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.
Purpose. To develop a method for calculating epimerisation parameters, find out if the kinetics of the independent reactions can be established, and elucidate primary structure-chemical degradation relationships in the degradation kinetics of three gonadorelin analogues. Methods. The influences of pH, temperature, and buffer concentration on the degradation of the three gonadorelin analogues buserelin, goserelin, and triptorelin were investigated using RP-HPLC. A method was developed to calculate epimerisation and hydrolysis rate constants independently. Results. Explicit structure-degradation mechanism relations were found in the degradation of all three compounds. The L-serine residue was found to be involved in both a solvent-catalysed backbone hydrolysis and a hydroxyl-catalysed epimerisation whereas, the O-tertiary butyl D-serine residue was only involved in proton-catalysed ether hydrolysis. The kinetics of identical reactions in different analogues were generally comparable. Conclusions. The degradation of the gonadorelin analogues is located at a relatively small number of chemical residues and prediction of the degradation mechanisms and kinetics of other peptides with similar structural elements appears to be possible.  相似文献   

2.
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   

3.
The degradation kinetics of phentolamine hydrochloride in aqueous solution over a pH range of 1.2 to 7.2 and its stability in propylene glycol- or polyethylene glycol 400-based solutions were investigated. The observed rate constants were shown to follow apparent first-order kinetics in all cases. The pKa determination for phentolamine hydrochloride was found to be 9.55 +/- 0.10 (n = 5) at 25 +/- 0.2 degrees C. This indicated the protonated form of phentolamine occurs in the pH range of this study. The pH-rate profile indicated a pH-independent region (pH 3.1-4.9) exists with a minimum rate around pH 2.1. The catalytic effect of acetate and phosphate buffer species is ordinary. The catalytic rate constants imposed by acetic acid, acetate ion, dihydrogen phosphate ion, and monohydrogen phosphate ion were determined to be 0.018, 0.362, 0.036, and 1.470 L mol-1 h-1, respectively. The salt effect in acetate and phosphate buffers followed the modified Debye-Huckel equation quite well. The ZAZB value obtained from the experiment closely predicts the charges of the reacting species. The apparent energy of activation was determined to be 19.72 kcal/mol for degradation of phentolamine hydrochloride in pH 3.1, 0.1 M acetate buffer solution at constant ionic strength (mu = 0.5). Irradiation with 254 nm UV light at 25 +/- 0.2 degrees C showed a ninefold increase in the degradation rate compared with the light-protected control. Propylene glycol had little or no effect on the degradation of phentolamine hydrochloride at 90 +/- 0.2 degrees C; however, polyethylene glycol 400 had a definite effect.  相似文献   

4.
Purpose. A theory and computer programs running on Microsoft® Excel for Windows for calculation of solution pH, buffer formula, and buffer capacity at a given ionic strength and temperature were developed. The theory does not limit the category of buffer components, the number of buffer components, or the number of ionizations for each buffer component. The usefulness of the programs was examined. Methods. The formulas for 7 single component buffer solutions and 2 multiple component buffer solutions composed of citrate, phosphate, Tris, borate, and glycine were calculated. The solution pH values were measured at 25, 40, 55 and 70°C for comparison with the calculated pH values. Results. Of the 108 predictions made, 96 were of pH values within ± 0.1 pH unit of the measured values, at temperatures ranging from 25°C to 70°C and at ionic strengths ranging from 0.1 M to 0.5 M. Conclusions. These programs will be useful for identifying appropriate buffer solutions at various temperatures and/or ionic strengths.  相似文献   

5.
Purpose. To determine the kinetics and mechanism of hydrolysis of efavirenz [(S)-6-chloro-4-(cyclopropylethynyl)-1,4-dihydro-4-(trifluoromethyl)-2H-3,1-benzoxazin-2-one] in aqueous solutions. Methods. The solution stability was examined at 60°C and an ionic strength of 0.3 M over the pH range of 0.6 to 12.8. The loss of efavirenz and the appearance of degradants were followed with a reverse-phase high-performance liquid chromatography assay. Characterization of the degradants was accomplished with liquid chromatography-mass spectrometry. Results. The degradation of efavirenz followed apparent first-order kinetics over the pH range of 0.6 to 12.8 at 60°C. The catalytic effect of phosphate and borate buffers was negligible while acetate and citrate demonstrated buffer catalysis. The overall rate constant indicated a pH minimum (the pH of maximum stability) of approximately 4. Mass spectra data identified a degradant with a molecular weight consistent with hydrolysis of the cyclic carbamate to the corresponding amino alcohol. The degradation route was confirmed with spiking experiments with an authentic sample of the amino alcohol indicating that the carbamate hydrolysis pathway was the predominant reaction throughout the pH range studied. Subsequent degradation of the amino alcohol proceeded at the extremes of the pH range studied via rearrangement to the quinoline. Conclusions. The pH-rate profile was consistent with a combination of a V-shaped profile in the pH range of 0-9 and a sigmoid-shaped profile in the pH range of 4-13. The plateau that began at pH 10-11 is a result of the ionization of the amine of the carbamate inhibiting the base-catalyzed hydrolysis of efavirenz, given that the ionized form of the carbamate is resonance-stabilized toward hydroxide-catalyzed degradation. Thus, increasing the pH resulted in a parallel decrease in the unionized fraction and increase in hydroxide ion concentration resulting in a constant kobs value.  相似文献   

6.
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   

7.
Many cyclic peptides are formed using a disulfide bond to increase their conformational rigidity; this provides receptor selectivity and increased potency. However, degradation of the disulfide bond in formulation can lead to a loss of structural stability and biological activity of the peptide. Therefore, the objective of this study was to study the stability of peptide 1 (cyclo(1,4)-Cys-Gly-Phe-Cys-Gly-OH). This cyclic peptide was synthesized using Boc strategy via solution-phase peptide synthesis and purified using semi-preparative HPLC. The accelerated stability studies of the cyclic peptide were conducted in buffer solutions at pH 1.0-11.0 with controlled ionic strengths at 70 degrees C. The pH-rate profile shows that the peptide has an optimal stability around pH 3.0 with a V-shape between pH 1.0 and 5.0. Two small plateaus are observed at pH 5.0-7.0 and pH 8.0-10.0, indicating hydrolysis on different ionized forms of the cyclic peptide. One product was observed at acidic pH due to peptide bond hydrolysis at Gly2-Phe3. The number of degradation products increases as the pH increases from neutral to basic, and most of the degradation products at neutral and basic pH are derived from the degradation at the disulfide bond.  相似文献   

8.
Abstract

The objective of this study was to investigate the effect of the physiological parameters (pH, buffer capacity, and ionic strength) of the gastrointestinal (GI) fluid on the dissolution behavior of the class II weakly acidic (BCS class IIa) drug valsartan. A series of in vitro dissolution studies was carried out on Diovan® immediate release tablets using media that cover the physiological range of pH (1.2–7.8), buffer capacity (0–0.047?M/ΔpH), and ionic strength (0–0.4?mol/L) of the GI fluid during fasted and fed states using the conventional USP II apparatus. Valsartan exhibited pH- and buffer capacity-dependent dissolution behavior, where valsartan release was slow and incomplete in media simulating gastric fluid with low pH, and fast and complete in media simulating intestinal fluid with high pH. In addition, the rate of valsartan release increased with increasing the buffer capacity of the dissolution medium. In water and NaCl solutions, valsartan release was incomplete and the dissolution profiles were similar regardless of the ionic strength of the medium, indicating an ionic strength-independent dissolution behavior. These results highlight the significant effect of the physiological parameters of the GI fluid on the dissolution behavior of BCS class IIa drugs.  相似文献   

9.
Purpose. PNU-67590A is a water-soluble micellar prodrug of methyl-prednisolone (MP). The major products of degradation of PNU-67590A are MP by hydrolysis and methylprednisolone 17-suleptanate (17-E) by 21 17 acyl migration. The effect of ionic strength on micelle formation and stability of PNU-67590A in aqueous solution was examined. Methods. PNU-67590A solutions at pH 2 and 8 and ionic strength of 0.05, 0.1, 0.2, and 0.4 M were maintained at 25°C in the dark to measure MP and 17-E levels over time. Results. The rate of degradation of micellar PNU-67590A at pH 8 was less than that of monomeric PNU-67590A, and vice versa at pH 2. Increase in ionic strength decreased both the critical micelle concentration of PNU-67590A and the degradation of micelle PNU-67590A at both pHs, resulting in improved overall stability of PNU-67590A. Conclusions. Formulation of PNU-67590A in a concentrated solution with high ionic strength will maximize stability and shelf-life.  相似文献   

10.
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   

11.
The effects of pH (3.7—13), ionic strength, buffer composition (acetate, phosphate and borate) and buffer concentration (15–200 mM) on the rate of degradation of melphalan in the presence 0.3 M chloride at 50 ± 0.1 °C were investigated using high-performance liquid chromatography. In addition, the data published in the literature for the degradation of phosphoramide mustard have been compared with those of melphalan, placing emphasis on mechanisms of hydrolysis and the effects of pH and chloride. In the presence of chloride, the degradation rate of melphalan was influenced by pH and buffer composition but not by ionic strength. These effects were not seen in the absence of added chloride and have been explained in terms of competition between chloride and other nucleophiles such as the hydroxide ion, water and buffer components for the active intermediate of the alkylating agent. These results help to explain differences in reported values for the rates of hydrolysis of various alkylating agents in the presence of chloride.  相似文献   

12.
Methazolamide (MTZ) is an anti-glaucoma drug. The present paper aims to characterize the physicochemical properties and degradation kinetics of MTZ to provide a basis for topical ophthalmic delivery. With the increase in pH (pH 5.5–8.0) of aqueous solution, the solubility of the compound increased while the partition coefficient (Ko/w) which was estimated in the system n-octanol/aqueous solution decreased. The degradation of MTZ in aqueous solution followed pseudo-first-order kinetic. The degradation rate kpH is the rate in the absence of buffer catalysis. Plotting the natural logarithm of kpH versus the corresponding pH value gave a V-shaped pH-rate profile with a maximum stability at pH 5.0. The degradation rate constants as a function of the temperature obeyed the Arrhenius equation (R2 = 0.9995 at pH 7.0 and R2 = 0.9955 at pH 9.0, respectively). A decrease in ionic strength and buffer concentration displayed a stabilizing effect on MTZ. Buffer species also influenced the MTZ hydrolysis. Phosphate buffer system was more catalytic than tris and borate buffer systems. In brief, it is important to consider the physicochemical properties and the stability of MTZ during formulation.  相似文献   

13.
The degradation kinetics of methanolic solution of danazol (0.020% w/v) in aqueous buffers and sodium hydroxide was investigated using stability-indicating HPLC method. The drug degrades in alkaline medium through a base-catalysed proton abstraction rather than via an oxidative mechanism involving oxygen species. The degradation followed pseudo-first-order kinetics. The rates pH-profile exhibited specific base catalysis. The stability of the drug was found to be dependent on pH, buffer concentration, buffer species (acetate, borate, phosphate) and temperature. The ionic strength did not affect the stability of the drug. The energy of activation according to Arrhenius plot was estimated to be 22.62 kcal mol(-1) at pH 12 and temperatures between 30 and 60 degrees C. The effect of simulated gastric and intestinal fluids on the drug stability was also investigated. Two major hydrolytic degradation products were separated and identified by IR, NMR and mass spectrometry and the degradative pathway suggested.  相似文献   

14.
Mometasone furoate (MF) is a synthetic glucocorticoid. There is little information available on the stability of MF and no degradation products have been unequivocally identified. Thus, the primary objective of this study was to characterize the degradation of MF, qualitatively and quantitatively. Stability of MF decreased with increasing pH (>4) and decreasing ionic strength in aqueous media. The chemical stability of MF in aqueous systems was significantly dependent on pH. MF appeared to be stable at pH < 4 but degraded to four products at higher pH. The degradation of MF in aqueous solutions follows pseudo-first-order kinetics and involved a series of parallel and consecutive reactions. The turnover of MF and its products appears to be catalyzed by the hydroxide ion. The pH dependence of these reactions should be considered, when formulating or extemporaneously compounding MF formulations. An optimal pH of stability was below pH 4. The changes in pH, however, do not appear to be the only factor of importance, since an increase in ionic strength and buffer concentration displayed a stabilizing effect on this glucocorticoid in the buffers tested. Trace metal ions are unlikely to be involved in degradation of MF in aqueous solution.  相似文献   

15.
Purpose. Eudragit RL and RS 30D are pseudolatexes frequently used in the coating of solid dosage forms. They are based on cationic copolymers stabilized with quaternary ammonium groups (poly(ethylacrylate-methylmethacrylate-trimethylammonioethyl methacrylate chloride). A pH-independent drug release is expected because of the quaternary nature of the cationic groups. The objective was to explain a distinct “pH-dependent” drug release in various buffer media with coated diltiazem beads. Methods. The diltiazem HC1 release from and water uptake of Eudragit RS/RL-coated beads was determined in various buffers of different buffer species, pH or concentration. Results. The drug release in the different buffer media was in the following order: pH 5.0 acetate > pH 3.5 formate > pH 7.4 phosphate buffer > 0.1M HC1). This “pH-dependent” drug release could be explained with an anion exchange process; the chloride counterions of the quaternary groups were exchanged with the anionic buffer species during the dissolution study. The water uptake of the coated beads correlated well with the drug release from the beads. Increasing the buffer strength (acetate buffer) first increased and then decreased the drug release, while increasing the ionic strength of different buffers with NaCl decreased the drug release and eliminated the observed buffer effects because of the excess of chloride ions. Conclusions. The anionic buffer species and not the pH had a significant effect on the hydration and hence on the drug release from beads coated with the cationic polymers, Eudragit RS and RL.  相似文献   

16.
Purpose. It has been suggested that fluorocarbon liquid may be a unique vehicle for the delivery of drugs directly to the acutely injured lung. A prodrug approach was used as a means of enhancing the solubility of a model drug (nicotinic acid) in the fluorocarbon. The solubility, the chemical stability of the putative prodrugs, and the sensitivity to enzymatic hydrolysis was investigated. Methods. The solubility of each nicotinic acid ester was determined in buffer as a function of pH and in perflubron. The octanol/buffer partition coefficient was determined at pH 7.4. The chemical stability of the putative prodrugs was determined as a function of pH, temperature, buffer content, and ionic strength. In addition, sensitivity of the esters to enzymatic degradation was evaluated. Results. Compared with nicotinic acid, the solubility in perflubron of the esters was significantly enhanced. In aqueous buffers, the esters exhibited pseudo-first order degradation kinetics, with both acid and base catalyzed loss. Studies of the fluorobutyl ester indicate quantitative loss of the putative prodrug and release of the parent nicotinic acid. Porcine esterase accelerated the loss of fluorobutyl ester by a factor of over 200 compared with chemical hydrolysis at pH 7.4. Conclusions. The properties of the fluorinated esters suggest that they may be suitable candidates for further testing as possible prodrugs of nicotinic acid based upon higher solubility in perflubron, rapid release of the parent drug after simple hydrolysis, and sensitivity to the presence of a model esterase enzyme.  相似文献   

17.
Studies have been initiated to evaluate the catalytic effect of monohydrogen phosphate ions on the oxidation of ferrous (Fe2+) to ferric (Fe3+) ions in an aqueous solution under atmospheric oxygen conditions. The reactions were performed with an initial concentration of 1 × 10?4 M ferrous sulfate in solutions containing varying concentrations of phosphate buffer (0.005–0.0175 M) over the pH range of 6.6–7.1. The final ionic strength of the solutions were adjusted to 0.1 M with sodium chloride and the temperature was kept constant at 25 ± 0.5 °C. The rates of oxidation reactions were measured by following the increase in UV absorbance due to the formation of ferric ion in solution. The reactions appeared to follow pseudo-first-order kinetics and were very prone to catalysis by monohydrogen phosphate at any given pH. H2PO4? seemed to have no effect on the reaction. HPOs2? was the sole catalytic species with a second-order rate constant of 116.74 M?1 · min?1. The buffer independent pH-rate profile showed a sigmoidal behavior with the pseudo-first-order rate constant increasing with increasing pH. The sigmoidal nature of the experimental pH-rate profile could possibly suggest a change in the reactivity of the oxidizing species which might follow complex kinetics. The effects of ionic strength and temperature on the reaction rates were also evaluated.  相似文献   

18.
目的 研究罗红霉素片的溶出行为,确定有区分力的溶出度检查方法。方法 考察罗红霉素在不同pH溶液中的稳定性及溶解度,测定不同处方自研制剂与参比制剂在4种不同溶出介质中的溶出度,绘制溶出曲线,用相似因子法进行拟合。结果 罗红霉素在pH6.8和7.4磷酸盐缓冲液中8 h内稳定性较好,自研制剂A与参比制剂在4种溶出介质中的溶出曲线均具有较好的相似性,自研制剂B与参比制剂在pH 6.8磷酸盐缓冲液中的溶出曲线不相似。结论 以pH 6.8磷酸盐缓冲液为溶出介质,采用浆法测定,75 r/min对本品溶出具有较好的区分力,可为本品的质量控制和一致性评价提供参考。  相似文献   

19.
The apparent rate of degradation of penicillin G potassium micellar solutions of 500,000 units/ml, a concentration commonly encountered in vials reconstituted for storage in the refrigerator, was investigated and compared to that of nonmicellar solutions of 8000 units/ml at 25 degrees, ionic strength of 1.1 M, and pH range from 5.0 to 9.5. In the micellar solutions the apparent rate of the H+-catalyzed degradation was increased twofold but that of water- and OH minus-catalyzed hydrolysis was decreased two- to three-fold. Consequently, the pH-rate profile of the micellar solutions was shifted to higher pH values and the pH of minimum degradation was found to be at 7.0 compared to 6.5 for the nonmicellar solution of the same ionic strength. Compared at their respective pH-rate profile minima, micellar penicillin G is 2.5 times as stable as the nonmicellar solution under the conditions of constant pH and ionic strength.  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