throbber
Table IX-Extent of Complex Formation between 3,4-
`Dimethylphenol and Ethyl Myristate Determined by
`Partitioning Study
`
`Total Concentration
`of 3,4-Dimethylphenol,
`X 102 M
`
`Partition
`Coefficient
`
`[DMPCO"!Pl
`[DMPol
`
`0
`1.53
`3.07
`6.14
`9.21
`15.03
`30.00
`
`1.37
`1.73
`1.96
`2.61
`3.31
`4.57
`7.94
`
`0.27
`0.43
`0.91
`1.43
`2.35
`4.83
`
`for various cosolvent systems, and each showed a straight-line relation(cid:173)
`ship between the -two parameters (Fig. 4). The stability constants, K 1,1
`and K 1,2, for each ester were calculated from Fig. 4 and are shown in Table
`VII. It is evident from these results that 4-hexylresorcinol forms not only
`1:1 but also 1:2 complexes with esters in hexane. The stability constant
`values obtained for ethyl myristate were somewhat higher than those for
`the other esters. This result is probably due to the fact that ethyl my(cid:173)
`ristate has a larger hydrocarbon chain, which results in a better interac(cid:173)
`tion with the hydrophobic portion of phenols.
`To ascertain whether the formation of 1:2 complexes is due to the in(cid:173)
`volvement of the two hydroxy groups of 4-hexylresorcinol, the parti(cid:173)
`tioning study was repeated with 3,4-dimethylphenol. The data obtained
`from this study are shown in Tables VIII-X. A plot of [DMPcompJ/
`
`Table X-Extent of Complex Formation between 3,4-
`Dimethylphenol and Ethyl Pivalate Determined by Partitioning
`Study
`
`Total Concentration
`of 3,4-Dimethylphenol,
`X 102 M
`
`Partition
`Coefficient
`
`[DMPcomp]
`[DMPo}
`
`0
`3.82
`7.68
`15.36
`23.04
`38.40
`
`1.37
`1.78
`2.33
`3.33
`4.32
`6.63
`
`0.30
`0.70
`1.44
`2.17
`3.86
`
`•
`
`•
`.
`
`• e
`
`e
`
`•
`
`~16
`~
`0
`~ 12
`lU ::::
`Cl. 8
`E
`0 " ~ 4
`a
`
`60
`50
`40
`30
`20
`1 0
`0
`CONCENTRATION OF TOTAL ESTER,
`[Erl. IN HEXANE, X 102 M
`Figure 5-Plot of [DMPcompl I [DMPtreJ [ET] as a function of the total
`ester concentration, [ET], in hexane; [DMPcompJ and [DMPtreel rep(cid:173)
`resent the concentrations of complexed and free forms of 3,4-dimeth(cid:173)
`ylphenol, respectively. ([DMPcomp] = [DMPTJ - [DMPol). Key: ■ ,
`ethyl acetate; e, ethyl myristate; and .._, ethyl pivalate.
`
`[DMP01[Er] versus the total ester concentration is shown in Fig. 5;
`[DMPcomp] is the concentration of 3,4-dimethylphenol in the complex
`form, and [DMP0] is the concentration of the free form.
`As seen in Fig. 5, the monohydroxy compound forms only a 1:1 complex
`with the esters. The stability constants calculated from Fig. 5 are given
`in Table VII. The results of this study substantiate the conclusion that
`the diffusion of 4-hexylresorcinol through ethylene-vinyl acetate co(cid:173)
`polymers involved the formation of 1:1 and 1:2 complexes between the
`drug and the vinyl acetate portion of the copolymers.
`
`REFERENCES
`
`(1) E. Akaho, Ph.D. dissertation, University of Kentucky, Lexington,
`Ky. 1979.
`(2) H.B. Kostenbauder and T. Higuchi, J. Am. Pharm. Assoc. Sci. Ed.,
`45, 518 (1956).
`(3) T. Higuchi, J. H. Richards, S.S. Dacis, A. Kamada, J.P. How, M.
`Nakano, N. I. Nakano, and I. H. Pitman, J . Pharm. Sci., 58, 661
`(1969) .
`(4) H. Fung and T. Higuchi, ibid., 60, 1782 (1971).
`
`High-Performance Liquid Chromatographic Analysis of
`Chemical Stability of 5-Aza-2' -deoxycytidine
`
`KUN-TSAN LIN*x, RICHARD L. MOMPARLER**, and GEORGES E. RIVARD*
`Received December 17, 1979, from the *Centre de Recherche Pediatrique, H6pital Sainte-Justine, and the I Departement de Pharmacologie,
`Universite de Montreal, Montreal, Quebec , Canada H3T !CS.
`Accepted for publication March 31, 1981.
`
`Abstract □ The chemical stability of 5-au-2' -deoxycytidine (I) in acidic,
`neutral, and alkaline solutions was analyzed by high-performance liquid
`chromatography. In alkaline solution, I underwent rapid reversible de(cid:173)
`composition to N-(formylamidino)-N'-P-D-2-deoxyribofuranosylurea
`(II), which decomposed irreversibly to form 1-P-n-2'-deoxyribofurano(cid:173)
`syl-3-guanylurea (III). The pseudo-first-order rate constants for this
`reaction were determined. The decomposition of I in alkaline solution
`was identical to that reported previously for the related analog, 5-aza(cid:173)
`cytidine. However, in neutral solution (or water), there was a marked
`difference in the decomposition of I and 5-azacytidine. The same de(cid:173)
`composition products were formed from 5-azacytidine in neutral solution
`
`as in alkaline solution. However, in neutral solution, I decomposed to II
`and three unknown compounds that were chromophoric at 254 nm.
`Compound I was most stable when stored in neutral solution at low
`temperature.
`
`Keyphrases Cl 5-Aza-2'-deoxycytidine-analysis of chemical stability
`using high-performance liquid chromatography □ Antileukemic
`agents-5-aza-2'-deoxycytidine, analysis of chemical stability using
`high-performance liquid chromatography □ High-performance liquid
`chromatography-analysis of chemical stability of 5-aza-2'-deoxycyti(cid:173)
`dine
`
`3). This antimetabolite is related to 5-azacytidine, an agent
`5-Aza-2'-deoxycytidine (I), a nucleoside antimetabolite,
`is a very active antileukemic agent in mice (1, 2) and a
`currently used in the clinical treatment of acute leukemia
`potent cytotoxic agent against neoplastic cells in vitro (2,
`(4).
`CELGENE 2021
`APOTEX v. CELGENE
`IPR2023-00512
`
`0022-354918111100-1228$01.0010
`© 1981, American Pharmaceutical Association
`
`1228 / Journal of Pharmaceutical Sciences
`Vol. 70, No. 11, November 1981
`
`

`

`NH2
`
`NJ_N
`
`NH~ 0
`J._ n
`N7 NHCH
`
`HO~
`
`HO~H
`
`OH H
`I
`
`OH H
`II
`
`NH2
`
`N~NH2
`
`HO~H
`
`OH H
`m
`
`Scheme I
`
`BACKGROUND
`
`One major problem encountered in the clinical formulation of 5-aza(cid:173)
`cytidine is its chemical instability, leading to solutions of decreasing
`potency on storage. The chemical stability of 5-azacytidine was first
`studied by Pithova et al. (5), who demonstrated that in alkaline solution
`the triazine ring of 5-azacytidine opens and loses a formyl group to form
`1-/1-D-ribofuranosyl-3-guanylurea. These workers proposed that the
`intermediate compound in this reaction was N-(formylamidino)-N'(cid:173)
`/1-D-ribofuranosylurea, but they were unable to isolate it using paper
`chromatography because of its chemical instability. Beisler (6), using
`high-performance liquid chromatography (HPLC), isolated and identi(cid:173)
`fied this intermediate compound and showed that it could be partially
`converted back to 5-azacytidine.
`The chemical decomposition of I at alkaline pH (Scheme I) presumably
`follows the same reaction steps as described previously for 5-azacytidine.
`In alkaline solutions, I undergoes a reversible hydrolytic reaction to form
`N-(~ormyla~idino)-N'-/1-D-2-deoxyribofuranosylurea (II), which, by
`the 1rrevers1ble loss of the formyl group, forms 1-/1-D-2' -deoxyribofura(cid:173)
`nosyl-3-guanylurea (Ill).
`The present study investigated the decomposition of I in acidic, neu(cid:173)
`tral, and alkaline solutions and at different temperatures to illustrate the
`chemical stability of this compound.
`
`EXPERIMENTAL
`
`5-Aza-2'-deoxycytidine (I) was synthesized1 by modification of an
`earlier method (7).
`Method of Analysis-HPLC2 was performed with a variable-wave(cid:173)
`le~gth detector by monitoring at 220 or 254 nm. Analytical and prepar(cid:173)
`ative work was accomplished with a 300 X 3.9-mm i.d. commercially
`packed octadecylsilane column3, which was eluted at a flow rate of 2
`ml/min with 10 mM potassium phosphate buffer (pH 6.8). For analytical
`and preparative work, 20- and 500-µl injector loops, respectively, were
`used.
`Stock solutions of 10 mM I in water were stored at -70°. Aliquots of
`this solution were thawed quickly and stored at 1-2°. The following
`buffers were used for the pH stability studies: phosphoric acid, pH 2.2;
`phosphoric acid- potassium phosphate, pH 3.2; sodium acetate buffer,
`pH 4.5 and 5.7; potassium phosphate buffer, pH 6.4 arid 7.0; and sodium
`borate, pH 8.5, 9.2, 9.8, and 10.4. Compound II was prepared by addition
`of 20 µl of 0.5 M sodium borate, pH 10.4, to a 1.0-ml solution of 10 mM
`I. The mixture was incubated for 2 min at 24° and then neutralized with
`
`i By Dr. A. Piskala,_Institute of Organic Chemistry and Biochemistry, Czecho(cid:173)
`slovak Academy of Science, 16610 Prague 6, Czechoslovakia.
`2 Model 110 pump and Hitachi variable-wavelength detector Altex Scientific
`Berkeley, Calif.
`'
`'
`3 µBondapak C1s, Waters Associates, Framingham, Mass.
`
`Omin
`
`I O.OlAU
`
`2min
`
`pH 10.4 24 °C '
`
`IOmin
`
`25min
`
`60min
`
`60min
`
`Omin
`
`I O.OIAU
`
`2min
`
`25min
`
`E
`C
`(h
`N
`w·
`CJ z
`Q:) er
`0
`CJ)
`CD
`<(
`
`<(
`
`E
`C
`
`~
`N
`w·
`CJ z
`CD er
`0
`CJ)
`CD
`<(
`
`<(
`
`Figure 1-Decomposition of I at pH 10.4 and 24°. Compound I (1 .0
`mM) was dissolved in 10 mM borate buffer (pH 10.4). At the indicated
`times, an aliquot of the solution was neutralized with phosphate buffer
`and analyzed by HPLC at 254 and 220 nm.
`
`20 µl of 1.0 M KH2P04, and a 500-µl sample was injected onto the column
`for isolation of TI.
`Kinetic Experiments-A series of tubes containing 10 µl of 0.5 M
`stock sodium borate buffer were incubated at 37° for 10 min. Then 100
`µl of isolated II (in 10 mM potassium phosphate, pH 6.8) was added to
`each tube to start the reaction. At timed intervals, an aliquot was with(cid:173)
`drawn, and 20 µI was injected onto the column. The absorbance at 220
`nm was recorded, and the relative concentration of the observed degra(cid:173)
`dation products was characterized by peak heights. Similar experimental
`procedures were followed for the kinetic study on 5-azacytidine4•
`
`RESULTS
`
`The decomposition of I at pH 10.4 and 24° as determined by HPLC
`is shown in Fig. 1. Measurement of the absorbance of the column eluate
`at 254 nm initially showed a single major peak (I) with a tR value of 5.5
`min. With time, a second peak (II) appeared with a tR value of 6.6 min.
`Both peaks showed a gradual decrease in peak size with time. At 220 nm,
`the column eluate initially showed a single major peak (I) with a tR = 5.5
`min. With time, peak II (tR = 6.6 min) and peak III (tR = 3.5 min) became
`evident. At 60 min, most of I decomposed to peak III.
`
`4 Drug SY!)thesis and Chemistry Branch, Division of Cancer Treatment, National
`Cancer Institute, Bethesda, Md.
`
`Journal of Pharmaceutical Sciences I 1229
`Vol. 70, No. 11, November 1981
`
`

`

`Omitt
`
`A
`
`~
`
`pH8 .J
`
`37°C
`
`E
`C:
`;1;
`N w
`() z
`
`<(
`a:i
`a:
`0
`Vl
`IX)
`<(
`
`Omin
`
`II
`
`pH 3.2 0°C
`
`0.5 min
`
`a
`
`e C:
`~
`w·
`u z
`a:i a: g
`
`<(
`
`Cll
`<(
`
`5 min
`
`30 mir.
`
`60min JlJ
`
`0
`Figure 2-Conversion of II to I at pH 8.3, 37° (A), and to unidentified compounds at pH 3.2, 0° (B). Peak II was isolated by HPLC and placed
`in 20 mM potassium phosphate buffer (at final pH 8.3 or 3.2). At the indicated times, aliquots of this solution were analyzed by HPLC.
`
`When II was isolated and placed at pH 8.3 and 37°, I appeared rapidly,
`indicating that II could be converted to I (Fig. 2A). However, when lI was
`placed in an acidic solution (pH 3.2), it did not produce I; three uniden (cid:173)
`tified peaks at 254 nm were observed. The latter reaction was very rapid,
`even at 0° (Fig. 2B).
`The decomposition of I and 5-azacytidine when stored in water and
`in pH 7.4 phosphate buffer at 24° for 24 hr is compared in Fig. 3. Ab(cid:173)
`sorbance measurements at 254 nm of the column eluate showed five major
`peaks for I and two major peaks for 5-azacytidine. At 220 nm, the column
`eluate showed seven major peaks for I and three major peaks for 5-aza(cid:173)
`cytidine.
`
`The decomposition rates of I in acidic, neutral, and basic solutions at
`24 and 37° are shown in Figs. 4A and 4B, respectively. Compound I de(cid:173)
`composed more rapidly at pH 2.2, 8.5, and 9.2 than at pH 5.7, 6.4, and 7.0.
`The decomposition rate of I was the slowest at pH 7.0. At pH 2.2 a peak
`with the same tR value (2.3 min) as 5-azacytosine appeared on the chro(cid:173)
`matogram. The decomposition rate of I was very temperature dependent.
`For example, at pH 7.0, the decomposition rate was about sevenfold
`greater at 37 than at 24°.
`A kinetic study at pH 8.1 and 9.5 was then carried out with II as the
`starting substance. According to Scheme I, the pseudo-first-order rate
`constants k1, k- 1, and k2 were determined using the following differential
`
`5-AZA-2 1-DEOXYCYTIDINE
`H20
`
`IOmin.
`
`Omin. :,
`"'
`N
`0.
`0
`
`"
`
`:,
`0
`o•
`
`pH7.4
`
`:,
`"'
`o.
`
`0
`
`)
`
`I
`
`/
`
`1
`I
`
`:,
`<(
`N
`0
`0
`
`:,
`"'
`6
`d
`
`:,
`"'
`6.
`0
`
`E
`C:
`fi3
`N
`w·
`() z
`ttJ a:
`0
`(I)
`IX)
`<(
`
`<(
`
`0
`
`E
`C:
`~
`N
`w·
`u z
`<(
`IX)
`a:
`0
`en
`a:,
`<(
`
`E
`C:
`o .
`N ,
`N
`w··
`u
`z
`<(
`ai
`a:
`0
`Vl
`tXl
`<(
`
`E
`C:
`;%
`N
`w·
`u
`z
`<(
`IX)
`a:
`0
`Vl
`IX)
`<(
`
`pH7.4
`
`5-AZACYTIDINE
`
`Omin
`
`IOmin.
`
`H20
`
`:,
`
`"'
`"' q
`
`0
`
`:,
`
`"' "'
`
`0.
`0
`
`0
`0
`Figure 3-Decomposition of I and 5-azacytidine in pH 7.4 phosphate buffer and water. 5-Azacytidine (0.5 mM) and I (0.5 mM) were dissolved
`in 10 mM potassium phosphate (pH 7.4) and water and incubated at 24° . At the indicated times, aliquots of these solutions were analyzed by HPLC
`at 2.54 and 220 nm.
`
`'
`
`1230 / Journal of Pharmaceutical Sciences
`Vol. 70, No. 11, November 1981
`
`

`

`Table I-Rate Constants for the Alkaline Decomposition of I and
`5-Azacytidine at 37°
`
`Compound and pH 0
`
`5-Azacytidine, 9.5
`5-Azacytidine, 8.1
`I, 9.5
`
`Pseudo-First-Order Rate Constant, min- 1
`k1
`k-1
`k2
`
`2.0 X 10-1
`1.7 X 10-2
`2.2 X 10-1
`
`2.2 X 10-1
`2.0 X 10-2
`2.8 X 10-1
`
`6.0 X 10- 1
`5.6 X 10-2
`7.5 X 10- 1
`
`0 The pH value was determined at 37° in a separate tube by mixing one volume
`of0.5 M stock sodium borate buffer with 10 volumes of 10 mM potassium phosphate
`(pH 6.8).
`
`9.2
`
`8.5
`
`80
`
`z
`0
`j'.:
`u
`<{
`a:
`u..
`w 40
`...I
`0
`:::E
`
`75
`
`* 6
`z
`z
`~ 50
`~
`w
`a:
`
`25
`
`10L
`
`75
`
`'#.
`(.!) z
`~ 50
`<{
`~
`w
`a:
`
`25
`
`25
`
`30
`
`35
`
`0
`
`5
`
`10
`
`20
`15
`HOURS
`B
`Figure 4-Ef f ect of pH and temperature on the stability of I. Com(cid:173)
`pound I (1.0 mM) was dissolved in 10 mM of the appropriate buffer with
`the indicated pH and incubated at either 24° (top) or37° (bottom). At
`the indicated times, an aliquot of the solution was neutralized with
`phosphate buffer and analyzed by HPLC.
`
`0
`
`20
`
`60
`40
`MINUTES
`Figure 5-Time-concentration profile of decomposition for I at pH 9.5
`and 37°. Key: .&, normalized data for I; ■ , normalized data for II; and
`•• normalized data for III. Solid lines were generated by computer for
`.&, ■, and • using rate constants determined independently.
`
`80
`
`100
`
`the kinetic study of 5-azacytidine, and the results are shown in Table I
`and Figs. 6 and 7.
`
`DISCUSSION
`
`The hydrolysis of 5-azacytidine in alkaline solution and water results
`in the opening of the triazine ring between C-6 and N-1 to form N(cid:173)
`(formylamidino)-N'-/3-D-ribofuranosylurea in a reversible reaction; when
`this latter compound loses the formyl group, it irreversibly forms 1-
`/3-D-ribofuranosyl-3-guanylurea (5, 6). By using quantum chemical cal-
`
`100
`
`80
`
`z
`0
`~ 60
`<{
`a:
`u..
`w
`...I 40
`0
`:::E
`
`20
`
`0
`
`20
`
`0 ~-~-=-::::i;=:::::I:=== ........ .._
`40
`80
`60
`MINUTES
`Figure 6-Time-concentration profile of decomposition for 5-azacy(cid:173)
`tidine at pH 9.5 and 37°. Key: •• normalized data for 5-azacy tidine; ■,
`normalized data for N-(formylamidino)-N'-/3-D-ribofuranosy lurea; and
`• normalized data for 1-/3-D-ribofuranosyl-3-guanylurea. Solid lines
`were generated by computer for •• ■, and • using rate constants de(cid:173)
`termined independently.
`
`100
`
`Journal of Pharmaceutical Sciences I 1231
`Vol. 70, No. 11, November 1981
`
`equations (8):
`Ii= L1II0(-l-e-At +-1-e-Bi)
`B-A
`A-B
`IIi = Ilo (hi -A e-At + k1 - Be-Rt)
`B-A
`A-B
`Ill = II (l - k2(k1 - A) -At_ k2(k1 - B) -Bi)
`o
`A(B-A) e
`B(A-B) e
`Here A and B are roots of the following quadratic equation taken with
`the reverse signs:
`
`(Eq. l)
`
`(Eq. 2)
`
`(Eq. 3)
`
`1
`
`The rate constants thus obtained are presented in Table I and were used
`to generate time-concentration profiles for I-III. Normalized experi(cid:173)
`mental data obtained by HPLC were superimposed on these generated
`time-concentration profiles (Fig. 5). Similar procedures were used for
`
`(Eq. 4)
`
`

`

`100
`
`80
`
`z
`0
`~ 60
`u
`<{
`ex:
`u.
`w
`...J 40
`0
`~
`
`20
`
`z
`0
`~ u
`<{
`ex:
`u.
`w
`...J
`0
`~
`
`80
`
`100
`
`120
`
`0
`
`20
`
`40
`
`60
`I1/IINUTES
`Figure 7-Time- concentration profile of decomposition for 5-azacy(cid:173)
`tidine at pH 8.1 and 37°. Key:• • normalized data for 5-azacytidine; ■,
`normalized data for N-(formylamidino)-N'-{3-o-ribofuranosy lurea; and
`• • normalized data for J-{3-D-ribofuranosyl-3-guanylurea. Solid lines
`were generated by computer for • • ■, and • using rate constants de(cid:173)
`termined independently.
`
`culations, it was shown (5) that the electron density at C-6 of 5-azacy(cid:173)
`tosine is much lower than cytosine, making this position more susceptible
`to nucleophilic attack by a hydroxyl ion.
`From the structural similarity of 5-azacytidine and I, it would seem
`that I would decompose in alkaline solution according to Scheme I. This
`study supports this reaction scheme in alkaline solution. As shown in Fig.
`1, the decomposition of peak I (tR = 5.5 min) at pH 10.4 produced peak
`II (tR = 6.6 min) when the column eluate was monitored at 254 nm. Peak
`III was not observed on this chromatogram because guanylurea deriva(cid:173)
`tives are nonchromophoric at 254 nm (5, 6).
`However, at 220 nm, peak III (tR = 3.5 min) became apparent following
`the decomposition of 11 to III. When peak II was isolated, it had the same
`UV max of 238 nm as reported previously (6) for N-(formylamidino)-N'(cid:173)
`{3-ribofuranosylurea. The hydrolysis of I in alkaline solution produced
`an initial increase in absorbance5 at 238 nm, supporting the formation
`of II, since the formylguanylurea derivative has a much higher extinction
`coefficient than 5-azacytosine derivatives at this wavelength (6). The
`conversion of II to I (Fig. 2A) indicates that this reaction is reversible,
`which is in agreement with other observations (5, 6) for 5-azacytidine.
`There were marked differences in the decomposition of I and 5-aza(cid:173)
`cytidine in phosphate buffer (pH 7.4) and water (Fig. 3). In these solvents,
`5-azacytidine decomposed to form formylguanylurea and guanylurea
`derivatives, as reported previously (6). However, the decomposition of
`I in phosphate buffer or water was very different from that of 5-azacyti(cid:173)
`dine as shown by the presence of five (254 nm) and seven (220 nm) peaks
`on the chromatograms of I. Two of these peaks represent I and II, whereas
`the other peaks were not identified. These unidentified peaks could also
`be produced rapidly by placing II at pH 3.2 (Fig. 2B), even at 0° . These
`unidentified compounds were not the intermediates that eventually
`formed III, since they were not converted to III when placed in alkaline
`buffer, as would happen to I and II under the same conditions.
`Determination of the pseudo-first-order rate constants for I and 5-
`azacytidine at pH 9.5 and 37° demonstrated that both compounds de(cid:173)
`composed at comparable rates, although the formylguanylurea derivative
`for ribose appeared to be slightly more stable than that of deoxyribose
`(Table I). A 22-fold decrease in the hydroxyl-ion concentration (pH
`9.5- 8.1) reduced all three rate constants for 5-azacytidine equally by
`about 11-fold. Lowering the pH from 9.5 to 8.1 also greatly reduced the
`conversion rate of lI to I and to III (Figs. 5 and 8). The kinetic data ob(cid:173)
`t.ained for I at pH 8.1 did not quite fit in Eqs. 1-3 because apparently some
`of the decomposition did not follow Scheme I precisely due to the for(cid:173)
`mation of minor unidentified peaks. Thus the rate constants were not
`determined. These unidentified peaks became increasingly prominent
`as the pH was lowered, and there was virtually no conversion of II to I at
`pH <4.6. (Fig. 2B) .
`
`0
`
`20
`
`40
`
`60
`MINUTES
`Figure 8-Time-concentration profile of decomposition for I at pH 8.1
`and 37° . Key: •• normalized data for I; ■, normalized data for II; and
`•• normalized data for III.
`
`80
`
`Comparison of the overall stability profile of I in aqueous buffer so(cid:173)
`lution at various pH levels showed that I was most stable at neutral pH
`and at low temperature (Fig. 4). The instability of I at high pH could be
`explained on the basis that the opening of the triazine ring (k 1) and the
`breakage of an amide bond (k 2) is facilitated by a hydroxyl ion. Reduction
`in pH would slow down both these events and thus stabilize I. However,
`I became increasing unstable as the pH was gradually reduced below 7 .0.
`This result might be due to the increase in the rate of ring opening as well
`as to the instability of II in acidic solutions; thus, once ring opening took
`place and produced II, the latter broke down to the unidentified peaks
`more rapidly and minimized the reversal back to I (Fig. 2B) . In strong
`acidic solution (with pH <2.2), breakage of the glycolytic bond of I oc(cid:173)
`curred, producing 5-azacytosine6 as reported previously for I (9) and for
`5-azacytidine (5, 10).
`The effect of pH and temperature on the chemical stability of 5-aza(cid:173)
`cytidine was studied by Chan et al. (11), who observed that this com(cid:173)
`pound is most stable at pH 7 .0 and that its rate of decomposition increases
`in solutions of high or low pH and with temperature.
`
`REFERENCES
`
`(1) F. Sorm and J. Vesely, Neoplasma, 15,339 (1968).
`(2) R. L. Momparler and F. A. Gonzales, Cancer Res., 38, 2673
`(1978).
`(3) R. L. Momparler and J. Goodman, ibid., 37, 1636 (1977).
`(4) W. R. Volger, D. S. Miller, and J . W. Keller, Blood, 48, 331
`(1976).
`(5) P. Pithova, A. Piskala, J. Pitha, and F. Sorm, Collect . Czech.
`Chem. Commun., 30, 2801 (1965).
`(6) J . A. Heisler, J . Med. Chem., 21, 204 (1978).
`(7) J. Pliml and F. Sorm, Collect. Czech. Chem. Commun., 29, 2576
`(1964).
`(8) N. M. Rodiguin and E. N. Rodiguina, "Consecutive Chemical
`Reactions," Van Nostrand, Princeton, N.J., 1979.
`(9) A. Piskala, M. Synackova, H. Tomankova, P. Fiedler, and V.
`Zizkovsky, Nucleic Acid Res., 4, s109 (1978).
`(10) R. E. Notari and J. L. DeYoung, J. Pharm. Sci., 64, 1148
`(1975).
`(11) K. K. Chan, D. D. Giannini, J. A. Staroscik, and W. Sadee, ibid.,
`68,807 (1979).
`
`ACKNOWLEDGMENTS
`
`Supported by U.S. Public Health Service Grant CA23340 from the
`National Cancer Institute.
`The authors thank Suzanne Beaudet for assistance in preparation of
`this manuscript.
`
`5 R. L. Momparler, unpublished data.
`
`6 Unpublished observations.
`
`1232 / Journal of Pharmaceutical Sciences
`Vol. 70, No. 11, November 1981
`
`

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