

`
`124
`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`(Wright, 1991) whereas the presence of significant
`secondary structure
`(Wearne and Creighton,
`1989) or the lack of solvent access to the potential
`deamidating residue due to tertiary structure may
`tend to slow reaction rates. Moreover, asparaginyl
`deamidation is reported to be more facile than
`glutaminyl deamidation (Robinson and Rudd,
`1974). It has been suggested that this kinetic
`difference may be due to the greater length of the
`glutamine side chain which precludes stabilization
`of a side chain carbonyl oxyanion transition state
`by hydrogen bonding with the neighboring pep-
`tide nitrogen (Wright and Robinson, 1982).
`In glucagon, the lack of asparaginyl deamida-
`tion at residue 28 in the presence of glutaminyl
`deamidation at residues 3, 20 and 24 was a some-
`what unexpected observation. Although the sole
`asparagine residue in glucagon is located in a
`region of the peptide associated with some helical
`
`Fig. 6. Proposed degradation scheme of glucagon in acidic
`aqueous solutions. Deamidation and aspartyl cleavage path-
`ways are represented by dashed and solid arrows respectively.
`Only observed, but not hypothesized, degradation products
`are depicted. * Starting materials; ** has not been chromato-
`graphically separated.
`
`structure, this structure has been shown to be lost
`as the temperature is raised in the range 22–50°C
`(Yi et al., 1992). Since the current study was
`conducted at 60°C, secondary structure effects
`were unlikely.
`Most of the studies that support the greater
`susceptibility of asparaginyl residues to deamida-
`tion were conducted at pH conditions approach-
`ing neutrality (Robinson and Rudd, 1974; Geiger
`and Clarke, 1987; Patel and Borchardt, 1990a).
`The relative reactivity of glutamine and as-
`paragine and the effects of primary, secondary,
`and tertiary structure on glutaminyl deamidation
`in acidic conditions have not been reported and
`are the focus of our ongoing studies.
`Preferential peptide cleavage at aspartic acid
`residues in the presence of acid is a well-known
`hydrolysis pathway that was historically used to
`fragment proteins
`for
`compositional analysis
`(Light, 1967). Cleavage can occur between aspar-
`tic acid and either the N(cid:28)1 or N(cid:27)1 neighboring
`residue. The mechanism is believed to involve
`intramolecular catalysis by the side chain car-
`boxylic acid and formation of a six- or five-mem-
`bered ring intermediate (Inglis, 1983). Peptide
`cleavage was observed at all three of the aspartic
`acid residues in glucagon. Cleavage on both sides
`of Asp-15 was observed. The relative rates of
`peptide cleavage at the three residues are expected
`to vary based on sequence effects. Future work
`will evaluate the relative reactivity of the aspartic
`acid residues.
`
`References
`
`Beaven, G.H., Gratzer, W.B., Davies, H.G., 1969. Formation
`and structure of gels and fibrils from glucagon. Eur. J.
`Biochem. 11, 37–42.
`Bromer, W.W., 1983. Chemical characteristics of glucagon. In:
`Lefe`bvre, P.J. (Ed.), Handbook of Experimental Pharma-
`cology, Vol 66:1. Springer-Verlag, Berlin, pp. 1–22.
`Darrington, R.T., Anderson, B.D., 1994. The role of in-
`tramolecular nucleophilic catalysis and the effects of self-
`association on the deamidation of human insulin at low
`pH. Pharm. Res. 11, 784–793.
`Diamant, B., Picazo, J., 1983. Spasmolytic action and clinical
`use of glucagon. In: Lefe`bvre, P.J. (Ed.), Handb. Exp.
`Pharm, vol. 66:2. Springer-Verlag, Berlin, pp. 611–643.
`
`Page 10
`
`

`
`A.B. Joshi et al. :International Journal of Pharmaceutics 203 (2000) 115–125
`
`125
`
`Geiger, T., Clarke, S., 1987. Deamidation, isomerization, and
`racemization at asparaginyl and aspartyl residues in pep-
`tides. Succinimide-linked reactions
`that contribute to
`protein degradation. J. Biol. Chem. 262, 785–794.
`Gratzer, W.B., Beaven, G.H., Rattle, H.W.E., Bradbury,
`E.M., 1968. A conformational study of glucagon. Eur. J.
`Biochem. 3, 276–283.
`Inglis, A.S., 1983. Cleavage at aspartic acid. Methods Enzy-
`mol. 91, 324–332.
`Kossiakoff, A.A., 1988. Tertiary structure is a principal deter-
`minant to protein deamidation. Science 240, 191–194.
`Light, A., 1967. Partial acid hydrolysis. Methods Enzymol. 11,
`417–420.
`Manning, M.C., Patel, K., Borchardt, R.T., 1989. Stability of
`protein pharmaceuticals. Pharm. Res. 6, 903–918.
`Marks, V., 1983. Glucagon in the diagnosis and treatment of
`hypoglycemia. In: Lefe`bvre, P.J. (Ed.), Handbook of Ex-
`perimental Pharmacology, vol. 66:2. Springer-Verlag,
`Berlin, pp. 645–672.
`Ota, I.M., Ding, L., Clarke, S., 1987. Methylation at specific
`altered aspartyl and asparaginyl residues in glucagon by
`the erythrocyte protein carboxyl methyltransferase. J. Biol.
`Chem. 262, 8522–8531.
`Patel, K.P., Borchardt, R.T., 1990. Chemical pathways of
`peptide degradation. II Kinetics of deamidation of an
`asparaginyl residue in a model hexapeptide. Pharm. Res. 7,
`703–711.
`Patel, K., Borchardt, R.T., 1990. Chemical pathways of pep-
`tide degradation. III Effect of primary sequence on the
`pathways of deamidation of asparaginyl
`residues
`in
`hexapeptides. Pharm. Res. 7, 787–793.
`Robinson, A.B., Rudd, C.J., 1974. Deamidation of glutaminyl
`
`and asparaginyl residues in peptides and proteins. Curr.
`Top. Cell Regul. 8, 247–295.
`Sasaki, K., Dockerill, S., Adamiak, D.A., Tickle, I.J., Blun-
`dell, T., 1975. X-ray analysis of glucagon and its relation-
`ship to receptor binding. Nature 257 (5529), 751–757.
`Schultz, J., 1967. Cleavage at aspartic acid. Methods Enzymol.
`11, 255–263.
`Srere, P.A., Brooks, G.C., 1969. The circular dichroism of
`glucagon solutions. Arch. Biochem. Biophys. 220, 708–
`710.
`US Pharmacopeia 24, 1999. US Pharmacopeial Convention,
`Rockville, MD, pp. 777.
`Wearne, S.J., Creighton, T.E., 1989. Effect of protein confor-
`mation on rate of deamidation: ribonuclease-A. Proteins 5,
`8–12.
`Windisch, V., DeLuccia, F., Duhau, L., Herman, F., Mencel,
`J.J., Tang, S.Y., Vuilhorgne, M., 1997. Degradation path-
`ways of salmon calcitonin in aqueous solution. J. Pharm.
`Sci. 86, 359–364.
`Wright, H.T., 1991. Sequence and structure determinants of
`the non-enzymatic deamidation of asparagine and glu-
`tamine in proteins. Protein Eng. 4, 283–291.
`Wright, H.T., Robinson, A.B., 1982. Cryptic amidase active
`sites catalyze deamidation in proteins. In: Kaplan, N.O.,
`Robinson, A.B. (Eds.), From Cyclotrons to Cytochromes.
`Academic Press, New York, pp. 727–743.
`Xie, M., Schowen, R., 1999. Secondary structure and protein
`deamidation. J. Pharm. Sci. 88, 8–13.
`Yi, G.S., Choi, B.S., Kim, H., 1992. The conformation of
`glucagon in dilute aqueous solutions as studied by 1H
`NMR. Biochem. Int. 28, 519–524.
`
`.
`
`Page 11

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