首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Nefopam hydrochloride degradation kinetics in solution   总被引:2,自引:0,他引:2  
A stability-indicating reversed-phase high performance liquid chromatographic method was developed for the detection of nefopam hydrochloride and its degradation products under accelerated degradation conditions. The degradation kinetics of nefopam hydrochloride in aqueous solutions over a pH range of 1.18 to 9.94 at 90 +/- 0.2 degrees C was studied. The degradation of nefopam hydrochloride was found to follow apparent first-order kinetics. The pH-rate profile shows that maximum stability of nefopam hydrochloride was obtained at pH 5.2-5.4. No general acid or base catalysis from acetate, phosphate, or borate buffer species was observed. The catalytic rate constants on the protonated nefopam imposed by hydrogen ion and water was determined to be 7.16 X 10(-6) M-1 sec-1, and 4.54 X 10(-9) sec-1, respectively. The pKa of nefopam hydrochloride in aqueous solution was determined to be 8.98 +/- 0.33 (n = 3) at 25 +/- 0.2 degrees C by the spectrophotometric method. The catalytic rate constant of hydroxyl ion on the degradation of nefopam in either protonated or nonprotonated form was determined to be 6.63 X 10(-6) M-1 sec-1 and 4.06 X 10(-6) M-1 sec-1, respectively. A smaller effect of hydroxyl ion on the degradation of nonprotonated than on the degradation of protonated nefopam was observed.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The degradation kinetics of nicotinyl 6-aminonicotinate in aqueous buffer solutions were studied over the pH range from 4.0 to 10.0. In all cases, pseudo-first-order kinetics were observed at constant hydronium ion concentration. The pH-rate profile indicated that the hydrolysis of nicotinyl 6-aminonicotinate may be described by at least two catalytic terms. In alkaline solution the hydrolysis is catalyzed primarily by hydroxyl ions. In acidic solution the hydrolysis may be attributed to either the water-catalyzed reaction of the protonated species or the hydronium ion catalyzed reaction of the free base. The resulting catalytic profile afforded a sharp pH minimum of approximately 5.90 at 65 degrees C. An activation energy of 16 Kcal/mol was obtained in a phosphate buffer solution at a pH of approximately 5.90 +/- 0.2. The first- and second-order reaction constants for water and hydroxyl ion catalysis were determined, and the temperature dependency of the reaction was studied. The buffer effect and solvent effect on the hydronium and hydroxyl ion catalysis was also investigated.  相似文献   

3.
The stability of N-cyclohexanecarbonyl-3-(4-morpholino)-sydnone imine hydrochloride (ciclosidomine) in solution was studied as a function of pH, temperature, ionic strength, and buffer species. The rate of hydrolysis in the absence of light was found to be apparent first order in drug and general acid- and base-catalyzed reactions. The pH rate profile at an ionic strength of 0.1 M at 60 degrees C had a minimum value near pH 6. Change in ionic strength in the range of 0.05 to 0.2 M did not affect the rate of degradation at pH 7 (carbonate buffer) or pH 2 (phosphate buffer) at 60 degrees C. Similar degradation rates were noticed in air or nitrogen in the dark at pH 3, 5, and 6. However, degradation in light was very rapid in either case at pH 3, 5, and 6, and, therefore, the protection of solutions from light was required during all studies. The time for 10% loss of drug in solution at pH 6 in dilute phosphate or citrate buffer at an ionic strength of 0.154 M was projected to be 9 months at 20 degrees C and 2.6 months at 30 degrees C.  相似文献   

4.
The degradation of batanopride hydrochloride, an investigational antiemetic drug, was studied in aqueous buffer solutions (pH 2-10; ionic strength, 0.5; 56 degrees C) in an attempt to improve drug stability for parenteral administration. Degradation occurs by two different mechanisms depending on the pH of the solution. In acidic media (pH 2-6), the predominant reaction was intramolecular cyclization followed by dehydration to form a 2,3-dimethylbenzofuran. There was no kinetic or analytical (high-performance liquid chromatography) evidence for the formation of an intermediate; therefore, the rate of dehydration must have been very rapid compared with the rate of cyclization. In alkaline media (pH 8-10), the primary route of degradation was cleavage of the C-O alkyl ether bond. In the intermediate pH range (pH 6-8), both reactions contributed to the overall degradation. Both degradation reactions followed apparent first-order kinetics. The pH-rate profile suggests that batanopride hydrochloride attains its optimal stability at pH 4.5-5.5. Citrate buffer was catalytic at pH 3 and 5, and phosphate buffer was catalytic at pH 8. No catalytic effect was observed for the borate buffer at pH 9-10.  相似文献   

5.
First-order rate constants (k) were determined for cefepime degradation at 45, 55, 65, and 75 degrees C, pH 0.5 to 8.6, using an HPLC assay. Each pH-rate profile exhibited an inflection between pH 1 and 2. The pH-rate expression was k(pH) = kH1 f1(aH+) + kH2 f2(aH+) + ks + kOH(aOH-), where kH1 and kH2 are the catalytic constants (M(-1) h(-1)) for hydrogen ion activity (aH+), kOH is the catalytic constant for hydroxyl ion activity (aOH-), and ks is the first-order rate constant (h(-1)) for spontaneous degradation. The protonated (f1) and unprotonated (f2) fractions were calculated from the dissociation constant, Ka = (8.32x10(-6))e(5295)/RT where T was absolute temperature (T). Accelerated loss due to formate, acetate, phosphate, and borate buffer catalysis was quantitatively described with the catalytic constant, kGA (M(-1) h(-1)) for the acidic component, [GA], and kGB (M(-1) h(-1) for the basic component, [GB], of each buffer. The temperature dependency for each rate constant was defined with experimentally determined values for A and E and the Arrhenius expression, kT = Ae-E/RT, where kT represented kH1, kH2 , kS, kOH, kGA, or kGB. Degradation rate constants were calculated for all experimental pH, temperature, and buffer conditions by combining the contributions from pH and buffer effects to yield, k = k(pH) + kGA[GA] + kGB[GB]. The calculated k values had <10% error for 103 of the 106 experimentally determined values. Maximum stability was observed in the pH-independent region, 4 to 6. Degradation rate constants were predicted and experimentally verified for cefepime solutions stored at 30 degrees C, pH 4.6 and 5.6. These solutions maintained 90% of their initial concentration (T90) for approximately 2 days.  相似文献   

6.
The objective of this study was to evaluate the stability of ceftiofur (1 mg mL(-1)) in aqueous solutions at various pH (1, 3, 5, 7.4 and 10) and temperature (0, 8, 25, 37 and 60 degrees C) conditions. The ionic strength of all these solutions was maintained at 0.5 M. Ceftiofur solutions at pH 5 and 7.4 and in distilled water (pH = 6.8) were tested at all the above temperatures. All other solutions were tested at 60 degrees C. Over a period of 84 h, the stability was evaluated by quantifying ceftiofur and its degradation product, desfuroylceftiofur, in the incubation solutions. HPLC was used to analyse these compounds. At 60 degrees C, the rate of degradation was significantly higher at pH 7.4 compared with pH 1, 3, 5 and distilled water. At both 60 degrees C and 25 degrees C, degradation in pH 10 buffer was rapid, with no detectable ceftiofur levels present at the end of 10 min incubation. Degradation rate constants of ceftiofur were 0.79+/-0.21, 0.61+/-0.03, 0.44+/-0.05, 1.27+/-0.04 and 0.39+/-0.01 day(-1) at pH 1, 3, 5, 74 and in distilled water, respectively. Formation of desfuroylceftiofur was the highest (65%) at pH 10. The rate of degradation increased in all aqueous solutions with an increase in the incubation temperature. At pH 7.4 the degradation rate constants were 0.06+/-0.01, 0.06+/-0.01, 0.65+/-0.17, and 1.27+/-0.05 day(-1) at 0, 8, 25, 37 and 67 degrees C, respectively. The energy of activation for ceftiofur degradation was 25, 42 and 28 kcal mol(-1) at pH 5, 7.4 and in distilled water, respectively. Desfurylceftiofur formation was the greatest at alkaline pH compared with acidic pH. Ceftiofur degradation accelerated the most at pH 7.4 and was most rapid at pH 10. The results of this study are consistent with rapid clearance of ceftiofur at physiological pH.  相似文献   

7.
The influence of pH, phosphate buffer components and temperature on the degradation rate of busulfan was studied. The analysis was performed using gas chromatography with electron capture detection and reversed-phase liquid chromatography with radioactivity monitoring. The degradation rate of busulfan showed no pH dependence in the range pH 1.5-11 and increased at higher pH values. The degradation rate constant was 0.034 +/- 0.001 h(-1) (S.E.M.) for the degradation of busulfan in pure water and 0.45 +/- 0.01 h(-1)M(-1) (S.E.M.) for the reaction of busulfan with the hydroxide ion at 37 degrees C. The reactivity of HPO(4)(-2) was six times higher than the reactivity of H(2)PO(4)(-1) towards busulfan. The hydrolysis products were identified as tetrahydrofuran and methanesulphonic acid by nuclear magnetic resonance spectroscopy.  相似文献   

8.
The kinetics and mechanism of degradation of mibolerone were studied in aqueous buffered solutions in the pH range of 1-8 at 67.5 degrees. Mibolerone showed maximum stability between pH 5.5 and 6.4. At pH 1-2, the major degradative pathway was dehydration followed by migration of the 18-methyl group to form 7alpha,17,17-trimethylgona-4,13-dien-3-one. While there was only one degradation product at pH 1-2, the degradation at pH 7-8 was complex. As many as 12 degradation products were detected by GLC. Mass spectral data indicated that the majority of these products were either oxidation products or isomers. At pH 7.6, the apparent first-order rate constants exhibited marked dependency on buffer concentration. Incorporation of a sequestering agent into the solutions eliminated this dependency, suggesting that trace metal impurities from the buffer reagents were catalyzing the degradation. This was confirmed by degradation studies of solutions in water for injection containing 5 ppm of trace metal ions. Sn+2, Cu"2, and Fe+2 accelerated the degradation, with Fe+2 having the most catalytic effect. The temperature dependence of the rate of degradation was studied in 0.05 M phosphate buffer at pH 6.4. The activation energy was 19.6 +/- 1.63 kcal/mole.  相似文献   

9.
The degradation of Ro-26-9228, 1alpha-fluoro-25-hydroxy-16,23E-diene-26,27-bishomo-20-epi-cholecalciferol, 2, was studied in aqueous solution in the pH range of 1.17-10.56 and in alcohol solutions, at 25, 40, and 50 degrees C. The degradation of Ro-26-9228 was found to be acid catalyzed and to be independent of potassium acetate buffer concentration. Above pH 4, the reaction rate is independent of pH, with a T90 of 14.3 h at 25 degrees C in pH 7.75 buffer. 19F nuclear magnetic resonance was used to study the ratio of the vitamin (6-s-trans) to previtamin form in acetonitrile at 40 degrees C. The equilibrium percentage of previtamin and the rate of approach to equilibrium were 13.8% and 0.2 h(-1), respectively. Nuclear magnetic resonance was used to elucidate the structure of the degradation products. Novel products were formed from the elimination of the fluorine and addition of solvent to C9, with formation occurring through the previtamin form. Additional degradation products result from reaction of the side chain 25-hydroxyl and addition of solvent to C1.  相似文献   

10.
Light and temperature have considerable effect on the degradation of piroxicam in aqueous solutions. The pH and acetate buffer ions also affect the degradation process. The apparent first-order rate constants for the photochemical and thermal degradation of piroxicam have been determined as 2.04-10.01 and 0.86-3.06×10(-3) min(-1), respectively. The first-order plots for the degradation of piroxicam showed good linearity within a range of 20-50% loss of piroxicam at pH 2.0-12.0. The rate-pH profile for the photodegradation of piroxicam is a U-shaped curve and for the thermal degradation a bell-shaped curve in the pH range of 2.0-12.0. The thermal degradation of piroxicam was maximum around pH 6.0. It is increased in the presence of acetate ions but was not affected by citrate and phosphate ions.  相似文献   

11.
Chlorpheniramine maleate was incorporated into disks consisting of glyceryl fatty acid esters, polyethylene glycol fatty acid esters or a combination of the two. A melt-extrusion process was used to prepare the matrix disks containing the drug. The release of the drug into distilled water, pH 1.2 buffer, and pH 7.5 buffer showed the expected square root of time dependence. An increase in the fatty acid ester hydrophilic-lipophilic balance (HLB) from 1 to 14 resulted in a 10-fold increase in the drug release rate from 0.25 +/- 0.01 to 25.84 +/- 1.29 mg cm-2 h-1/2. The maximum release rate was seen from the fatty acid ester with a melting point of 44 degrees C. The pH of the dissolution medium had a small impact on the rate of drug release, but the rate of agitation had no significant influence on the rate of drug release. By blending a fatty acid ester of a high melting point (64 degrees C) and a low HLB value of 2 with esters of lower melting points (33 to 50 degrees C) or higher HLB values (10 to 14), it was possible to modify the release from 10.0 +/- 0.70 to 21.5 +/- 0.57 mg cm-2 h-1/2.  相似文献   

12.
L-NDDP is a liposome-entrapped platinum compound currently in phase 2 clinical trials that has been shown to undergo intraliposomal activation. The degradation/activation kinetics of liposome entrapped cis-bis-neodecanoato-trans-R,R-1,2-diamminocyclohexane platinum (II) [L-NDDP] at different conditions of pH, and temperature is presented. Liposomes were reconstituted in a solution of 0.9% sodium chloride (NaCl) in water (pH 5) at room temperature (formulation conditions currently used in the ongoing clinical trials). In the temperature experiments, L-NDDP 0.9% sodium chloride liposomes were incubated in a water-bath at 40, 60, and 80 degrees C. In the pH experiments, these solutions were compared to water, phosphate with and without chloride ion present, phosphate buffer without chloride ion at pH 3.1, 5.0, and 7.4, and glycine buffer with and without chloride ion. In 0.9% sodium chloride at room temperature, the chemical degradation/activation of liposome-bound NDDP was biphasic, with most of the degradation (approximately 45% conversion) occurring during the first hour after formation of the liposome suspension. NDDP degradation was pH dependent: when using pH 3 phosphate buffer as a reconstituting solution, liposome-bound NDDP degraded rapidly, whereas in pH 7.4 phosphate buffer it was stable for > 72 h. NDDP degradation was also temperature-dependent, the 50% point decreasing from 12 h at 25 degrees C to 9.5 h at 40 degrees C, 3.8 h at 60 degrees C, and 0.3 h at 80 degrees C when using 0.9% NaCl in water as a reconstituting solution. Using glycine buffer solution with and without NaCl at room temperature, no NDDP degradation over a 72 h period was observed at 25 degrees C; however, at 40 degrees C, only 68% NDDP remained intact at 72 h. Atomic absorption spectrophotometry (AAS) analysis of the eluting fractions after injection of L-NDDP samples reconstituted in chloride-containing and non chloride-containing solutions clearly indicated that the formation of DACH-Pt-Cl2 was only observed when chloride-containing solutions were used and was first detected at 3 h when using 0.9% NaCl in water as a reconstituting solution. These results indicate that pH and temperature, and not the presence of chloride ion, are the main factors leading to the activation of NDDP. Since 45% of NDDP is already degraded at 1 h in the same conditions, it is concluded that (1) the first active intermediates of L-NDDP formed within the liposomes are the DACH-Pt chloro-aquo and diaquo intermediates, and (2) the in vivo, antitumour activity of L-NDDP is most likely mediated by direct intracellular delivery of the active species.  相似文献   

13.
The stability of cefazolin sodium reconstituted in four artificial tear solutions, two acetate buffer solutions, phosphate buffer solution, and 0.9% sodium chloride injection was studied. Cefazolin was reconstituted in Tearisol, Isopto Tears, Liquifilm Forte, and Liquifilm Tears; acetate buffer solution at pH 4.5 and pH 5.7; phosphate buffer solution at pH 7.5; and 0.9% sodium chloride injection. The solutions were stored at 4 degrees C, 25 degrees C, and 35 degrees C for seven days. All of the solutions were inspected for particulates, turbidity, color, and odor. Five assay determinations on each of three samples of each formulation were performed using a stability-indicating high-performance liquid chromatographic assay. Cefazolin stability was influenced primarily by pH and storage temperature. Reconstitution of cefazolin sodium in the alkaline tear solutions Isopto Tears and Tearisol and in phosphate buffer solution resulted in particulate and color formation at 25 degrees C and 35 degrees C. Turbidity was noted after cefazolin sodium was reconstituted in Isopto Tears. No color or precipitate formation was evident after seven days at 25 degrees C and 35 degrees C in the formulations of acidic pH containing Liquifilm Tears, Liquifilm Forte, 0.9% sodium chloride injection or acetate buffer solution as the vehicles. The extent of degradation of cefazolin was substantially higher in the formulations of alkaline pH than in those of acidic pH at 35 degrees C and 25 degrees C. All of the formulations retained more than 90% of their initial concentration when stored at 4 degrees C. Cefazolin sodium, when reconstituted in artificial tear solutions with an acidic pH, is stable for up to three days at room temperature.  相似文献   

14.
In the absence of strong buffer catalysts, the degradation of codeine sulfate (7,8-didehydro-4,5 alpha-epoxy-3-methoxy-17-methylmorphinan-6 alpha-ol sulfate) in aqueous solution is described by the expression kobs = kH+ [H+] + k0 + kHO-[HO-], where kH+ = (3.9 +/- 1.3) X 10(-8) M-1 X S-1, k0 = (2.7 +/- 0.5) X 10(-8) S-1, and kHO- = (5.1 +/- 1.0) X 10(-6) M-1 X S-1 at 80 degrees C. The activation energies associated with these rate constants are 27.7, 21.0, and 28.3 kcal X mol-1, respectively. In the absence of buffer catalysis, codeine sulfate is predicted to have a room temperature shelf life of approximately 44 years between pH 1 and 10, significantly longer than the 1.1 year shelf life of codeine phosphate reported earlier.  相似文献   

15.
Mbah CJ 《Die Pharmazie》2007,62(10):747-749
The degradation kinetics of avobenzone in aqueous solution was studied at 60 +/- 0.2 degrees C over a pH range of 2.0-10.0. The degradation rates were determined by high performance liquid chromatography. The reaction is found to follow first-order kinetics and the rate constant for the decomposition at 25 degrees C is estimated by extrapolation. The breakdown of avobenzone is shown to be hydroxide ion catalyzed and the Arrhenius plots showed the temperature dependence of avobenzone degradation.  相似文献   

16.
Fomepizole (4-Methylpyrazole, 4MP) is used as an antidote for ethylene glycol and methanol poisoning. In France it is administered intravenously as the sulfate or hydrochloride salt formulation and in the United States, as the free base formulation. Since its pKa was unknown, it was unknown if the free base, hydrochloride and sulfate salt formulations of 4MP are chemically equivalent, and if 4MP is in chemically equivalent forms in blood when these base and salt formulations are administered intravenously. Using UV spectrophotometry, the pKa of 4MP was determined to be 2.91 +/- 0.05 (n = 7) at a low concentration of 0.006 M in formate buffers of various pH, and 3.00 +/- 0.16 (n = 7) when high concentration of 4MP (0.06 M) was titrated with HCl at 25 degrees C. The hydrochloride salt formulation (pH 6.64) was ionically and hence chemically equivalent to the free base formulation (pH 7.02), while the sulfate salt formulation, due to its lower pH of 2.33, showed presence of some ionized 4MP indicating chemical inequivalence. In order to determine chemical equivalence upon intravenous administration, these formulations were diluted with phosphate buffer (pH 7.4) with identical ionic strength and buffer capacity as blood. In spite of chemical inequivalence of the sulfate salt formulation, 4MP free base was observed from all three formulations when diluted with physiological buffer suggesting chemical equivalence under physiological conditions due to the strong buffering action of blood.  相似文献   

17.
The stability of the new antileukemic kojic acid derivative, 5-benzyloxy-2-thiocyanatomethyl-4-pyranone (BTMP) was investigated. The degradation of BTMP was studied using specific and reproducible HPLC and LC-MS methods. Accelerated stability studies of BTMP were conducted in 0.1 M hydrochloric acid solution, physiological phosphate buffer solution (pH 7.5) and basic phosphate buffer solution (pH 9.0) at 30, 40 and 60 degrees C, respectively. The degradation of BTMP was found to follow pseudo-first order kinetics. In basic solution (pH 9.0) BTMP underwent rapid hydrolysis at a degradation rate constant (0.183-0.638 h-1) and degradation half-life (3.67-1.06 h) depending on the temperature setting. On the other hand, BTMP was significantly stable in 0.1 M hydrochloric acid solution (kdeg: 0.0017-0.0052 h-1; degradation half-life t1/2: 408.6-135.7 h), whereas in physiological phosphate buffer solution (pH 7.5), BTMP was only moderately stable (kdeg: 0.006-0.231 h-1; degradation half-life: 117.7-3.0 h). Arrhenius plots were constructed to predict the degradation kinetic parameters of BTMP at 25 degrees C and 4 degrees C. LC-MS analyses confirmed the degradation of BTMP in basic solutions and indicated at least two degradation products; namely 5-benzyloxypyran-2-ol-4-one (m/z 217.8) and 2-thiocyanatomethylpyran-5-ol-4-one (m/z 181.6).  相似文献   

18.
Temperature stability studies were conducted on 0.36 M (6.5% W/V) aspirin solutions including either 0.02 M L-methionine or 0.02 M histidine in propylene glycol. Aspirin was determined spectrophoto-fluorometrically as salicylic acid content at 412 nm. A 0.36 M aspirin in polyethylene glycol 400 solution was studied concurrently. Aspirin degradation rate constants, k, obtained from semilogarithmic plots of percent drug remaining versus time at 30-70 +/- 0.5 degrees were used for preparing Arrhenius plots. Good correlation was seen between predicted aspirin stability and experimental k25 degrees values. L-Methionine and histidine markedly reduced aspirin stability.  相似文献   

19.
The solubility of bumetanide in vehicles of various polarities, suitable for intranasal administration in acute situations, has been investigated. The solubility at 4 degrees C in glycofurol and polyethylene glycol 200 was high (167 and 143 mg/mL, respectively), decreasing exponentially with addition of phosphate buffer or coconut oil. Vehicles containing coconut oil and glycofurol did not seem to improve the solubility relative to mixtures between glycofurol and buffer. Adequate solubility (approximately 50 mg/mL) was achieved in vehicles containing about 80% cosolvent. The stability of bumetanide was studied at 5 degrees C and 57 degrees C. No degradation was observed at low temperature. At high temperature, bumetanide decomposes in nonaqueous vehicles with half-lifes ranging from 69 to 400 days, but sufficient stability may be obtained by adjustment of pH to 7.4. It may be concluded that it is possible to prepare a clinically relevant formulation for intranasal delivery of bumetanide.  相似文献   

20.
Hydrolysis of mazindol to form 2-(2-aminoethyl)-3-(p-chlorophenyl)-3-hydroxyphthalimidine was followed spectro-photometrically in aqueous solutions at temperatures between 37 and 70degree, pH values up to 7.6, and an ionic strength of 0.2. The effects of acetate, formate, and phosphate buffers as well as ionic strength on the observed rate constants were investigated. An interesting nonlinear dependency of the kobs with buffer concentration was noted. The velocity constants declined with increasing hydrogen-ion concentration; the log k-pH profile and rate law are given along with other relevant data.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号